SlideShare a Scribd company logo
1 of 6
Download to read offline
H O S T E D B Y Contents lists available at ScienceDirect
Progress in Natural Science: Materials International
journal homepage: www.elsevier.com/locate/pnsmi
Original Research
CO2 activation on bimetallic CuNi nanoparticles☆
Natalie Austin, Brandon Butina, Giannis Mpourmpakis
⁎
Department of Chemical Engineering, University of Pittsburgh, Pittsburgh, PA 15261, United States
A R T I C L E I N F O
Keywords:
Density functional theory
Alloys
CO2 adsorption
Catalysis
Stability
A B S T R A C T
Density functional theory calculations have been performed to investigate the structural, electronic, and CO2
adsorption properties of 55-atom bimetallic CuNi nanoparticles (NPs) in core-shell and decorated architectures,
as well as of their monometallic counterparts. Our results revealed that with respect to the monometallic Cu55
and Ni55 parents, the formation of decorated Cu12Ni43 and core-shell Cu42Ni13 are energetically favorable. We
found that CO2 chemisorbs on monometallic Ni55, core-shell Cu13Ni42, and decorated Cu12Ni43 and Cu43Ni12,
whereas, it physisorbs on monometallic Cu55 and core-shell Cu42Ni13. The presence of surface Ni on the NPs is
key in strongly adsorbing and activating the CO2 molecule (linear to bent transition and elongation of C˭O
bonds). This activation occurs through a charge transfer from the NPs to the CO2 molecule, where the local
metal d-orbital density localization on surface Ni plays a pivotal role. This work identifies insightful structure-
property relationships for CO2 activation and highlights the importance of keeping a balance between NP
stability and CO2 adsorption behavior in designing catalytic bimetallic NPs that activate CO2.
1. Introduction
Novel methods for CO2 capture, storage, and utilization are of
marked interest for mitigating elevated levels of CO2 in the atmo-
sphere, which contributes to the greenhouse effect [1–3]. As an
example, industrial processes, utilizing transition metal (TM) catalysts,
use CO2 as a chemical feedstock to produce important products such as
formic acid and methanol [4]. Given the high thermodynamic stability
of CO2, strong reducing species such as H2, are necessary to activate
CO2 through hydrogenation reactions [5]. For example, industrial
synthesis of methanol, which takes place on a Cu/ZnO/Al2O3 catalyst,
requires a syngas feedstock containing CO, CO2 and H2 [4]. Studies on
alkali-doped metal catalysts [6–8], bare metal catalysts [6,9,10], and
metal oxide catalysts [7,11,12] have shown that CO2 can be activated in
the absence of reducing species (i.e. H2). The activation of CO2 on these
catalysts involves charge transfer from the metal system to the CO2
molecule, which results in a geometric change of CO2 from a linear to a
bent configuration and elongation of the C˭O bonds [5,13]. A recent
periodic Density Functional Theory (DFT) study on CO2 activation and
dissociation on fcc(100) surfaces of Fe, Co, Ni, and Cu, showed that Co
and Ni could favorably chemisorb CO2, transferring significant charge
to the molecule from the metal, exhibiting the lowest barriers for CO2
dissociation [14]. Despite considerable research on TM interactions
with CO2, kinetic limitations are still present in industrial processes
that utilize CO2 as a chemical feedstock [15]. Therefore, it is important
to identify catalysts and catalyst properties that activate CO2, in order
to improve the activity of CO2 conversion reactions.
Alloyed systems often exhibit very different properties from their
monometallic counterparts, which make them very attractive for
catalytic applications [16–18]. For example, experimental studies on
Ni/Cu(100) surfaces have shown enhanced catalytic activity for metha-
nol synthesis compared to the monometallic Cu(100) surface [19,20].
Additionally, a combined DFT and kinetic Monte Carlo study on the
effects of metal doping (metal = Au, Pd, Rh, Pt, and Ni) of Cu(111) on
methanol synthesis found that the Ni/Cu(111) surface showed the
highest rate for methanol production compared to the other alloyed
systems and the monometallic Cu(111) surface [21]. Temperature
programmed desorption (TPD) experiments on a Ni film grown on a
Cu(110) surface showed high rates of CO2 chemisorption compared to
the Cu(110) surface on which only physisorption was observed [22].
The ability of CuNi systems to adsorb CO2 have been theoretically
investigated on Ni-doped Cun (n=1–12) clusters [23]. The Ni-doped
Cun systems chemisorbed CO2 in a bent state, while the pure Cun
clusters only weakly interacted with CO2. In these clusters the strong
CO2 adsorption was observed when CO2 was in direct contact with the
Ni atom of the cluster. CuNi nanoparticle (NP) interactions with CO2
have also been studied on 55-atom NPs with Ni-doped and core-shell
CuNi compositions [24]. In the doped system a single Ni atom was
located in the core, surrounded by 54 Cu atoms and in the core-shell
system 13 Ni atoms were located in the core and 42 Cu atoms were on
http://dx.doi.org/10.1016/j.pnsc.2016.08.007
Received 31 July 2016; Accepted 16 August 2016
Peer review under responsibility of Chinese Materials Research Society.
⁎
Corresponding author.
E-mail address: gmpourmp@pitt.edu (G. Mpourmpakis).
Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx
1002-0071/ © 2016 Chinese Materials Research Society. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by/4.0/).
Available online xxxx
Please cite this article as: Austin, N., Progress in Natural Science: Materials International (2016), http://dx.doi.org/10.1016/j.pnsc.2016.08.007
the shell. Only weak CO2 adsorption was observed on these systems as
further indicated by the C˭O bond lengths and O˭C˭O bond angle of
adsorbed CO2 (~1.172 Å and ~180°) remaining similar to that of gas
phase CO2 (1.162 Å and 180°).
As we have pointed out, it appears that the presence of Ni on the NP
surface can facilitate strong CO2 adsorption on the catalyst. In addition,
very recently the active site for CO2 hydrogenation to methanol on the
industrial catalyst Cu/ZnO/Al2O3, was identified to be a Cu stepped
surface decorated with Zn atoms [4]. According to DFT calculations in
this study, the incorporation of Zn to the Cu(211) surface increased the
adsorption strength of reaction intermediates and decreased the
barriers for methanol synthesis compared to the monometallic
Cu(211) surface. Since the presence of Zn on a catalyst surface
enhances CO2 hydrogenation to methanol [4] and the presence of Ni
on the surface of small clusters facilitates CO2 adsorption [23], in this
study we investigate CO2 adsorption and activation on CuNi bimetallic
NPs, accounting for Ni and Cu surface decorations, core-shell bime-
tallic configurations, as well as, on their monometallic counterparts.
We provide a thorough analysis on the structural, thermodynamic, and
electronic properties of these NPs that could potentially lead to
identifying stable and active catalytic systems for CO2 conversion to
chemicals.
2. Computational Details
We used the BP86 functional [25,26], the resolution of the identity
(RI) approximation [27,28], and the def2-SV(P) [29,30] basis set as
implemented in the TURBOMOLE 7.02 computational program pack-
age [31]. Dispersion corrections have been taken into account in the
calculations using the D3 method [32,33]. The total electronic energies
of icosahedral 55-atom monometallic and bimetallic NPs composed of
Cu and Ni were determined by geometry optimizations. Specifically,
these structures include monometallic (Cu55 and Ni55), decorated
(Cu43Ni12 and Cu12Ni43), and core-shell (Cu42Ni13 and Cu13Ni42)
combinations of Cu and Ni atoms. Multiple spin states were considered
in our calculations and the lowest in energy were selected for further
analysis. Furthermore, all the optimized structures were verified as
minima with frequency calculations (absence of any imaginary modes).
The cohesive energy (binding energy per metal atom, BE/n) [34], a
descriptor for average bond strength of the NPs, was calculated using
Eq. 1:
BE
n
=
E -(55-x)E - (x)E
55
Cu Ni Cu Ni55-x x
(1)
where x is the number of Ni atoms in the NP (x =0, 12, 13, 42, 43, and
55) and ECu55-x,Nix, ECu, and ENi are the total energies of the Cu55−xNix
NPs, Cu atom, and Ni atom, respectively. Using Eq. (2) we calculated
the excess energy (Eexc) [35,36], a descriptor for the stability of the
bimetallic NPs relative to their monometallic counterparts:
E =
E - E - E
55
exc
Cu Ni
(55-x)
55 Cu
x
55 Ni55-x x 55 55
(2)
where ECu55 and ENi55 are the total energies of the Cu55 and Ni55 NPs.
The Eexc indicates if the formation of the bimetallic systems will be
favorable (negative total energy) or unfavorable. Eq. (3) was used to
calculate CO2 adsorption (binding energy, BE) on the metal NPs
BE=E -E -ENP_CO NP CO2 2 (3)
where ENP_CO2, ENP, and ECO2 are the total energies of CO2 interacting
with the NP, the isolated NP, and isolated CO2, respectively.
Adsorption was assessed with CO2 oriented towards one of the
equivalent corner sites (coordination number 6, CN6) on the surface
of the NPs, in two different adsorption configurations, horizontal and
vertical to the NP surface. In the case of Cu12Ni43, CO2 adsorption was
also investigated with CO2 oriented towards the edge site (CN8) of the
NP where Ni atoms are located (vide-infra analysis). During CO2
optimization, the coordinates of the NPs were kept frozen at their
optimized positions and the CO2 molecule was allowed to relax. The
calculated adsorption states were further verified as minima with
frequency calculations (absence of imaginary modes on CO2). Natural
Fig. 1. Optimized geometries of 55-atom NPs: monometallic (a) Cu55 and (b) Ni55, decorated (c) Cu43Ni12 and (d) Cu12Ni43 NP with 12 heteroatoms located at the corner (CN6) sites of
the NP, and core-shell (e) Cu13Ni42 and (f) Cu42Ni13 with 13 metal atoms located in the core and 42 heteroatoms in the shell. The equivalent CN6 and CN8 sites on the NPs are illustrated
in (a). The Cu atoms are colored brown and the Ni blue.
N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx
2
bond orbital (NBO) analysis was used to calculate charge distribution
on all the systems. Molecular orbital plots were visualized using
TmoleX [37], a graphical user interface for TURBOMOLE.
3. Results and discussion
3.1. Structural and electronic properties of Cu55−xNix NPs
The lowest energy structures of the six NPs studied in this work are
shown in Fig. 1. The icosahedral shape, which is the lowest energy
structure for monometallic Cu55 and Ni55[38], was also maintained for
the bimetallic systems. The conservation of the icosahedral geometry
can be attributed to the similar atomic radius of Cu (1.28 Å) [39] and
Ni (1.25 Å) [39] atoms. The calculated BE/n and Eexc of the systems are
shown in Table 1. The BE/n trend from the largest (most negative
value) to lowest value is as follows: Ni55 > Cu12Ni43 > Cu13Ni42 >
Cu42Ni13 > Cu43Ni12 > Cu55. Thus, we find that as the Ni fraction in
Cu55−xNix increases, the BE/n also increases. This observation agrees
with the melting points of the metals, with Ni (1728 K) [39] having
higher melting point than Cu (1358 K) [39], and as a result, the average
bond strength of their alloys shows larger values with higher Ni
content. Furthermore we found that the BE/n trend of the mono-
metallics followed that of experimental bulk (Cu 3.49 eV/atom and Ni
4.44 eV/atom) [39], where in this case the more positive BE/n
represents the more stable system.
The Eexc trend describes the stability of the bimetallic NPs relative
to the monometallic parents. The Eexc trend from most stable to least
stable is as follows: Cu42Ni13 > Cu12Ni43 > Cu43Ni12 > Cu13Ni42. Overall
we found that the core-shell structure, Cu42Ni13, is the most energe-
tically favorable formation for the bimetallic NPs. In theoretical work
by Yang et al. the icosahedral Cu42Ni13 NP was also found to be
energetically favorable. [24] The Eexc trends can be attributed to the
surface energy (Cu 1170 ergs·cm−2
and Ni 2240 ergs·cm−2
) [40] and
bulk cohesive energy of monometallic Cu and Ni. The surface and
cohesive energy values show that, in general, Cu prefers to reside on the
surface and Ni prefers to be in the core of the NP. This core-shell
preference has been shown to be present in nanoscale systems in the
recent work by Wang et al. [41], calculating segregation energies of
doped transition metal systems. In particular it was found that the Ni
atom preferred to reside in the core than on the surface of a 55-atom
cuboctohedral Cu54Ni. Due to the energetic preference of Cu for the
surface and Ni for the core of the NPs, we find that the core-shell
Cu42Ni13 is more energetically favorable than the corresponding
Cu13Ni42 NP, and the decorated Cu12Ni43 is more favorable than the
corresponding Cu43Ni12 NP. Overall the results show that the most
energetically favorable NPs have Cu (lower cohesive and surface
energy) on the lowest coordinated, surface sites, and Ni (higher
cohesive and surface energy) on the highest coordinated, bulk sites of
the NP.
Table 1
Calculated BE/n and Eexc of the Cu55-xNix NPs. The negative values indicate exothermi-
city.
BE/n (eV/atom) Eexc (eV/atom)
Cu55 −3.266 –
Ni55 −4.155 –
Cu43Ni12 −3.413 0.0560
Cu12Ni43 −3.980 −0.0218
Cu42Ni13 −3.535 −0.0692
Cu13Ni42 −3.849 0.113
1 2
1
2
1
2
1 2
1 2
1
2
Fig. 2. CO2 adsorption on the (a, b) monometallic, (c, d) decorated, and (e, f) core-shell Cu-Ni NPs. The color code is as depicted in Fig. 1 with the addition of CO2 (C colored gray and
oxygen colored red).
N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx
3
3.2. CO2 adsorption on the Cu55−xNix NPs
Previous theoretical work on adsorption of small molecules, such as
CO, on the surface of NPs showed that the stronger adsorption was
observed on surface NP sites exhibiting low CNs (e.g. corners, edges)
[42]. In turn, in the CO2 adsorption studies [5,6,14], strong adsorption
of the CO2 molecule on the metal surface results to its activation. Thus,
we primarily investigated the CO2 adsorption on the corner (CN6) site
(the lowest coordinated site in our systems) of the Cu55−xNix NPs. As
shown in Fig. 1(a), since all the CN6 sites are equivalent (same with the
CN8 edge sites) due to the Ih symmetry of the NPs, we studied the CO2
adsorption on one of these sites. The lowest energy structures for CO2
adsorption are shown in Fig. 2. We calculated the adsorption energy for
two orientations of the CO2 molecule: (i) horizontal CO2, where the C
atom of CO2 was interacting with the corner site and (ii) vertical CO2,
where an O atom of CO2 was interacting with the corner site. The
horizontal orientation was found to be the most preferred adsorption
configuration of CO2 on all the metal NPs in our study. Consistent with
literature [5,9,12–14], we found that the systems that activate CO2
always show strong adsorption and charge transfer from the metal NP
to CO2, which results in a linear to bent transition of the CO2 molecule,
and elongation of the C˭O bonds. Specifically from Table 2 we found
strong CO2 adsorption on Ni55, Cu43Ni12, Cu12Ni43, and Cu13Ni42 and
weak adsorption on Cu55 and Cu42Ni13. In addition as shown in Table 2
the systems with strong CO2 adsorption had more than −0.6|e|
transferred to the CO2 molecule while in the weakly adsorbed systems
there was no significant charge transfer. The geometric properties of
gas phase CO2 and CO2 interacting with the metal NPs are also shown
in Table 2. Compared to gas phase (non-interacting) CO2, the O1-C-O2
angle of the strongly adsorbed CO2 decreased ( < 150°) and the C˭O
bonds elongated ( > 1.2 Å) which indicates the activation of the CO2
molecule. In the weakly adsorbed systems the bond angles and bond
distances of CO2 remained similar to gas phase CO2. The deviations of
average bond lengths (black squares) and angles (blue circles) between
adsorbed and gas phase CO2 as a function of CO2 BE are illustrated in
Fig. 3. It is clear that both the CO2 bond distances and angles are
affected in a similar way with the CO2 BE. To demonstrate this
interdependence between the CO2 angle and bond distance deviation
(from CO2 gas) we plotted these geometric properties in the inset of
Fig. 3 and we observe a perfect linear trend. Overall, Fig. 3 shows that
the bending of the CO2 molecule and the elongation of its bonds show
the same behavior and they are enhanced (increased activation) with
stronger adsorption on the NP surface. In Fig. 4(a) we demonstrate the
CO2 BE as a function of the total charge transferred to the CO2
molecule. We notice that bimetallic systems with Ni atoms being on the
surface of the NP transfer significant charge to CO2. In turn, this results
in strong CO2 adsorption, while for the Cu42Ni13 system, where Ni is at
the core of the NP and inaccessible to CO2, significant charge transfer is
not observed, resulting to weak CO2 adsorption. It should be noticed
that so far, our discussion on the activation of the CO2 molecule is
entirely focused on the structural and electronic observations made on
the CO2 molecule itself.
From a catalyst design perspective, we need to identify a property of
the metal NPs that could correlate with the observed activation. As a
result, we made an effort to rationalize the CO2 adsorption behavior
using the d-band center (dC) model by Hammer and Norskov [43,44].
We calculated the local dC on a single metal site of CO2 adsorption (Cu
or Ni) on the Cu55−xNix NPs. In Fig. 4(b), the CO2 adsorption values
are plotted as a function of the local site dC of the NPs. The observed
linear trend indicates that there is an increase in CO2 adsorption with
decrease in the dC, shifting towards the energy level of the CO2 Lowest
Unoccupied Molecular Orbital (LUMO), which is located at −0.35 eV
(blue dashed line). The Cu43Ni12 showed the highest local dC and
strongest BE, while the Cu55 showed the lowest dC and the weakest BE.
Notice that the Cu43Ni12 positions its local-site dC at the energy level of
CO2 LUMO, showing the strongest adsorption. It is important to note
that for Cu12Ni43 we did not observe strong adsorption when CO2
interacted directly with the corner Cu sites of the NP. Even the
presence of Ni atoms in the neighboring positions (edge sites), did
not enhance CO2 adsorption on the corner site compared to the Cu55
system. However, since surface Ni is responsible for enhancing the CO2
adsorption (compare CO2 adsorption on Cu55 vs. Cu43Ni12), we
calculated the adsorption energy of CO2 on an edge site (CN8) of the
Cu12Ni43, where the Ni atoms are located. In this case we observe a
strong CO2 adsorption on Cu12Ni43 which was 0.4 eV lower in energy
than the case where the CO2 interacted with the Cu atoms of the NP.
This shows that the presence of Ni atoms on the surface of the NP
significantly enhances CO2 adsorption.
To further rationalize the effect of the metal's d-orbital density and
the charge transfer on the CO2 activation, we plotted the Highest
Occupied Molecular Orbitals (HOMO) of Cu55, Ni55, Cu43Ni12 and
Cu12Ni43 NPs as shown in Fig. 5. These are the monometallic parents
(Cu55 and Ni55), the bimetallic showing the strongest CO2 adsorption
(Cu43Ni12) among the NPs studied, and the bimetallic NP that shows
strong CO2 adsorption and favorable Eexc (Cu12Ni43). It has been
shown that charge transfer from the metal to CO2 occurs when the d-
orbitals of the metal NP significantly interacts with the 2Πu antibond-
ing orbital (LUMO) of CO2 [13]. In Fig. 5 we assessed the presence of
d-orbital character in the HOMO orbitals of the aforementioned NPs.
Figs. 5(a) and (b) illustrate the HOMO orbital distribution on mono-
metallic Cu55 and Ni55 respectively. From the shape of the orbitals we
observe primarily s-orbital localization on Cu55, while there is mainly
d-orbital localization on Ni55. By quantifying the orbital character
through the atomic orbital coefficients of the HOMO, we found that
Table 2
CO2 binding energies on the NPs, total charge (NBO) transferred to CO2, and geometric
properties of gas phase CO2 (non-interacting) and CO2 interacting with the NPs. Binding
energies are in eVs. Bond lengths of C-O1 and C-O2 are in angstroms, Å. Bond angle of
O1-C-O2 are in degrees, °.
BE (eV) |e| C-O1 (Å) C-O2 (Å) O1-C-O2 (°)
CO2 (gas) – – 1.175 1.175 180.000
Cu55 −0.368 0.00825 1.169 1.183 179.406
Ni55 −0.665 −0.658 1.243 1.238 141.610
Cu43Ni12 −0.919 −0.604 1.245 1.230 143.465
Cu12 Ni43 −0.803 −0.810 1.216 1.309 132.070
Cu42Ni13 −0.391 0.00207 1.183 1.170 178.989
Cu13 Ni42 −0.891 −0.652 1.231 1.255 140.476
Fig. 3. Deviation of adsorbed CO2 average C˭O bond length (left ordinate) and O˭C˭O
bond angle (right ordinate) from gas phase CO2 as functions of CO2 BE. The inset figure
at the bottom left shows a linear relationship between average C˭O bond length and
O˭C˭O bond angle.
N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx
4
there is a much greater fraction of d-orbital character in the HOMO of
Ni55, which shows strong binding to CO2 than in Cu55, which weakly
adsorbs CO2. The bimetallic systems Cu43Ni12 and Cu12Ni43 show high
fractions d-orbital character with Cu12Ni43 having the highest.
Cu43Ni12 has a lower fraction of d-orbital character than Cu12Ni43
because of the lower fraction of Ni atoms in Cu43Ni12 than in Cu12Ni43.
It should also be noticed that the d-orbitals are localized on surface
CN6 atoms of the Cu43Ni12, whereas, on the CN8 of the Cu12Ni43. In
other words, the surface sites where the Ni atoms are located show
strong CO2 adsorption. Overall we demonstrate that adding Ni on the
surface of Cu NPs increases the presence of d-orbital character, which
in turn, results in a favorable interaction of the NP with CO2, and
subsequent activation of the CO2 molecule. Although we found that
Cu43Ni12 binds CO2 the strongest, the most promising NP from our
study is the Cu12Ni43 because in addition to the strong CO2 adsorption
(BE =−0.8 eV) and activation, it shows favorable energetics for its
synthesis (Eexc=−0.02 eV/atom). This study highlights that in bimetal-
lic catalyst design it is important to achieve a balance between catalyst
stability (the most stable CuNi NPs prefer Cu to be on the surface) and
interaction strength of the catalyst with adsorbates (CO2 strongly
adsorbs and is being activated on surface Ni). Although computational
studies like this one identify bimetallic NPs that are stable and
promising for CO2 activation, there is a synthetic challenge in forming
nanostructures with (in-silico) predefined architecture. Recent experi-
mental advances on the controlled synthesis of bimetallic NPs with
atomic precision can pave the way towards achieving this goal [45,46].
4. Conclusions
In summary, we performed a DFT investigation on the structural,
electronic and CO2 adsorption properties of Cu55−xNix (x=0, 12, 13, 42,
43, and 55) NPs with a monometallic, core-shell, and decorated
distribution of Cu and Ni atoms. We found that the BE/n of the
bimetallic systems was a linear combination of the BE/n of the
monometallic systems. We also calculated the excess energy (Eexc) of
the bimetallic NPs with respect to the monometallic NPs and showed
that the formation of decorated Cu12Ni43 and core-shell Cu42Ni13 NPs
were energetically favorable, while the formation of core-shell Cu13Ni42
and decorated Cu43Ni12 were less favorable. These trends rationalize
the preference of Cu to be located at the surface of the NPs rather than
Ni. CO2 adsorption calculations revealed weak interaction (physisorp-
tion) with the monometallic Cu55 and core-shell Cu42Ni13, while the
monometallic Ni55, decorated Cu12Ni43 and Cu43Ni12, and core-shell
Cu13Ni42 chemisorbed CO2. In the chemisorbed cases we found strong
adsorption of CO2 on the corner sites of all NPs except Cu12Ni43 where
strong CO2 adsorption was found on the edge sites. The sites of strong
adsorption on the NPs were always surface sites which were occupied
by Ni atoms. Thus, the location of Ni on the NP plays an important role
in the resulting adsorption behavior. The chemisorption behavior on
the NPs was attributed to charge transferred from the metal NPs to CO2
which led to the activation of the molecule. Additionally, we calculated
the local-site d-band center (dC) and we found a linear relationship
between the dC and CO2 adsorption energy. The sites of strong
adsorption localize HOMO orbitals with increased d-character.
Fig. 4. CO2 BE as a function of (a) total charge on CO2 and (b) local dC of the Cu55−xNix NPs. The dashed black lines in both figures serve as a guide to the eye. In (b) the vertical blue line
represents the LUMO orbital energy of the CO2 molecule.
Fig. 5. Visual representation of the HOMO orbitals and fractional distribution of the HOMO orbital character of the (a) Cu55, (b) Ni55, (c) Cu43Ni12, (c) Cu12Ni43 NPs.
N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx
5
Overall this study demonstrates that the presence of surface Ni on CuNi
bimetallic NPs can significantly enhance CO2 adsorption, resulting in
the activation of the CO2 molecule. Furthermore, among the different
nanostructures in this study we identified the Cu12Ni43, which can be
potentially experimentally synthesized and activate CO2 for dissocia-
tion and hydrogenation reactions due to its exothermic Eexc and strong
adsorption behavior towards CO2, respectively.
Acknowledgements
This material is based upon work supported by the National Science
Foundation Graduate Research Fellowship Program to N.A. under
Grant No. (1247842). The authors would like to acknowledge compu-
tational support from the Center for Simulation and Modeling (SAM)
and start-up funding from the University of Pittsburgh.
References
[1] W. Wang, S.P. Wang, X.B. Ma, J.L. Gong, Chem Soc. Rev. 40 (2011) 3703–3727.
[2] A. Goeppert, M. Czaun, G.K.S. Prakash, G.A. Olah, Energy Environ. Sci. 5 (2012)
7833–7853.
[3] G. Centi, E.A. Quadrelli, S. Perathoner, Energy Environ. Sci. 6 (2013) 1711–1731.
[4] M. Behrens, F. Studt, I. Kasatkin, S. Kuehl, M. Haevecker, F. Abild-Pedersen, et al.,
Science 336 (2012) 893–897.
[5] H.J. Freund, M.W. Roberts, Surf. Sci. Rep. 25 (1996) 225–273.
[6] F. Solymosi, J. Mol. Catal. 65 (1991) 337–358.
[7] D. Ochs, M. Brause, B. Braun, W. Maus-Friedrichs, V. Kempter, Surf. Sci. 397
(1998) 101–107.
[8] F. Solymosi, G. Klivenyi, Surf. Sci. 315 (1994) 255–268.
[9] J. Ko, B.K. Kim, J.W. Han, J. Phys. Chem C. 120 (2016) 3438–3447.
[10] S.G. Wang, D.B. Cao, Y.W. Li, J.G. Wang, H.J. Jiao, J. Phys. Chem B. 109 (2005)
18956–18963.
[11] Y. Wang, R. Kovacik, B. Meyer, K. Kotsis, D. Stodt, V. Staemmler, et al., Angew.
Chem.-Int. Ed. 46 (2007) 5624–5627.
[12] K.R. Hahn, M. Iannuzzi, A.P. Seitsonen, J. Hutter, J. Phys. Chem C. 117 (2013)
1701–1711.
[13] S.G. Wang, X.Y. Liao, D.B. Cao, C.F. Huo, Y.W. Li, J.G. Wang, et al., J. Phys. Chem
C. 111 (2007) 16934–16940.
[14] C. Liu, T.R. Cundari, A.K. Wilson, J. Phys. Chem C. 116 (2012) 5681–5688.
[15] X.M. Liu, G.Q. Lu, Z.F. Yan, Beltramini, J. Ind. ENG Chem Res. 42 (2003)
6518–6530.
[16] F. Studt, I. Sharafutdinov, F. Abild-Pedersen, C.F. Elkjaer, J.S. Hummelshoj,
S. Dahl, et al., Nat. Chem 6 (2014) 320–324.
[17] M.S. Chen, D. Kumar, C.W. Yi, D.W. Goodman, Science 310 (2005) 291–293.
[18] A. Sandoval, A. Aguilar, C. Louis, A. Traverse, R. Zanella, J. Catal. 281 (2011)
40–49.
[19] J. Nerlov, I. Chorkendorff, J. Catal. 181 (1999) 271–279.
[20] J. Nerlov, S. Sckerl, J. Wambach, I. Chorkendorff, Appl Catal. a-Gen. 191 (2000)
97–109.
[21] Y.X. Yang, M.G. White, P. Liu, J. Phys. Chem C. 116 (2012) 248–256.
[22] E. Vesselli, E. Monachino, M. Rizzi, S. Furlan, X.M. Duan, C. Dri, et al., Acs Catal. 3
(2013) 1555–1559.
[23] S.L. Han, X.L. Xue, X.C. Nie, H. Zhai, F. Wang, Q. Sun, et al., Phys. Lett. A 374
(2010) 4324–4330.
[24] Y. Yang, D. Cheng, J. Phys. Chem C. 118 (2014) 250–258.
[25] A.D. Becke, Phys. Rev. A 38 (1988) 3098–3100.
[26] J.P. Perdew, Phys. Rev. B 33 (1986) 8822–8824.
[27] K. Eichkorn, F. Weigend, O. Treutler, R. Ahlrichs, Theor. Chem Acc. 97 (1997)
119–124.
[28] K. Eichkorn, O. Treutler, H. Ohm, M. Haser, R. Ahlrichs, Chem Phys. Lett. 240
(1995) 283–289.
[29] A. Schafer, H. Horn, R. Ahlrichs, J. Chem Phys. 97 (1992) 2571–2577.
[30] F. Weigend, M. Haser, H. Patzelt, R. Ahlrichs, Chem Phys. Lett. 294 (1998)
143–152.
[31] R. Ahlrichs, M. Bar, M. Haser, H. Horn, C. Kolmel, Chem Phys. Lett. 162 (1989)
165–169.
[32] S.J. Grimme, Comput Chem 25 (2004) 1463–1473.
[33] S.J. Grimme, Comput Chem 27 (2006) 1787–1799.
[34] N. Austin, G. Mpourmpakis, J. Phys. Chem C. 118 (2014) 18521–18528.
[35] R. Ferrando, J. Jellinek, R.L. Johnston, Chem Rev. 108 (2008) 845–910.
[36] M.J. Piotrowski, P. Piquini, J.L.F. Da Silva, J. Phys. Chem C. 116 (2012)
18432–18439.
[37] C. Steffen, K. Thomas, U. Huniar, A. Hellweg, O. Rubner, A. Schroer, J. Comput
Chem. 31 (2010) 2967–2970.
[38] J.P.K. Doye, D.J. Wales, New J. Chem. 22 (1998) 733–744.
[39] C. Kittel, Introduction to Solid State Physics, 7 ed, Wiley, New York, 1996.
[40] Y.-N. Wen, H.-M. Zhang, Solid State Commun. 144 (2007) 163–167.
[41] L.-L. Wang, D.D. Johnson, J. Am. Chem Soc. 131 (2009) 14023–14029.
[42] G. Mpourmpakis, A.N. Andriotis, D.G. Vlachos, Nano Lett. 10 (2010) 1041–1045.
[43] Hammer B, Norskov JK. Theoretical surface science and catalysis – calculations
and concepts. In: Gates BC, Knozinger H, editors. Advances in Catalysis, Vol 45:
Impact of Surface Science on Catalysis, 2000, pp. 71–129.
[44] B. Hammer, J.K. Norskov, Surf. Sci. 343 (1995) 211–220.
[45] R. Jin, K. Nobusada, Nano Res. 7 (2014) 285–300.
[46] H. Qian, Li.G. Jiang D-e, C. Gayathri, A. Das, R.R. Gil, et al., J. Am. Chem Soc. 134
(2012) 16159–16162.
N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx
6

More Related Content

What's hot

Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Pawan Kumar
 
Balucan and Steel_2015_A regenerable precipitant-solvent system for CO2 mitig...
Balucan and Steel_2015_A regenerable precipitant-solvent system for CO2 mitig...Balucan and Steel_2015_A regenerable precipitant-solvent system for CO2 mitig...
Balucan and Steel_2015_A regenerable precipitant-solvent system for CO2 mitig...Reydick D Balucan
 
Bicrystalline Titania Photocatalyst for Reduction of CO2 to Solar Fuels
Bicrystalline Titania Photocatalyst for Reduction of CO2 to Solar FuelsBicrystalline Titania Photocatalyst for Reduction of CO2 to Solar Fuels
Bicrystalline Titania Photocatalyst for Reduction of CO2 to Solar FuelsA'Lester Allen
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Pawan Kumar
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Pawan Kumar
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Pawan Kumar
 
Nanostructured composite materials for CO2 activation
Nanostructured composite materials for CO2 activationNanostructured composite materials for CO2 activation
Nanostructured composite materials for CO2 activationPawan Kumar
 
Investigations on Physical Properties of Sulfur Based Composite Cathodes in L...
Investigations on Physical Properties of Sulfur Based Composite Cathodes in L...Investigations on Physical Properties of Sulfur Based Composite Cathodes in L...
Investigations on Physical Properties of Sulfur Based Composite Cathodes in L...IRJET Journal
 
Sdarticle (2)
Sdarticle (2)Sdarticle (2)
Sdarticle (2)52900339
 
Core–shell structured reduced graphene oxide wrapped magneticallyseparable rG...
Core–shell structured reduced graphene oxide wrapped magneticallyseparable rG...Core–shell structured reduced graphene oxide wrapped magneticallyseparable rG...
Core–shell structured reduced graphene oxide wrapped magneticallyseparable rG...Pawan Kumar
 
Simulation of Flash Pyrolysis of Maiganga Coal Using Modified Straight First ...
Simulation of Flash Pyrolysis of Maiganga Coal Using Modified Straight First ...Simulation of Flash Pyrolysis of Maiganga Coal Using Modified Straight First ...
Simulation of Flash Pyrolysis of Maiganga Coal Using Modified Straight First ...Premier Publishers
 
Studying and Comparing Sensing Capability of Single Walled Carbon Nanotubes f...
Studying and Comparing Sensing Capability of Single Walled Carbon Nanotubes f...Studying and Comparing Sensing Capability of Single Walled Carbon Nanotubes f...
Studying and Comparing Sensing Capability of Single Walled Carbon Nanotubes f...IJERA Editor
 
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...Ary Assuncao
 
Development of novel catalytic systems for photoreduction of CO2 to fuel and ...
Development of novel catalytic systems for photoreduction of CO2 to fuel and ...Development of novel catalytic systems for photoreduction of CO2 to fuel and ...
Development of novel catalytic systems for photoreduction of CO2 to fuel and ...Pawan Kumar
 
A bridged ruthenium dimer (ru–ru) for photoreduction of co2 under visible lig...
A bridged ruthenium dimer (ru–ru) for photoreduction of co2 under visible lig...A bridged ruthenium dimer (ru–ru) for photoreduction of co2 under visible lig...
A bridged ruthenium dimer (ru–ru) for photoreduction of co2 under visible lig...Pawan Kumar
 

What's hot (19)

Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
 
Balucan and Steel_2015_A regenerable precipitant-solvent system for CO2 mitig...
Balucan and Steel_2015_A regenerable precipitant-solvent system for CO2 mitig...Balucan and Steel_2015_A regenerable precipitant-solvent system for CO2 mitig...
Balucan and Steel_2015_A regenerable precipitant-solvent system for CO2 mitig...
 
Bicrystalline Titania Photocatalyst for Reduction of CO2 to Solar Fuels
Bicrystalline Titania Photocatalyst for Reduction of CO2 to Solar FuelsBicrystalline Titania Photocatalyst for Reduction of CO2 to Solar Fuels
Bicrystalline Titania Photocatalyst for Reduction of CO2 to Solar Fuels
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
 
66
6666
66
 
Nanostructured composite materials for CO2 activation
Nanostructured composite materials for CO2 activationNanostructured composite materials for CO2 activation
Nanostructured composite materials for CO2 activation
 
CO2 Presentation
CO2 PresentationCO2 Presentation
CO2 Presentation
 
Investigations on Physical Properties of Sulfur Based Composite Cathodes in L...
Investigations on Physical Properties of Sulfur Based Composite Cathodes in L...Investigations on Physical Properties of Sulfur Based Composite Cathodes in L...
Investigations on Physical Properties of Sulfur Based Composite Cathodes in L...
 
Sdarticle (2)
Sdarticle (2)Sdarticle (2)
Sdarticle (2)
 
Core–shell structured reduced graphene oxide wrapped magneticallyseparable rG...
Core–shell structured reduced graphene oxide wrapped magneticallyseparable rG...Core–shell structured reduced graphene oxide wrapped magneticallyseparable rG...
Core–shell structured reduced graphene oxide wrapped magneticallyseparable rG...
 
ncomms13869
ncomms13869ncomms13869
ncomms13869
 
Simulation of Flash Pyrolysis of Maiganga Coal Using Modified Straight First ...
Simulation of Flash Pyrolysis of Maiganga Coal Using Modified Straight First ...Simulation of Flash Pyrolysis of Maiganga Coal Using Modified Straight First ...
Simulation of Flash Pyrolysis of Maiganga Coal Using Modified Straight First ...
 
Studying and Comparing Sensing Capability of Single Walled Carbon Nanotubes f...
Studying and Comparing Sensing Capability of Single Walled Carbon Nanotubes f...Studying and Comparing Sensing Capability of Single Walled Carbon Nanotubes f...
Studying and Comparing Sensing Capability of Single Walled Carbon Nanotubes f...
 
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
A closed loop ammonium salt system for recovery of high-purity lead tetroxide...
 
Development of novel catalytic systems for photoreduction of CO2 to fuel and ...
Development of novel catalytic systems for photoreduction of CO2 to fuel and ...Development of novel catalytic systems for photoreduction of CO2 to fuel and ...
Development of novel catalytic systems for photoreduction of CO2 to fuel and ...
 
A bridged ruthenium dimer (ru–ru) for photoreduction of co2 under visible lig...
A bridged ruthenium dimer (ru–ru) for photoreduction of co2 under visible lig...A bridged ruthenium dimer (ru–ru) for photoreduction of co2 under visible lig...
A bridged ruthenium dimer (ru–ru) for photoreduction of co2 under visible lig...
 
silver
silversilver
silver
 

Similar to CO2_activation_on_bimetallic_CuNi_nanoparticles

Structural aspect on carbon dioxide capture in nanotubes
Structural aspect on carbon dioxide capture in nanotubesStructural aspect on carbon dioxide capture in nanotubes
Structural aspect on carbon dioxide capture in nanotubesIJRES Journal
 
Simulation, modeling and fabrication of the bio-inspired flexible tactile sen...
Simulation, modeling and fabrication of the bio-inspired flexible tactile sen...Simulation, modeling and fabrication of the bio-inspired flexible tactile sen...
Simulation, modeling and fabrication of the bio-inspired flexible tactile sen...dc230400176dhu
 
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...Pawan Kumar
 
acs%2Ejpcc%2E5b10334
acs%2Ejpcc%2E5b10334acs%2Ejpcc%2E5b10334
acs%2Ejpcc%2E5b10334Victor Mao
 
Theoretical study of two dimensional Nano sheet for gas sensing application
Theoretical study of two dimensional Nano sheet for gas sensing  applicationTheoretical study of two dimensional Nano sheet for gas sensing  application
Theoretical study of two dimensional Nano sheet for gas sensing applicationvivatechijri
 
F041033947
F041033947F041033947
F041033947IOSR-JEN
 
Catal Commun 45 (2014) 153-Author
Catal Commun 45 (2014) 153-AuthorCatal Commun 45 (2014) 153-Author
Catal Commun 45 (2014) 153-AuthorHo Huu Trung
 
Wamg -graphene tubes_Nanoscale
Wamg -graphene tubes_NanoscaleWamg -graphene tubes_Nanoscale
Wamg -graphene tubes_NanoscaleHaiyang Sheng
 
Characterization of Structural and Surface Properties of Nanocrystalline TiO2...
Characterization of Structural and Surface Properties of Nanocrystalline TiO2...Characterization of Structural and Surface Properties of Nanocrystalline TiO2...
Characterization of Structural and Surface Properties of Nanocrystalline TiO2...Shingo Watanabe (渡邊真悟)
 
Böhler et al. - 2014 - Recent advances in the study of the UO2–PuO2 phase dia...
Böhler et al. - 2014 - Recent advances in the study of the UO2–PuO2 phase dia...Böhler et al. - 2014 - Recent advances in the study of the UO2–PuO2 phase dia...
Böhler et al. - 2014 - Recent advances in the study of the UO2–PuO2 phase dia...shyam933392
 
A first principles study into the properties and activities of rare-earth and...
A first principles study into the properties and activities of rare-earth and...A first principles study into the properties and activities of rare-earth and...
A first principles study into the properties and activities of rare-earth and...Hnakey Lora
 
H2 OXIDATION AT Pd100 A FIRST-PRINCIPLES CONSTRAINED THERMODYNAMICS STUDY
H2 OXIDATION AT Pd100 A FIRST-PRINCIPLES CONSTRAINED THERMODYNAMICS STUDYH2 OXIDATION AT Pd100 A FIRST-PRINCIPLES CONSTRAINED THERMODYNAMICS STUDY
H2 OXIDATION AT Pd100 A FIRST-PRINCIPLES CONSTRAINED THERMODYNAMICS STUDYNi Zhenjuan
 
Electrochemical reduction of Carbon Dioxide
Electrochemical reduction of Carbon DioxideElectrochemical reduction of Carbon Dioxide
Electrochemical reduction of Carbon DioxideSaurav Ch. Sarma
 
Carbon based catalysts for oxygen reduction reaction (ORR)
Carbon based catalysts for oxygen reduction reaction (ORR) Carbon based catalysts for oxygen reduction reaction (ORR)
Carbon based catalysts for oxygen reduction reaction (ORR) Lav Kumar Kasaudhan
 
Modification of CWZ-22 with KOH to enhance CO2 adsorption
Modification of CWZ-22 with KOH to enhance CO2 adsorptionModification of CWZ-22 with KOH to enhance CO2 adsorption
Modification of CWZ-22 with KOH to enhance CO2 adsorptionumut mutlu
 
CNT Hydrogen Storage Brief
CNT Hydrogen Storage BriefCNT Hydrogen Storage Brief
CNT Hydrogen Storage BriefAndy Zelinski
 
Martin Owen Jones (May 27th 2014)
Martin Owen Jones (May 27th 2014)Martin Owen Jones (May 27th 2014)
Martin Owen Jones (May 27th 2014)Roadshow2014
 

Similar to CO2_activation_on_bimetallic_CuNi_nanoparticles (20)

Structural aspect on carbon dioxide capture in nanotubes
Structural aspect on carbon dioxide capture in nanotubesStructural aspect on carbon dioxide capture in nanotubes
Structural aspect on carbon dioxide capture in nanotubes
 
ZE2159266272.pdf
ZE2159266272.pdfZE2159266272.pdf
ZE2159266272.pdf
 
C121326
C121326C121326
C121326
 
Simulation, modeling and fabrication of the bio-inspired flexible tactile sen...
Simulation, modeling and fabrication of the bio-inspired flexible tactile sen...Simulation, modeling and fabrication of the bio-inspired flexible tactile sen...
Simulation, modeling and fabrication of the bio-inspired flexible tactile sen...
 
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
 
acs%2Ejpcc%2E5b10334
acs%2Ejpcc%2E5b10334acs%2Ejpcc%2E5b10334
acs%2Ejpcc%2E5b10334
 
Theoretical study of two dimensional Nano sheet for gas sensing application
Theoretical study of two dimensional Nano sheet for gas sensing  applicationTheoretical study of two dimensional Nano sheet for gas sensing  application
Theoretical study of two dimensional Nano sheet for gas sensing application
 
F041033947
F041033947F041033947
F041033947
 
CARBON-CUPROUS OXIDE COMPOSITE NANOPARTICLES ON GLASS TUBES FOR SOLAR HEAT CO...
CARBON-CUPROUS OXIDE COMPOSITE NANOPARTICLES ON GLASS TUBES FOR SOLAR HEAT CO...CARBON-CUPROUS OXIDE COMPOSITE NANOPARTICLES ON GLASS TUBES FOR SOLAR HEAT CO...
CARBON-CUPROUS OXIDE COMPOSITE NANOPARTICLES ON GLASS TUBES FOR SOLAR HEAT CO...
 
Catal Commun 45 (2014) 153-Author
Catal Commun 45 (2014) 153-AuthorCatal Commun 45 (2014) 153-Author
Catal Commun 45 (2014) 153-Author
 
Wamg -graphene tubes_Nanoscale
Wamg -graphene tubes_NanoscaleWamg -graphene tubes_Nanoscale
Wamg -graphene tubes_Nanoscale
 
Characterization of Structural and Surface Properties of Nanocrystalline TiO2...
Characterization of Structural and Surface Properties of Nanocrystalline TiO2...Characterization of Structural and Surface Properties of Nanocrystalline TiO2...
Characterization of Structural and Surface Properties of Nanocrystalline TiO2...
 
Böhler et al. - 2014 - Recent advances in the study of the UO2–PuO2 phase dia...
Böhler et al. - 2014 - Recent advances in the study of the UO2–PuO2 phase dia...Böhler et al. - 2014 - Recent advances in the study of the UO2–PuO2 phase dia...
Böhler et al. - 2014 - Recent advances in the study of the UO2–PuO2 phase dia...
 
A first principles study into the properties and activities of rare-earth and...
A first principles study into the properties and activities of rare-earth and...A first principles study into the properties and activities of rare-earth and...
A first principles study into the properties and activities of rare-earth and...
 
H2 OXIDATION AT Pd100 A FIRST-PRINCIPLES CONSTRAINED THERMODYNAMICS STUDY
H2 OXIDATION AT Pd100 A FIRST-PRINCIPLES CONSTRAINED THERMODYNAMICS STUDYH2 OXIDATION AT Pd100 A FIRST-PRINCIPLES CONSTRAINED THERMODYNAMICS STUDY
H2 OXIDATION AT Pd100 A FIRST-PRINCIPLES CONSTRAINED THERMODYNAMICS STUDY
 
Electrochemical reduction of Carbon Dioxide
Electrochemical reduction of Carbon DioxideElectrochemical reduction of Carbon Dioxide
Electrochemical reduction of Carbon Dioxide
 
Carbon based catalysts for oxygen reduction reaction (ORR)
Carbon based catalysts for oxygen reduction reaction (ORR) Carbon based catalysts for oxygen reduction reaction (ORR)
Carbon based catalysts for oxygen reduction reaction (ORR)
 
Modification of CWZ-22 with KOH to enhance CO2 adsorption
Modification of CWZ-22 with KOH to enhance CO2 adsorptionModification of CWZ-22 with KOH to enhance CO2 adsorption
Modification of CWZ-22 with KOH to enhance CO2 adsorption
 
CNT Hydrogen Storage Brief
CNT Hydrogen Storage BriefCNT Hydrogen Storage Brief
CNT Hydrogen Storage Brief
 
Martin Owen Jones (May 27th 2014)
Martin Owen Jones (May 27th 2014)Martin Owen Jones (May 27th 2014)
Martin Owen Jones (May 27th 2014)
 

CO2_activation_on_bimetallic_CuNi_nanoparticles

  • 1. H O S T E D B Y Contents lists available at ScienceDirect Progress in Natural Science: Materials International journal homepage: www.elsevier.com/locate/pnsmi Original Research CO2 activation on bimetallic CuNi nanoparticles☆ Natalie Austin, Brandon Butina, Giannis Mpourmpakis ⁎ Department of Chemical Engineering, University of Pittsburgh, Pittsburgh, PA 15261, United States A R T I C L E I N F O Keywords: Density functional theory Alloys CO2 adsorption Catalysis Stability A B S T R A C T Density functional theory calculations have been performed to investigate the structural, electronic, and CO2 adsorption properties of 55-atom bimetallic CuNi nanoparticles (NPs) in core-shell and decorated architectures, as well as of their monometallic counterparts. Our results revealed that with respect to the monometallic Cu55 and Ni55 parents, the formation of decorated Cu12Ni43 and core-shell Cu42Ni13 are energetically favorable. We found that CO2 chemisorbs on monometallic Ni55, core-shell Cu13Ni42, and decorated Cu12Ni43 and Cu43Ni12, whereas, it physisorbs on monometallic Cu55 and core-shell Cu42Ni13. The presence of surface Ni on the NPs is key in strongly adsorbing and activating the CO2 molecule (linear to bent transition and elongation of C˭O bonds). This activation occurs through a charge transfer from the NPs to the CO2 molecule, where the local metal d-orbital density localization on surface Ni plays a pivotal role. This work identifies insightful structure- property relationships for CO2 activation and highlights the importance of keeping a balance between NP stability and CO2 adsorption behavior in designing catalytic bimetallic NPs that activate CO2. 1. Introduction Novel methods for CO2 capture, storage, and utilization are of marked interest for mitigating elevated levels of CO2 in the atmo- sphere, which contributes to the greenhouse effect [1–3]. As an example, industrial processes, utilizing transition metal (TM) catalysts, use CO2 as a chemical feedstock to produce important products such as formic acid and methanol [4]. Given the high thermodynamic stability of CO2, strong reducing species such as H2, are necessary to activate CO2 through hydrogenation reactions [5]. For example, industrial synthesis of methanol, which takes place on a Cu/ZnO/Al2O3 catalyst, requires a syngas feedstock containing CO, CO2 and H2 [4]. Studies on alkali-doped metal catalysts [6–8], bare metal catalysts [6,9,10], and metal oxide catalysts [7,11,12] have shown that CO2 can be activated in the absence of reducing species (i.e. H2). The activation of CO2 on these catalysts involves charge transfer from the metal system to the CO2 molecule, which results in a geometric change of CO2 from a linear to a bent configuration and elongation of the C˭O bonds [5,13]. A recent periodic Density Functional Theory (DFT) study on CO2 activation and dissociation on fcc(100) surfaces of Fe, Co, Ni, and Cu, showed that Co and Ni could favorably chemisorb CO2, transferring significant charge to the molecule from the metal, exhibiting the lowest barriers for CO2 dissociation [14]. Despite considerable research on TM interactions with CO2, kinetic limitations are still present in industrial processes that utilize CO2 as a chemical feedstock [15]. Therefore, it is important to identify catalysts and catalyst properties that activate CO2, in order to improve the activity of CO2 conversion reactions. Alloyed systems often exhibit very different properties from their monometallic counterparts, which make them very attractive for catalytic applications [16–18]. For example, experimental studies on Ni/Cu(100) surfaces have shown enhanced catalytic activity for metha- nol synthesis compared to the monometallic Cu(100) surface [19,20]. Additionally, a combined DFT and kinetic Monte Carlo study on the effects of metal doping (metal = Au, Pd, Rh, Pt, and Ni) of Cu(111) on methanol synthesis found that the Ni/Cu(111) surface showed the highest rate for methanol production compared to the other alloyed systems and the monometallic Cu(111) surface [21]. Temperature programmed desorption (TPD) experiments on a Ni film grown on a Cu(110) surface showed high rates of CO2 chemisorption compared to the Cu(110) surface on which only physisorption was observed [22]. The ability of CuNi systems to adsorb CO2 have been theoretically investigated on Ni-doped Cun (n=1–12) clusters [23]. The Ni-doped Cun systems chemisorbed CO2 in a bent state, while the pure Cun clusters only weakly interacted with CO2. In these clusters the strong CO2 adsorption was observed when CO2 was in direct contact with the Ni atom of the cluster. CuNi nanoparticle (NP) interactions with CO2 have also been studied on 55-atom NPs with Ni-doped and core-shell CuNi compositions [24]. In the doped system a single Ni atom was located in the core, surrounded by 54 Cu atoms and in the core-shell system 13 Ni atoms were located in the core and 42 Cu atoms were on http://dx.doi.org/10.1016/j.pnsc.2016.08.007 Received 31 July 2016; Accepted 16 August 2016 Peer review under responsibility of Chinese Materials Research Society. ⁎ Corresponding author. E-mail address: gmpourmp@pitt.edu (G. Mpourmpakis). Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx 1002-0071/ © 2016 Chinese Materials Research Society. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by/4.0/). Available online xxxx Please cite this article as: Austin, N., Progress in Natural Science: Materials International (2016), http://dx.doi.org/10.1016/j.pnsc.2016.08.007
  • 2. the shell. Only weak CO2 adsorption was observed on these systems as further indicated by the C˭O bond lengths and O˭C˭O bond angle of adsorbed CO2 (~1.172 Å and ~180°) remaining similar to that of gas phase CO2 (1.162 Å and 180°). As we have pointed out, it appears that the presence of Ni on the NP surface can facilitate strong CO2 adsorption on the catalyst. In addition, very recently the active site for CO2 hydrogenation to methanol on the industrial catalyst Cu/ZnO/Al2O3, was identified to be a Cu stepped surface decorated with Zn atoms [4]. According to DFT calculations in this study, the incorporation of Zn to the Cu(211) surface increased the adsorption strength of reaction intermediates and decreased the barriers for methanol synthesis compared to the monometallic Cu(211) surface. Since the presence of Zn on a catalyst surface enhances CO2 hydrogenation to methanol [4] and the presence of Ni on the surface of small clusters facilitates CO2 adsorption [23], in this study we investigate CO2 adsorption and activation on CuNi bimetallic NPs, accounting for Ni and Cu surface decorations, core-shell bime- tallic configurations, as well as, on their monometallic counterparts. We provide a thorough analysis on the structural, thermodynamic, and electronic properties of these NPs that could potentially lead to identifying stable and active catalytic systems for CO2 conversion to chemicals. 2. Computational Details We used the BP86 functional [25,26], the resolution of the identity (RI) approximation [27,28], and the def2-SV(P) [29,30] basis set as implemented in the TURBOMOLE 7.02 computational program pack- age [31]. Dispersion corrections have been taken into account in the calculations using the D3 method [32,33]. The total electronic energies of icosahedral 55-atom monometallic and bimetallic NPs composed of Cu and Ni were determined by geometry optimizations. Specifically, these structures include monometallic (Cu55 and Ni55), decorated (Cu43Ni12 and Cu12Ni43), and core-shell (Cu42Ni13 and Cu13Ni42) combinations of Cu and Ni atoms. Multiple spin states were considered in our calculations and the lowest in energy were selected for further analysis. Furthermore, all the optimized structures were verified as minima with frequency calculations (absence of any imaginary modes). The cohesive energy (binding energy per metal atom, BE/n) [34], a descriptor for average bond strength of the NPs, was calculated using Eq. 1: BE n = E -(55-x)E - (x)E 55 Cu Ni Cu Ni55-x x (1) where x is the number of Ni atoms in the NP (x =0, 12, 13, 42, 43, and 55) and ECu55-x,Nix, ECu, and ENi are the total energies of the Cu55−xNix NPs, Cu atom, and Ni atom, respectively. Using Eq. (2) we calculated the excess energy (Eexc) [35,36], a descriptor for the stability of the bimetallic NPs relative to their monometallic counterparts: E = E - E - E 55 exc Cu Ni (55-x) 55 Cu x 55 Ni55-x x 55 55 (2) where ECu55 and ENi55 are the total energies of the Cu55 and Ni55 NPs. The Eexc indicates if the formation of the bimetallic systems will be favorable (negative total energy) or unfavorable. Eq. (3) was used to calculate CO2 adsorption (binding energy, BE) on the metal NPs BE=E -E -ENP_CO NP CO2 2 (3) where ENP_CO2, ENP, and ECO2 are the total energies of CO2 interacting with the NP, the isolated NP, and isolated CO2, respectively. Adsorption was assessed with CO2 oriented towards one of the equivalent corner sites (coordination number 6, CN6) on the surface of the NPs, in two different adsorption configurations, horizontal and vertical to the NP surface. In the case of Cu12Ni43, CO2 adsorption was also investigated with CO2 oriented towards the edge site (CN8) of the NP where Ni atoms are located (vide-infra analysis). During CO2 optimization, the coordinates of the NPs were kept frozen at their optimized positions and the CO2 molecule was allowed to relax. The calculated adsorption states were further verified as minima with frequency calculations (absence of imaginary modes on CO2). Natural Fig. 1. Optimized geometries of 55-atom NPs: monometallic (a) Cu55 and (b) Ni55, decorated (c) Cu43Ni12 and (d) Cu12Ni43 NP with 12 heteroatoms located at the corner (CN6) sites of the NP, and core-shell (e) Cu13Ni42 and (f) Cu42Ni13 with 13 metal atoms located in the core and 42 heteroatoms in the shell. The equivalent CN6 and CN8 sites on the NPs are illustrated in (a). The Cu atoms are colored brown and the Ni blue. N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx 2
  • 3. bond orbital (NBO) analysis was used to calculate charge distribution on all the systems. Molecular orbital plots were visualized using TmoleX [37], a graphical user interface for TURBOMOLE. 3. Results and discussion 3.1. Structural and electronic properties of Cu55−xNix NPs The lowest energy structures of the six NPs studied in this work are shown in Fig. 1. The icosahedral shape, which is the lowest energy structure for monometallic Cu55 and Ni55[38], was also maintained for the bimetallic systems. The conservation of the icosahedral geometry can be attributed to the similar atomic radius of Cu (1.28 Å) [39] and Ni (1.25 Å) [39] atoms. The calculated BE/n and Eexc of the systems are shown in Table 1. The BE/n trend from the largest (most negative value) to lowest value is as follows: Ni55 > Cu12Ni43 > Cu13Ni42 > Cu42Ni13 > Cu43Ni12 > Cu55. Thus, we find that as the Ni fraction in Cu55−xNix increases, the BE/n also increases. This observation agrees with the melting points of the metals, with Ni (1728 K) [39] having higher melting point than Cu (1358 K) [39], and as a result, the average bond strength of their alloys shows larger values with higher Ni content. Furthermore we found that the BE/n trend of the mono- metallics followed that of experimental bulk (Cu 3.49 eV/atom and Ni 4.44 eV/atom) [39], where in this case the more positive BE/n represents the more stable system. The Eexc trend describes the stability of the bimetallic NPs relative to the monometallic parents. The Eexc trend from most stable to least stable is as follows: Cu42Ni13 > Cu12Ni43 > Cu43Ni12 > Cu13Ni42. Overall we found that the core-shell structure, Cu42Ni13, is the most energe- tically favorable formation for the bimetallic NPs. In theoretical work by Yang et al. the icosahedral Cu42Ni13 NP was also found to be energetically favorable. [24] The Eexc trends can be attributed to the surface energy (Cu 1170 ergs·cm−2 and Ni 2240 ergs·cm−2 ) [40] and bulk cohesive energy of monometallic Cu and Ni. The surface and cohesive energy values show that, in general, Cu prefers to reside on the surface and Ni prefers to be in the core of the NP. This core-shell preference has been shown to be present in nanoscale systems in the recent work by Wang et al. [41], calculating segregation energies of doped transition metal systems. In particular it was found that the Ni atom preferred to reside in the core than on the surface of a 55-atom cuboctohedral Cu54Ni. Due to the energetic preference of Cu for the surface and Ni for the core of the NPs, we find that the core-shell Cu42Ni13 is more energetically favorable than the corresponding Cu13Ni42 NP, and the decorated Cu12Ni43 is more favorable than the corresponding Cu43Ni12 NP. Overall the results show that the most energetically favorable NPs have Cu (lower cohesive and surface energy) on the lowest coordinated, surface sites, and Ni (higher cohesive and surface energy) on the highest coordinated, bulk sites of the NP. Table 1 Calculated BE/n and Eexc of the Cu55-xNix NPs. The negative values indicate exothermi- city. BE/n (eV/atom) Eexc (eV/atom) Cu55 −3.266 – Ni55 −4.155 – Cu43Ni12 −3.413 0.0560 Cu12Ni43 −3.980 −0.0218 Cu42Ni13 −3.535 −0.0692 Cu13Ni42 −3.849 0.113 1 2 1 2 1 2 1 2 1 2 1 2 Fig. 2. CO2 adsorption on the (a, b) monometallic, (c, d) decorated, and (e, f) core-shell Cu-Ni NPs. The color code is as depicted in Fig. 1 with the addition of CO2 (C colored gray and oxygen colored red). N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx 3
  • 4. 3.2. CO2 adsorption on the Cu55−xNix NPs Previous theoretical work on adsorption of small molecules, such as CO, on the surface of NPs showed that the stronger adsorption was observed on surface NP sites exhibiting low CNs (e.g. corners, edges) [42]. In turn, in the CO2 adsorption studies [5,6,14], strong adsorption of the CO2 molecule on the metal surface results to its activation. Thus, we primarily investigated the CO2 adsorption on the corner (CN6) site (the lowest coordinated site in our systems) of the Cu55−xNix NPs. As shown in Fig. 1(a), since all the CN6 sites are equivalent (same with the CN8 edge sites) due to the Ih symmetry of the NPs, we studied the CO2 adsorption on one of these sites. The lowest energy structures for CO2 adsorption are shown in Fig. 2. We calculated the adsorption energy for two orientations of the CO2 molecule: (i) horizontal CO2, where the C atom of CO2 was interacting with the corner site and (ii) vertical CO2, where an O atom of CO2 was interacting with the corner site. The horizontal orientation was found to be the most preferred adsorption configuration of CO2 on all the metal NPs in our study. Consistent with literature [5,9,12–14], we found that the systems that activate CO2 always show strong adsorption and charge transfer from the metal NP to CO2, which results in a linear to bent transition of the CO2 molecule, and elongation of the C˭O bonds. Specifically from Table 2 we found strong CO2 adsorption on Ni55, Cu43Ni12, Cu12Ni43, and Cu13Ni42 and weak adsorption on Cu55 and Cu42Ni13. In addition as shown in Table 2 the systems with strong CO2 adsorption had more than −0.6|e| transferred to the CO2 molecule while in the weakly adsorbed systems there was no significant charge transfer. The geometric properties of gas phase CO2 and CO2 interacting with the metal NPs are also shown in Table 2. Compared to gas phase (non-interacting) CO2, the O1-C-O2 angle of the strongly adsorbed CO2 decreased ( < 150°) and the C˭O bonds elongated ( > 1.2 Å) which indicates the activation of the CO2 molecule. In the weakly adsorbed systems the bond angles and bond distances of CO2 remained similar to gas phase CO2. The deviations of average bond lengths (black squares) and angles (blue circles) between adsorbed and gas phase CO2 as a function of CO2 BE are illustrated in Fig. 3. It is clear that both the CO2 bond distances and angles are affected in a similar way with the CO2 BE. To demonstrate this interdependence between the CO2 angle and bond distance deviation (from CO2 gas) we plotted these geometric properties in the inset of Fig. 3 and we observe a perfect linear trend. Overall, Fig. 3 shows that the bending of the CO2 molecule and the elongation of its bonds show the same behavior and they are enhanced (increased activation) with stronger adsorption on the NP surface. In Fig. 4(a) we demonstrate the CO2 BE as a function of the total charge transferred to the CO2 molecule. We notice that bimetallic systems with Ni atoms being on the surface of the NP transfer significant charge to CO2. In turn, this results in strong CO2 adsorption, while for the Cu42Ni13 system, where Ni is at the core of the NP and inaccessible to CO2, significant charge transfer is not observed, resulting to weak CO2 adsorption. It should be noticed that so far, our discussion on the activation of the CO2 molecule is entirely focused on the structural and electronic observations made on the CO2 molecule itself. From a catalyst design perspective, we need to identify a property of the metal NPs that could correlate with the observed activation. As a result, we made an effort to rationalize the CO2 adsorption behavior using the d-band center (dC) model by Hammer and Norskov [43,44]. We calculated the local dC on a single metal site of CO2 adsorption (Cu or Ni) on the Cu55−xNix NPs. In Fig. 4(b), the CO2 adsorption values are plotted as a function of the local site dC of the NPs. The observed linear trend indicates that there is an increase in CO2 adsorption with decrease in the dC, shifting towards the energy level of the CO2 Lowest Unoccupied Molecular Orbital (LUMO), which is located at −0.35 eV (blue dashed line). The Cu43Ni12 showed the highest local dC and strongest BE, while the Cu55 showed the lowest dC and the weakest BE. Notice that the Cu43Ni12 positions its local-site dC at the energy level of CO2 LUMO, showing the strongest adsorption. It is important to note that for Cu12Ni43 we did not observe strong adsorption when CO2 interacted directly with the corner Cu sites of the NP. Even the presence of Ni atoms in the neighboring positions (edge sites), did not enhance CO2 adsorption on the corner site compared to the Cu55 system. However, since surface Ni is responsible for enhancing the CO2 adsorption (compare CO2 adsorption on Cu55 vs. Cu43Ni12), we calculated the adsorption energy of CO2 on an edge site (CN8) of the Cu12Ni43, where the Ni atoms are located. In this case we observe a strong CO2 adsorption on Cu12Ni43 which was 0.4 eV lower in energy than the case where the CO2 interacted with the Cu atoms of the NP. This shows that the presence of Ni atoms on the surface of the NP significantly enhances CO2 adsorption. To further rationalize the effect of the metal's d-orbital density and the charge transfer on the CO2 activation, we plotted the Highest Occupied Molecular Orbitals (HOMO) of Cu55, Ni55, Cu43Ni12 and Cu12Ni43 NPs as shown in Fig. 5. These are the monometallic parents (Cu55 and Ni55), the bimetallic showing the strongest CO2 adsorption (Cu43Ni12) among the NPs studied, and the bimetallic NP that shows strong CO2 adsorption and favorable Eexc (Cu12Ni43). It has been shown that charge transfer from the metal to CO2 occurs when the d- orbitals of the metal NP significantly interacts with the 2Πu antibond- ing orbital (LUMO) of CO2 [13]. In Fig. 5 we assessed the presence of d-orbital character in the HOMO orbitals of the aforementioned NPs. Figs. 5(a) and (b) illustrate the HOMO orbital distribution on mono- metallic Cu55 and Ni55 respectively. From the shape of the orbitals we observe primarily s-orbital localization on Cu55, while there is mainly d-orbital localization on Ni55. By quantifying the orbital character through the atomic orbital coefficients of the HOMO, we found that Table 2 CO2 binding energies on the NPs, total charge (NBO) transferred to CO2, and geometric properties of gas phase CO2 (non-interacting) and CO2 interacting with the NPs. Binding energies are in eVs. Bond lengths of C-O1 and C-O2 are in angstroms, Å. Bond angle of O1-C-O2 are in degrees, °. BE (eV) |e| C-O1 (Å) C-O2 (Å) O1-C-O2 (°) CO2 (gas) – – 1.175 1.175 180.000 Cu55 −0.368 0.00825 1.169 1.183 179.406 Ni55 −0.665 −0.658 1.243 1.238 141.610 Cu43Ni12 −0.919 −0.604 1.245 1.230 143.465 Cu12 Ni43 −0.803 −0.810 1.216 1.309 132.070 Cu42Ni13 −0.391 0.00207 1.183 1.170 178.989 Cu13 Ni42 −0.891 −0.652 1.231 1.255 140.476 Fig. 3. Deviation of adsorbed CO2 average C˭O bond length (left ordinate) and O˭C˭O bond angle (right ordinate) from gas phase CO2 as functions of CO2 BE. The inset figure at the bottom left shows a linear relationship between average C˭O bond length and O˭C˭O bond angle. N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx 4
  • 5. there is a much greater fraction of d-orbital character in the HOMO of Ni55, which shows strong binding to CO2 than in Cu55, which weakly adsorbs CO2. The bimetallic systems Cu43Ni12 and Cu12Ni43 show high fractions d-orbital character with Cu12Ni43 having the highest. Cu43Ni12 has a lower fraction of d-orbital character than Cu12Ni43 because of the lower fraction of Ni atoms in Cu43Ni12 than in Cu12Ni43. It should also be noticed that the d-orbitals are localized on surface CN6 atoms of the Cu43Ni12, whereas, on the CN8 of the Cu12Ni43. In other words, the surface sites where the Ni atoms are located show strong CO2 adsorption. Overall we demonstrate that adding Ni on the surface of Cu NPs increases the presence of d-orbital character, which in turn, results in a favorable interaction of the NP with CO2, and subsequent activation of the CO2 molecule. Although we found that Cu43Ni12 binds CO2 the strongest, the most promising NP from our study is the Cu12Ni43 because in addition to the strong CO2 adsorption (BE =−0.8 eV) and activation, it shows favorable energetics for its synthesis (Eexc=−0.02 eV/atom). This study highlights that in bimetal- lic catalyst design it is important to achieve a balance between catalyst stability (the most stable CuNi NPs prefer Cu to be on the surface) and interaction strength of the catalyst with adsorbates (CO2 strongly adsorbs and is being activated on surface Ni). Although computational studies like this one identify bimetallic NPs that are stable and promising for CO2 activation, there is a synthetic challenge in forming nanostructures with (in-silico) predefined architecture. Recent experi- mental advances on the controlled synthesis of bimetallic NPs with atomic precision can pave the way towards achieving this goal [45,46]. 4. Conclusions In summary, we performed a DFT investigation on the structural, electronic and CO2 adsorption properties of Cu55−xNix (x=0, 12, 13, 42, 43, and 55) NPs with a monometallic, core-shell, and decorated distribution of Cu and Ni atoms. We found that the BE/n of the bimetallic systems was a linear combination of the BE/n of the monometallic systems. We also calculated the excess energy (Eexc) of the bimetallic NPs with respect to the monometallic NPs and showed that the formation of decorated Cu12Ni43 and core-shell Cu42Ni13 NPs were energetically favorable, while the formation of core-shell Cu13Ni42 and decorated Cu43Ni12 were less favorable. These trends rationalize the preference of Cu to be located at the surface of the NPs rather than Ni. CO2 adsorption calculations revealed weak interaction (physisorp- tion) with the monometallic Cu55 and core-shell Cu42Ni13, while the monometallic Ni55, decorated Cu12Ni43 and Cu43Ni12, and core-shell Cu13Ni42 chemisorbed CO2. In the chemisorbed cases we found strong adsorption of CO2 on the corner sites of all NPs except Cu12Ni43 where strong CO2 adsorption was found on the edge sites. The sites of strong adsorption on the NPs were always surface sites which were occupied by Ni atoms. Thus, the location of Ni on the NP plays an important role in the resulting adsorption behavior. The chemisorption behavior on the NPs was attributed to charge transferred from the metal NPs to CO2 which led to the activation of the molecule. Additionally, we calculated the local-site d-band center (dC) and we found a linear relationship between the dC and CO2 adsorption energy. The sites of strong adsorption localize HOMO orbitals with increased d-character. Fig. 4. CO2 BE as a function of (a) total charge on CO2 and (b) local dC of the Cu55−xNix NPs. The dashed black lines in both figures serve as a guide to the eye. In (b) the vertical blue line represents the LUMO orbital energy of the CO2 molecule. Fig. 5. Visual representation of the HOMO orbitals and fractional distribution of the HOMO orbital character of the (a) Cu55, (b) Ni55, (c) Cu43Ni12, (c) Cu12Ni43 NPs. N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx 5
  • 6. Overall this study demonstrates that the presence of surface Ni on CuNi bimetallic NPs can significantly enhance CO2 adsorption, resulting in the activation of the CO2 molecule. Furthermore, among the different nanostructures in this study we identified the Cu12Ni43, which can be potentially experimentally synthesized and activate CO2 for dissocia- tion and hydrogenation reactions due to its exothermic Eexc and strong adsorption behavior towards CO2, respectively. Acknowledgements This material is based upon work supported by the National Science Foundation Graduate Research Fellowship Program to N.A. under Grant No. (1247842). The authors would like to acknowledge compu- tational support from the Center for Simulation and Modeling (SAM) and start-up funding from the University of Pittsburgh. References [1] W. Wang, S.P. Wang, X.B. Ma, J.L. Gong, Chem Soc. Rev. 40 (2011) 3703–3727. [2] A. Goeppert, M. Czaun, G.K.S. Prakash, G.A. Olah, Energy Environ. Sci. 5 (2012) 7833–7853. [3] G. Centi, E.A. Quadrelli, S. Perathoner, Energy Environ. Sci. 6 (2013) 1711–1731. [4] M. Behrens, F. Studt, I. Kasatkin, S. Kuehl, M. Haevecker, F. Abild-Pedersen, et al., Science 336 (2012) 893–897. [5] H.J. Freund, M.W. Roberts, Surf. Sci. Rep. 25 (1996) 225–273. [6] F. Solymosi, J. Mol. Catal. 65 (1991) 337–358. [7] D. Ochs, M. Brause, B. Braun, W. Maus-Friedrichs, V. Kempter, Surf. Sci. 397 (1998) 101–107. [8] F. Solymosi, G. Klivenyi, Surf. Sci. 315 (1994) 255–268. [9] J. Ko, B.K. Kim, J.W. Han, J. Phys. Chem C. 120 (2016) 3438–3447. [10] S.G. Wang, D.B. Cao, Y.W. Li, J.G. Wang, H.J. Jiao, J. Phys. Chem B. 109 (2005) 18956–18963. [11] Y. Wang, R. Kovacik, B. Meyer, K. Kotsis, D. Stodt, V. Staemmler, et al., Angew. Chem.-Int. Ed. 46 (2007) 5624–5627. [12] K.R. Hahn, M. Iannuzzi, A.P. Seitsonen, J. Hutter, J. Phys. Chem C. 117 (2013) 1701–1711. [13] S.G. Wang, X.Y. Liao, D.B. Cao, C.F. Huo, Y.W. Li, J.G. Wang, et al., J. Phys. Chem C. 111 (2007) 16934–16940. [14] C. Liu, T.R. Cundari, A.K. Wilson, J. Phys. Chem C. 116 (2012) 5681–5688. [15] X.M. Liu, G.Q. Lu, Z.F. Yan, Beltramini, J. Ind. ENG Chem Res. 42 (2003) 6518–6530. [16] F. Studt, I. Sharafutdinov, F. Abild-Pedersen, C.F. Elkjaer, J.S. Hummelshoj, S. Dahl, et al., Nat. Chem 6 (2014) 320–324. [17] M.S. Chen, D. Kumar, C.W. Yi, D.W. Goodman, Science 310 (2005) 291–293. [18] A. Sandoval, A. Aguilar, C. Louis, A. Traverse, R. Zanella, J. Catal. 281 (2011) 40–49. [19] J. Nerlov, I. Chorkendorff, J. Catal. 181 (1999) 271–279. [20] J. Nerlov, S. Sckerl, J. Wambach, I. Chorkendorff, Appl Catal. a-Gen. 191 (2000) 97–109. [21] Y.X. Yang, M.G. White, P. Liu, J. Phys. Chem C. 116 (2012) 248–256. [22] E. Vesselli, E. Monachino, M. Rizzi, S. Furlan, X.M. Duan, C. Dri, et al., Acs Catal. 3 (2013) 1555–1559. [23] S.L. Han, X.L. Xue, X.C. Nie, H. Zhai, F. Wang, Q. Sun, et al., Phys. Lett. A 374 (2010) 4324–4330. [24] Y. Yang, D. Cheng, J. Phys. Chem C. 118 (2014) 250–258. [25] A.D. Becke, Phys. Rev. A 38 (1988) 3098–3100. [26] J.P. Perdew, Phys. Rev. B 33 (1986) 8822–8824. [27] K. Eichkorn, F. Weigend, O. Treutler, R. Ahlrichs, Theor. Chem Acc. 97 (1997) 119–124. [28] K. Eichkorn, O. Treutler, H. Ohm, M. Haser, R. Ahlrichs, Chem Phys. Lett. 240 (1995) 283–289. [29] A. Schafer, H. Horn, R. Ahlrichs, J. Chem Phys. 97 (1992) 2571–2577. [30] F. Weigend, M. Haser, H. Patzelt, R. Ahlrichs, Chem Phys. Lett. 294 (1998) 143–152. [31] R. Ahlrichs, M. Bar, M. Haser, H. Horn, C. Kolmel, Chem Phys. Lett. 162 (1989) 165–169. [32] S.J. Grimme, Comput Chem 25 (2004) 1463–1473. [33] S.J. Grimme, Comput Chem 27 (2006) 1787–1799. [34] N. Austin, G. Mpourmpakis, J. Phys. Chem C. 118 (2014) 18521–18528. [35] R. Ferrando, J. Jellinek, R.L. Johnston, Chem Rev. 108 (2008) 845–910. [36] M.J. Piotrowski, P. Piquini, J.L.F. Da Silva, J. Phys. Chem C. 116 (2012) 18432–18439. [37] C. Steffen, K. Thomas, U. Huniar, A. Hellweg, O. Rubner, A. Schroer, J. Comput Chem. 31 (2010) 2967–2970. [38] J.P.K. Doye, D.J. Wales, New J. Chem. 22 (1998) 733–744. [39] C. Kittel, Introduction to Solid State Physics, 7 ed, Wiley, New York, 1996. [40] Y.-N. Wen, H.-M. Zhang, Solid State Commun. 144 (2007) 163–167. [41] L.-L. Wang, D.D. Johnson, J. Am. Chem Soc. 131 (2009) 14023–14029. [42] G. Mpourmpakis, A.N. Andriotis, D.G. Vlachos, Nano Lett. 10 (2010) 1041–1045. [43] Hammer B, Norskov JK. Theoretical surface science and catalysis – calculations and concepts. In: Gates BC, Knozinger H, editors. Advances in Catalysis, Vol 45: Impact of Surface Science on Catalysis, 2000, pp. 71–129. [44] B. Hammer, J.K. Norskov, Surf. Sci. 343 (1995) 211–220. [45] R. Jin, K. Nobusada, Nano Res. 7 (2014) 285–300. [46] H. Qian, Li.G. Jiang D-e, C. Gayathri, A. Das, R.R. Gil, et al., J. Am. Chem Soc. 134 (2012) 16159–16162. N. Austin et al. Progress in Natural Science: Materials International xx (xxxx) xxxx–xxxx 6