You are on page 1of 9

Physica E 73 (2015) 12–20

Contents lists available at ScienceDirect

Physica E
journal homepage: www.elsevier.com/locate/physe

NO adsorption on nickel and nickel–manganese bimetallic clusters:


A density functional study
Ngangbam Bedamani Singh, B. Indrajit Sharma, Utpal Sarkar n
Department of Physics, Assam University, Silchar, 788011, India

H I G H L I G H T S G R A P H I C A L A B S T R A C T

 Using first-principle calculations, we Bimetallic Nin  1Mn clusters have higher magnetic moment and lower binding energy than pure Nin
investigate Ni–Mn clusters. clusters. NO adsorption is more favorable on bimetallic clusters.
 Introducing a Mn atom in Nin clus-
ters causes increase of magnetic
moment.
 Nin  1Mn clusters are more suitable
than Nin clusters for NO adsorption.

art ic l e i nf o a b s t r a c t

Article history: Using first principle calculations, geometry, binding energies and magnetic moments of Nin and Nin  1Mn
Received 24 November 2014 (n ¼2  8) clusters are reported. The stability of the clusters is discussed in terms of binding energy. The
Received in revised form relative stability is examined using the second order energy difference. Structures of Nin and Nin  1Mn
2 May 2015
clusters are similar. The bimetallic clusters have higher magnetic moments. We have investigated the
Accepted 12 May 2015
Available online 14 May 2015
adsorption of NO molecule on these clusters since it is one of the crucial steps for activation of catalytic
oxidation/reduction process. The adsorption of NO molecule is energetically more favored on bimetallic
Keywords: clusters. NO molecule prefers to be adsorbed near the Mn atom site. Bader charge analysis shows that
Bimetallic cluster electrons are transferred from metal cluster to NO molecule.
Stability
& 2015 Elsevier B.V. All rights reserved.
Magnetic moment
NO adsorption
Density functional theory

1. Introduction Bimetallic clusters, sometimes known as nanoalloys, are used


in nanotechnologies, catalysis and electrochemistry [9,10]. In ad-
In the last few decades, great interest has been developed to dition to the large surface area and quantum confinement, the
study the metal clusters [1–3]. Metal clusters exhibit unique bimetallic clusters have the advantage of tuning properties
chemical, magnetic and optical properties [4]. They show pro- through doping or alloying. Transition metal bimetallic clusters
mising potential applications in catalysis [5], hydrogen storage have potential applications in nanoelectronics and nanosensors
[6,7] and electronic devices [8]. Due to their large surface to vo- [11]. They are of special importance because of their unique cat-
lume ratio, their physical, chemical and electronic properties are alytic properties [12]. The highly stable PdAu nanoparticles act as
extremely dependent on their shape, size and composition. excellent catalysts in a number of reactions, such as reduction of
toxic pollutant NO [13], CO oxidation [14], CO and H2 adsorption
n
Corresponding author. [15]. The catalytic properties of the bimetallic nanoparticles are
E-mail address: utpalchemiitkgp@yahoo.com (U. Sarkar). highly related to the arrangement of atoms [15,16]. The basic aim

http://dx.doi.org/10.1016/j.physe.2015.05.012
1386-9477/& 2015 Elsevier B.V. All rights reserved.
N. Bedamani Singh et al. / Physica E 73 (2015) 12–20 13

Fig. 1. Lowest energy structures of Nin (n¼ 2  ) clusters. Here, B’s represent bond length in units of angstrom and bond angles are represented by A’s.

of these studies is to search for more stable catalysts with en- high activity for reduction of NO and NH3[27]. The adsorption of
hanced selectivity toward oxidation/reduction processes and bet- NO on Pd2, Cu2 and PdCu clusters have been reported using den-
ter magnetic properties. Most of the catalysts studies are done on sity functional calculations [28]. Endou et al. [29] investigated
nanoparticles consisting of Pt, Pd and Rh, which are precious and adsorption properties of precious metal clusters (Rh, Pd, Ag, Ir, Pt
expensive metals. Due to high cost, their large-scale practical ap- and Au) towards NO molecule. The adsorption process of NO
plication is limited and investigation of an alternative less ex- molecule on Pd clusters at various adsorption sites has been stu-
pensive metal like Ni is attractive. In this context, reducing the cost died using density functional calculations [30]. Further theoretical
of catalyst materials as well as improving the catalytic activity of studies of NO adsorption on charged and neutral gold clusters [31]
such low cost materials is highly important. The focus of many and silver clusters [32] are reported.
scientific investigations is to replace these precious metals with Among the 3d transition metals. Nickel has received highest
transition metal such as Fe, Co, Ni, Mn, Cu [17–21]. attention from research communities [20,21,33–37]. Nickel be-
NO emitted from industries, and automobiles is one of the cause of its better activity and low cost is one of the commonly
major pollutants in the present-day world. It is related to forma- used catalysts’ material [20,21]. Recently, it was reported that
tion of acid rain, destruction of ozone layer and global warming. nickel acts as an excellent catalyst in the liquid phase hydrode-
For the removal of this toxic pollutant, catalytic reduction is the chlorination of chlorobenzene, giving 99.3% conversion [33].
most investigated approach [22,23]. Another approach for re- Nickel exhibits a greater adsorption ability than Pt, Pd and Rh for
moving NO is the catalytic oxidation of NO to NO2, which can be molecule such as ethyne [34]. Moreover, in several research works,
easily absorbed by alkali solutions [24]. The adsorption of mole- nickel cluster has been the focus of interests due to its important
cules on metal clusters is an essential part of the activation steps magnetic properties [35–37]. While the majority of the cluster
during the catalytic oxidation/reduction reaction. Very recently, studies are focussed on the properties of mono-atomic clusters,
Shu et al. [25] have shown that bimetallic Ni–Mn, Co–Mn and Zn– properties of heteroatomic clusters have also been investigated in
Mn oxides have enhanced efficiencies in NO removal as compared some studies. It is reported that bimetallic clusters exhibit better
with corresponding single metal oxides. Wu and Ho [26] in- catalytic activities than the corresponding pure metal clusters due
vestigated the interaction of NO, NO2 and N2O with nickel. The to better activation, stability and selectivity [38,39]. Again, mag-
adsorbed species exhibited a tendency to dissociate and produce netic transition metal clusters such as Ni, Mn are promising can-
atomic N and O on the nickel surface. Mn based catalysts displayed didate for ultrahigh density magnetic storage devices [40]. Clusters
14 N. Bedamani Singh et al. / Physica E 73 (2015) 12–20

Fig. 2. Lowest energy structures of bimetallic Nin  1Mn (n ¼2  8) clusters. Nickel atoms are represented by blue color spheres and violet spheres are Mn atoms. Here, B’s
represent bond length in units of angstrom and bond angles are represented by A’s.

have unusual magnetic properties as compared to their bulk form used for conversion of semiconductor material into ferromagnetic
[41,42]. In contrast to bulk form, increase in magnetic moment phase [46] and to introduce ultrahigh magnetic moment [47] in
with Mn concentration was reported for bimetallic Co–Mn cluster host cluster.
[42]. For Ni–Mn, Fe–Mn and Co–Mn clusters, monotonic change in
magnetic moment is observed with increasing Mn concentration
[42]. 2. Calculation method
To design more efficient catalysts nanomaterials, a detailed
knowledge of atomic and electronic structures of these nano- In the first-principle study of clusters, the possibility of ob-
particles is crucial. The remarkable progress of theoretical meth- taining the global minimum structure is dependent on the selec-
ods enables the determination of cluster geometry accurately. The tion of initial geometries. The number of possible initial geome-
present study is aimed at understanding how the electronic, cat- tries for clusters increases rapidly with cluster size. Initial geo-
alytic and magnetic properties of Ni clusters change on doping metries of the clusters are generated from a constant temperature
with a single Mn atom. To the best of our knowledge, chemi- ab-initio molecular dynamics simulation near the melting tem-
sorption studies of NO molecule on bimetallic Ni–Mn clusters have perature of bulk nickel. Ab-initio molecular dynamics are carried
not been reported. It is expected that our results will be helpful for out for 5000 fs in the NVT ensemble and temperature is controlled
understanding the unique properties of Nin  1Mn (n¼ 2–8) using the Nose thermostat. Newton’s equations of motion are in-
clusters. tegrated using the Verlet algorithm with a time step of 1.00 fs.
In the present work, we examined how single Mn atom affects During the simulation process, the structure of the cluster goes
the structure, magnetic and NO adsorption properties of Nin through various changes. Based on the lowest energy criteria, we
clusters. We have selected Mn because Mn is the third most chose some of these structures with reasonably different geome-
abundant transition metal on earth and has special importance in tries. For our choice of initial structures to be more extensive, we
catalystic applications [43–45]. As a dopant material Mn can tune have also considered those geometries of Nin clusters reported in
the magnetic moment of the host cluster and this tunable mag- the literature [48–50]. While choosing the initial structures of bi-
netic property can be used to engineer nanomagnets. It can also be metallic clusters, we considered those structures which are
N. Bedamani Singh et al. / Physica E 73 (2015) 12–20 15

Table 1 Table 2
Structure parameters for Nin clusters. Here, relative energy is the difference in Structure parameters for Nin  1Mn clusters. Here, relative energy is the difference in
energy with respect to energy of the most stable structure. energy with respect to energy of the most stable structure.

Cluster Geometry Relative energy Symmetry Cluster Geometry Relative Symmetry


energy
Ni2 Linear (stable) 0.000 Dαh
Linear (isomer) 0.010 Dαh NiMn Linear (stable) 0.000 Cαv

Ni3 Equilateral (stable) 0.000 D3h Linear (isomer) 0.012 Cαv


Linear (isomer) 0.259 C2v
Isosceles (isomer) 0.279 D3h

Ni4 Parallelogram (stable) 0.000 CS Ni2Mn Isosceles triangle 0.000 C2v


Tetrahedron (isomer) 0.518 D2d
Square (isomer) 0.890 D4h Linear with Mn at edge (isomer) 0.559 CS

Ni5 Trigonal bipyramid (stable) 0.000 D3h Linear with Mn at center (isomer) 0.910 C2v
Planar W-shaped (isomer) 1.267 C2V
Side capped square (isomer) 2.284 C2V

Ni6 Tetragonal bipyramid (stable) 0.000 Ci Ni3Mn Planar diamond-shape (stable) 0.000 CS
Capped trigonal bipyramid (isomer) 0.337 C1
Square prism (isomer) 1.636 D3h Tetrahedron (isomer) 0.206 C3v
Ni7 Pentagonal bipyramid (stable) 0.000 CS
Capped octahedron (isomer) 0.177 C1 Distorted planar diamond-shape with 0.827 C1
Disorted capped octahedron (isomer) 0.961 C1 Mn at different site (isomer)

Ni8 Bicapped octahedron (stable) 0.000 D2d


Capped pentagonal bipyramid (isomer) 0.291 C1
Cube (isomer) 1.073 Oh Ni4Mn Trigonal bipyramid (stable) 0.000 CS

Planar (isomer) 0.152 C2v


obtained by placing one Mn atom at each possible site of Nin  1
Trigonal bipyramid with Mn at different 0.373 CS
clusters. We also considered the structures which are obtained position (isomer)
when one nickel atom of the Nin cluster is replaced by a manga-
nese atom.
Our calculations are performed in the framework of spin-po-
Ni5Mn Tetragonal bipyramid (stable) 0.000 C4v
larized density functional theory as implemented in SIESTA 3.2
[51]. The SIESTA solves the single-particle Kohn–Sham equations Tetragonal bipyramid with different Mn 0.105 C1
using a flexible, numerical LCAO basis set and this approach has position (isomer)
been proven to successfully describe various systems such as free
Capped trigonal bipyramid (isomer) 0.521 C1
and supported clusters, molecules, surfaces, nanotubes, etc. [52].
Since Local Density Approximation (LDA) does not take into ac-
count gradient corrections, we have employed Generalized Gra-
dient Approximation (GGA) using the Perdew–Burke–Ernzerhof Ni6Mn Pentagonal bipyramid (stable) 0.000 C1
(PBE) [53] exchange correlation functional. The atomic cores are
Capped tetragonal bipyramid (isomer) 0.568 C1
replaced by nonlocal norm-conserving Troullier–Martins pseudo-
potential [54]. The valence states are described with double-ζ Pentagonal bipyramid with different 1.002 C1
polarized basis set. The geometry relaxations are carried out using Mn position (isomer)
the conjugate gradient method. The mesh cutoff has been set at
300 Rydberg to define the finite real-space grid for numerical
calculations. We have also tested triple-ζ polarized basis set and Ni7Mn Bicapped octahedron with Mn at cap- 0.000 C1
other mesh cutoff values but they do not substantially change the ping site (stable)
results and hence the results obtained using triple-ζ polarized
Bicapped octahedron with Mn at octa- 0.110 C1
basis set are presented as supplementary information. The ap-
hedron (isomer)
proach of GGA with PBE functional, has been successfully applied
to describe nickel cluster [55], manganese cluster [56], nickel– Capped pentagonal bipyramid (isomer) 0.464 C1
manganese [57] and bimetallic transition metal clusters [58,59].
Also, adsorption of NO on Rh cluster was investigated by the same
approach, where double-ζ polarized basis set was employed to clusters can be predicted by the second-order difference of energy
describe the valence states for N, O and Rh atoms [60]. Full geo- (ΔE2) as
metry relaxation is done until the forces on the atoms are less than
0.02 eV/Å. We have used 25  25  25 Å3 supercell, which is suf- ΔE2(Nin) = E(Nin − 1) + E(Nin + 1) − 2E(Nin) (2)
ficiently large enough to neglect interactions among clusters in the
neighboring cells. For the Mn-doped bimetallic Nin  1Mn (n ¼2–8) clusters, the
Binding energy per atom (BE) is calculated as binding energy has been calculated as

nE(Ni) − E(Nin) (n − 1)E(Ni) + E(Mn) − E(Nin − 1Mn)


BE(Nin) = BE(Nin − 1Mn) =
n (1) n (3)
where E(Nin) is the energy of nickel cluster having n atoms and E where E(Nin  1Mn) is the energy of bimetallic Ni–Mn cluster with
(Ni) is the energy of a single Ni atom. The relative stability of the (n  1) number of Ni atoms and E(Mn) is the energy of a single Mn
16 N. Bedamani Singh et al. / Physica E 73 (2015) 12–20

Fig. 3. (a) Binding energy and (b) second order energy difference of Nin and bimetallic Nin  1Mn clusters.

η = IP − EA (7)

where IP is calculated as the energy difference between positive


ion and neutral system at the neutral geometry. EA is obtained as
the difference in neutral energy and negative ion energy at the
neutral geometry.

3. Results and discussion

3.1. Geometry of Ni and Ni–Mn cluster


Fig. 4. Calculated chemical hardness of pure Nin clusters and bimetallic Nin  1Mn
clusters as a function of cluster size. The most stable structures of pure Nin clusters and Mn-doped
bimetallic Nin  1Mn (n ¼2–8) clusters are shown in Figs. 1 and 2. In
Tables 1 and 2, the relative energies and symmetries of most stable
structures of the clusters along with low-lying isomers are re-
ported. Figures of these low-lying isomers are provided in sup-
plementary information, Fig. S1 and Fig. S2.
The Ni dimer has a bond length 2.111 Å, which agrees with the
experimental value of 2.200 Å [62]. An isomer with bond length
2.063 Å is observed at slightly higher energy. For NiMn, the lowest
energy structure has a bond length of 2.794 Å, and its isomer with
bond length 1.894 Å is 0.012 eV higher in energy. The most stable
structure for Ni3 is found to be an equilateral triangle, however its
bimetallic counterpart, i.e. Ni2Mn is an isosceles triangle. The
lowest energy structure of Ni4 cluster is found to be a parallelo-
Fig. 5. Calculated magnetic moments of pure Nin clusters and bimetallic Nin  1Mn
gram. Replacing one Ni atom by Mn changes this parallelogram
clusters as a function of cluster size.
structure (Ni4) into a planar diamond-shaped structure (Ni3Mn).
Ni5 and Ni4Mn clusters are observed to be trigonal bipyramid. An
atom. The second-order difference of energy (ΔE2) is calculated as
experimental study also reported the same structure for Ni5[63]. In
ΔE2(Nin − 1Mn) = E(Nin − 2Mn) + E(NinMn) − 2E(Nin − 1Mn) (4) case of Ni6 and Ni5Mn clusters, the structures are tetragonal bi-
pyramid whereas for Ni7 and Ni6Mn clusters, the structures are
The adsorption energy of NO on Nin (n ¼2–8) clusters is cal-
pentagonal bipyramid. Experimental study on the uptake of am-
culated as
monia on clusters suggests that the pentagonal bipyramid is the
EAds(Nin) = E(Nin) + E(NO) − E(NinNO) (5) most likely structure of Ni7[64]. Bicapped octahedron is observed
to be the lowest energy structure for Ni8 and Ni7Mn cluster. Our
where E(NO) is the energy of NO molecule and E(NinNO) is the results suggest that most of the stable structures of Nin clusters
energy of NO adsorbed Nin cluster system. Similarly, the adsorp- (n ¼2  8) are similar to those of corresponding Nin  1Mn clusters.
tion energy of NO on Nin  1Mn (n ¼2–8) clusters is calculated as The average Ni–Ni bond length in Nin (n ¼2  8) clusters are
EAds(Nin − 1Mn) = E(Nin − 1Mn) + E(NO) − E(Nin − 1MnNO) (6) 2.111, 2.140, 2.310, 2.360, 2.340, 2.350, and 2.360 Å for Ni2, Ni3, Ni4,
Ni5, Ni6, Ni7 and Ni8 respectively. The Ni–Ni average bond length
Chemical hardness (η) which gives a measure of the resistance increases as the cluster size grows from n ¼2 5. But the average
of the system to changes in number of electrons and is calculated bond length of Ni6 becomes smaller than Ni5 and also from Ni7 and
as difference between ionization potential (IP) and electron affinity Ni8. This decrease in bond length suggests a more compact con-
(EA) [61]. figuration of Ni6. Similarly, smaller Ni–Ni average bond length for
N. Bedamani Singh et al. / Physica E 73 (2015) 12–20 17

Fig. 6. Lowest energy structures of NO adsorbed on Nin clusters. Here, B’s represent bond length in units of angstrom and bond angles are represented by A’s.

six atom cluster is observed in the case of Nin  1Mn. The Ni–Ni particular stability. In the case of Nin  1Mn clusters, highest peak is
average bond length for Ni2Mn, Ni3Mn, Ni4Mn, Ni5Mn, Ni6Mn and observed for Ni5Mn cluster. Six-atom clusters are relatively more
Ni7Mn are 2.160, 2.270, 2.350, 2.300, 2.350 and 2.350 respectively. stable than their neighbors. The high stability of Ni6 and Ni5Mn
The six atom cluster, which has higher stability than their neigh- may be attributed to its geometry with high coordination
bors have a more compact configuration. numbers.
The chemical hardness (η) value is reported in Fig. 4. Clusters
3.2. Stability of cluster with six-atoms, viz. Ni6 and Ni5Mn clusters are observed to have
higher hardness values than their neighboring clusters. It again
The binding energy (BE) per atom of the clusters are presented suggests enhanced stability of six-atom clusters which was seen
in Fig. 3a. Two important points are worth to be noted here. First, from the second order energy difference (ΔE2). Chemical hardness
the BE gradually increases with the cluster size. It is due to in- of Ni3 nd Ni2Mn are lower than their corresponding neighbors
creased average coordination on increasing the cluster size. Sec- which shows the lower stability of these clusters as previously
ondly, clusters having single Mn atom (i.e. Nin  1Mn clusters) have seen from second order energy difference (ΔE2).
lower BEs than the corresponding Nin clusters. Thus, doping with
Mn atom leads to decrease the cluster stability. The observation 3.3. Magnetic properties
can be understood from the fact that BEs of Mn clusters [65] are
smaller than those of pure Ni clusters [66]. In cluster form, nickel has enhanced magnetic moment per
The second-order energy difference (ΔE2) is an important and atom than the corresponding bulk [67]. It is reported that certain
very sensitive quantity which reflects the relative stability of metals (such as V, Rh and Pd), which are non-magnetic in bulk
clusters. The calculated second order energy differences are de- form can have net magnetic moments in their cluster forms
picted in Fig. 3b. Among Nin clusters, Ni6 has the highest value of [68,69]. In Fig. 5, we have presented the magnetic moments of
second-order energy difference, suggesting this cluster possess a pure Nin and bimetallic Nin  1Mn (n¼ 2  8) clusters. The
18 N. Bedamani Singh et al. / Physica E 73 (2015) 12–20

Fig. 7. Lowest energy structures of NO adsorbed on bimetallic Nin  1Mn clusters. Here, B’s represent bond length in units of angstrom and bond angles are represented by
A’s.

Table 3
Bader charges on N, O and Mn atoms in Nin  1Mn–NO complex. Charges are in units
of |e|.

System Mn N O

Ni1Mn–NO þ 0.549  0.256  0.274


Ni2Mn–NO þ 0.529  0.146  0.428
Ni3Mn–NO þ 0.144  0.183  0.343
Ni4Mn–NO þ 0.544  0.208  0.364
Ni5Mn–NO þ 0.486  0.216  0.309
Ni6Mn–NO þ 0.555  0.065  0.467
Ni7Mn–NO þ 0.596  0.127  0.384

3d orbital), which results in a magnetic moment of 5μB. However,


Fig. 8. Adsorption energies for NO molecule on pure Nin clusters and bimetallic
Nin  1Mn clusters as function of cluster size. a Ni atom has eight electrons in its valence electronic configura-
tion, out of which six electrons are in pair and, the remaining two
substitutional doping of Ni2 by one Mn atom increases the total electrons are left unpaired. This configuration of Ni atom resulted
magnetic moment from 3.924 to 7.000μB. For Nin (n ¼3  8), sub- to a magnetic moment of 2μB. The Mn atom has larger magnetic
stitution with one Mn atom increases the total magnetic moment moment than Ni atom. Therefore, the magnetic moments of Nin
by 3μB. This may be attributed to the redistribution of electrons in clusters are enhanced when substituted with a Mn atom.
the molecular orbitals of Ni and Mn atoms. When a Ni atom
(4s23d8) is replaced by a Mn atom (4s23d5), the corresponding 3.4. NO adsorption on Ni cluster and bimetallic Ni–Mn cluster
cluster loses three valence electrons. The Mn atom has five un-
paired electrons in its valence electronic configuration (half-filled The interaction of NO molecule with Nin and Nin  1Mn
N. Bedamani Singh et al. / Physica E 73 (2015) 12–20 19

(n ¼2  8) clusters are investigated. All the possible adsorption systems, electrons from Nin cluster are transferred to NO molecule.
sites, i.e., on-top of Ni atoms, on-top of Mn atom, Ni–Ni bridge site, The transferred electrons are primarily accumulated on the ni-
Mn–Ni bridge site, hollow adsorption sites formed by Ni–Ni–Ni, trogen atom. The Bader charge analysis shows that charges on O
and Ni–Ni–Mn have been considered. The resulting most stable atom are  0.283,  0.511,  0.505,  0.494,  0.366, 0.449,
structures of NO adsorbed clusters are shown in Fig. 6 and higher  0.348 |e| when NO molecules are adsorbed to Ni2, Ni3, Ni4, Ni5,
energy structures are presented in supplementary information Ni6, Ni7 and Ni8 respectively. And, charges on N atoms are  0.049,
(Fig. S3 and Fig. S4). Our calculated bond length for NO in pure  0.137,  0.198,  0.168,  0.020,  0.217,  0.193 |e| in Ni2–NO,
state is 1.160 Å. The NO molecule is slightly tilted from the top Ni3–NO, Ni4–NO, Ni5–NO, Ni6–NO, Ni7–NO and Ni8–NO respec-
position in Ni2–NO. The Ni–Ni bond is slightly elongated from tively. In all the Nin–NO complexes, we observed that the nitrogen
2.111 Å in bare Ni2 cluster to 2.114 Å in Ni2–NO. NO adsorbed on atom accepts electrons. Whereas, the oxygen atom loses electron
the top and bridge sites are 0.064 and 0.083 eV higher in energy. and becomes electron deficient, except for Ni3–NO, Ni4–NO and
The equilateral triangular structure of bare Ni3 cluster changes to Ni5–NO complexes. Table 3 shows the Bader charges on Mn, N and
form an isosceles triangle after NO adsorption. Top and bridge site O atoms for the Nin  1Mn–NO complexes. We found that there is
adsorption of NO molecule are 0.431 and 1.092 eV higher in en- electron transfer from manganese atom, which becomes positively
ergy. For Ni4–NO, the NO molecule is adsorbed at the bridge site. charged as high as þ0.596 |e| (for Ni7Mn–NO) to the NO molecule.
The planar parallelogram structure of Ni4 cluster is observed to The Bader charges of Nin  1Mn before and after NO adsorption are
form a tetrahedron after NO adsorption. Adsorption of NO at the presented in supplementary information (Table S1). In Nin  1Mn–
hollow site is 0.568 eV higher in energy. In Ni5–NO, the NO mo- NO, electrons are transferred from metal cluster to NO molecule.
lecule prefers to be adsorbed at the hollow site and top site ad- After forming the Nin  1Mn–NO complex, manganese and oxygen
sorption is 0.754 eV higher in energy. In the lowest energy state of become electron deficient, whereas, nitrogen gains electrons. Due
Ni6–NO, the NO molecule is adsorbed on the top site and energy of to the redistribution of electron, a little portion of nickel atoms are
bridge site adsorption is 0.004 eV above the energy of most stable electron rich, except in NiMn–NO and Ni2Mn–NO complexes.
structure. For Ni7–NO, the NO molecule prefers to be adsorbed at
the hollow site but NO molecule adsorbed at the bridge site lies
0.607 eV higher than the minimum energy state. Adsorption of NO 4. Conclusion
molecule takes place at the bridge site for Ni8–NO, however the
second-lowest energy structure in which the absorption of NO We have studied the effect of Mn doping on Nin (n ¼2–8)
takes place on top site is found to be 0.312 eV higher in energy. clusters using density functional theory. The most stable struc-
From Fig. 7, it is observed that NO molecule prefers to bind tures of Nin  1Mn (n¼ 2–8) are similar to those of Nin clusters. The
with the Mn atom of Nin  1Mn alloy clusters. In NiMn–NO, the Ni– substitution of Mn leads to an enhanced magnetic moment in the
Mn–N angle is observed to be 115°. In Ni2Mn–NO complex, the NO clusters. Monometallic clusters (Nin) have higher BE than the bi-
molecule adsorbed at the side of Mn atom is 0.233 eV lower in metallic Nin  1Mn (n ¼2–8) clusters. The second order energy
energy than adsorption on the hollow site. For the case of Ni3Mn– difference suggests that Ni6 cluster has the highest stability among
NO, Ni4Mn–NO, Ni5Mn–NO, Ni6Mn–NO and Ni7Mn–NO, the NO Nin clusters. Ni5Mn is more stable than its neighboring clusters.
molecule is adsorbed on the top-site of Mn atom and for theses Adsorption of NO molecule leads to structural changes of the
clusters, energy of the next higher configuration lie 0.235, 0.674, clusters but do not lead to any change in magnetic moment. The
0.200, 0.125 and 0.418 eV above the energy of their corresponding adsorption of NO molecule is energetically more favored on
most stable structure. In both, mono-atomic nickel and bimetallic Nin  1Mn than Nin cluster. NO molecule prefers to be adsorbed
nickel–manganese clusters, the N–O bond is elongated, albeit near the Mn atom. From the Bader charge analysis, the NO mo-
different for different cluster size, after adsorption in the clusters. lecule is observed to acquire electrons from the cluster.
We do not observe any change in magnetic moment of the cluster
systems after adsorption of NO molecule.
The adsorption energies of NO molecule with bare Nin clusters Acknowledgment
and bimetallic Nin  1Mn clusters are shown in Fig. 8. It is observed
that the bimetallic clusters have higher adsorption energy than Dr. Utpal Sarkar acknowledge the support from SHARCNET
pure nickel clusters. Thus, doping of Nin clusters with a manganese Canada for providing the computational facilities for this research
atom enhances the NO adsorption energies. If we compare the work.
distance between N and nearest Ni atom in Nin–NO with the dis-
tance between N and Mn atom in corresponding Nin  1Mn–NO, we
see that N–Ni bond is shorter than N–Mn bond, except Ni6–NO in Appendix A. Supplementary material
which N–Ni bond length is 0.003 Å smaller than N–Mn bond
length of Ni5Mn–NO. This smaller bond length of N–Mn may be Supplementary data associated with this article can be found in
the possible reason for higher adsorption energy of Nin–1Mn the online version at http://dx.doi.org/10.1016/j.physe.2015.05.012.
cluster compared to Nin cluster. Ni3–NO, Ni7–NO and their bime-
tallic counterparts are relatively more stable than their neighbor-
ing complexes, whereas Ni6NO and its bimetallic counterpart are References
relatively less stable than their neighbors.
[1] U. Sarkar, S. Blundell, Phys. Rev. B 79 (2009) 125441–125447.
3.5. Charge transfer analysis [2] S. Pan, S. Jalife, R.M. Kumar, V. Subramanian, G. Merino, P.K. Chattaraj,
ChemPhysChem 14 (2013) 2511–2517.
[3] N.B. Singh, U. Sarkar, Mol. Simul. 40 (2014) 1255–1264.
We have carried out Bader charge analysis [70] on the lowest
[4] R.L. Johnston, Atomic and Molecular Clusters, Taylor and Francis, London,
energy structures of Nin–NO and Nin  1Mn–NO complexes. The 2002.
“Bader charge” is the sum of the electronic valence charge density [5] S. Vajda, M.J. Pellin, J.P. Greeley, C.L. Marshall, L.A. Curtiss, G.A. Ballentine, J.
and a set of “model core charges” placed at the atomic sites [70]. W. Elam, S. Catillon-Mucherie, P.C. Redfern, F. Mehmood, P. Zapol, Nat. Mater. 8
(2009) 213–216.
On the basis of Bader charge analysis, charge transfer between the [6] S. Bandaru, A. Chakraborty, S. Giri, P.K. Chattaraj, Int. J. Quant. Chem. 112
clusters and adsorbed NO molecule is analyzed. In the Nin–NO (2011) 695–702.
20 N. Bedamani Singh et al. / Physica E 73 (2015) 12–20

[7] S. Giri, A. Chakraborty, P.K. Chattaraj, J. Mol. Model. 17 (2011) 777–784. [38] D. Bazin, C. Mottet, G. Treglia, Appl. Catal.: Gen. 200 (2000) 47–54.
[8] D.M. Adams, L. Brus, C.E.D. Chidsey, S. Creager, C. Creutz, C.R. Kagan, P. [39] A.M. Molenbroek, J.K. Norskov, B.S. Clausen, J. Phys. Chem. B 105 (2001)
V. Kamat, M. Lieberman, S. Lindsay, R.A. Marcus, R.M. Metzger, M.E. Michel- 5450–5458.
Beyerle, J.R. Miller, M.D. Newton, D.R. Rolison, O. Sankey, K.S. Schanze, [40] S. Sun, C.B. Murray, D. Weller, L. Folks, A. Moser, Science 287 (2000)
J. Yardley, X.Y. Zhu, J. Phys. Chem. B 107 (2003) 6668–6697. 1989–1992.
[9] R. Ferrando, J. Jellinek, R.L. Johnston, Chem. Rev. 108 (2008) 845–910. [41] I.M.L. Billas, A. Châtelain, W.A. de Heer, Science 265 (1994) 1682–1684.
[10] V. Caps, S. Arrii, F. Morfin, G. Bergeret, J.L. Rouss, Faraday Discuss. 138 (2008) [42] G. Rollmann, S. Sahoo, A. Hucht, P. Entel, Phys. Rev. B 78 (2008) 134404/1-5.
241–256. [43] M.M. Najafpour, M.Z. Ghobadi, B. Haghighi, T. Tomo, J.R. Shen, S.
[11] P.C. Chen, J.Y. Ma, L.Y. Chen, G.L. Lin, C.C. Shih, T.Y. Lin, H.T. Chang, Nanoscale 6 I. Allakhverdiev, Biochim. Biophys. Acta 1847 (2015) 294–306.
(2014) 3503–3507. [44] M.M. Najafpour, F. Rahimi, E.M. Aro, C.H. Lee, S.I. Allakhverdiev, J. R. Soc. In-
[12] J.D. Aiken III, R.G. Finke, J. Mol. Catal. A: Chem. 145 (1999) 1–44. terface 9 (2012) 2383–2395.
[13] M. Valden, J. Aaltonen, E. Kuusisto, M. Pessa, C.J. Barnes, Surf. Sci. 307 (1994) [45] R.K. Hockling, R. Brimblecombe, L.Y. Chang, A. Singh, M.H. Cheah, C. Glover, W.
193–198. H. Casey, L. Spiccia, Nat. Chem. 3 (2011) 461–466.
[14] S.L. Peng, L.Y. Gan, R.Y. Tian, Y.J. Zhao, Computat. Theor. Chem. 977 (2011) [46] H. Ohno, Science 281 (1998) 951–956.
62–68. [47] J. Wng, J. Bai, J. Jellinek, X.C. Zeng, J. Am. Chem. Soc. 129 (2007) 4110–4111.
[15] F. Maroun, F. Ozanam, O. Magnussen, R. Behm, Science 293 (2001) 1811–1814. [48] V.G. Grigoryan, M. Springborg, Chem. Phys. Lett. 375 (2003) 219–226.
[16] M. Chen, D. Kumar, C. Yi, D. Goodman, Science 310 (2005) 291–293. [49] C. Luo, Model. Simul. Mater. Sci. Eng. 8 (2000) 95–101.
[17] H. Hamada, Y. Kintaichi, M. Sasaki, T. Ito, M. Tabata, Appl. Catal 75 (1919) [50] S. Goel, A.E. Masunov, J. Mol. Model. 18 (2012) 783–790.
L1–L8. [51] J.M. Soler, E. Artacho, J.D. Gale, A. García, J. Junquera, P. Ordejón, D. Sánchez-
[18] A. Obuchi, A. Ohi, M. Nakamura, K. Ogata, K. Mizuno, H. Ohuchi, Appl. Catal. B Portal, J. Phys.: Condens. Matter 14 (2002) 2745–2779.
2 (1993) 71–80. [52] D.S. Portal, P. Ordejon, E. Canadell, Struct. Bonding 113 (2004) 103–170.
[19] M. Iwamoto, H. Yahiro, Catal. Today 22 (1994) 5–18. [53] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865–3868.
[20] K. Hata, D.N. Futuba, K. Mizuno, T. Numai, M. Yumura, S. Ijima, Science 306 [54] N. Troullier, J.L. Martins, Phys. Rev. B 43 (1991) 1993–2006.
(2004) 1362–1364. [55] F.A. Granja, A.G. Fuente, A. Vega, Phys. Rev. B 78 (2008) 134425/1-9.
[21] J. Xu, M. Saeys, J. Phys. Chem. C 112 (2008) 9679–9685. [56] R.C. Longo, M.M.G. Alemany, J. Ferrer, A. Vega, L.J. Gallego, J. Chem. Phys. 128
[22] Z.M. Liu, J.H. Li, S.I. Woo, Energy Environ. Sci. 5 (2012) 8799–8814. (2008) 114315/1-5.
[23] J.J. Yu, X.P. Wang, L.D. Li, Z.P. Hao, Z.P. Xu, G.Q. Lu, Adv. Funct. Mater. 17 (2007) [57] K. Park, Phys. Rev. B 83 (2011) 064423/1-9.
3598–3606. [58] F.A. Granja, M.B. Torres, A. Vega, L.C. Balbas, J. Phys. Chem. A 116 (2012)
[24] P. Nikolov, M. Khristova, D. Mehandjiev, Colloids Surf. A 295 (2007) 239–245. 9353–9360.
[25] Z. Shu, W. Huang, Z. Hua, L. Zhang, X. Cui, Y. Chen, H. Chen, C. Wei, Y. Wang, [59] A.R. Diaz, Y. Coronado, L.A. Pere, I.L. Garzon, Eur. Phys. J. D 52 (2009) 127–130.
X. Fan, H. Yao, D. He, J. Shi, J. Mater. Chem. A 1 (2013) 10218–10227. [60] M.B. Torres, F.A. Granja, L.C. Balbas, A. Vega, J. Phys. Chem. A 115 (2011)
[26] S.Y. Wu, J.J. Ho, Phys. Chem. Chem. Phys. 12 (2010) 13707–13714. 8350–8360.
[27] M. Stanciulescu, G. Caravaggio, A. Dobri, J. Moir, R. Burich, J.P. Charland, [61] R.G. Parr, R.G. Pearson, J. Am. Chem. Soc. 105 (1983) 7512–7516.
P. Bulsink, Appl. Catal. B: Environ. 123 (2012) 229–240. [62] M. Moskovits, J.E. Hulse, J. Chem. Phys. 66 (1997) 3988–3994.
[28] A. Rochefort, R. Fournier, J. Phys. Chem. 100 (1996) 13506–13513. [63] E.K. Parks, L. Zhu, J. Ho, S.J. Riley, J. Chem. Phys. 100 (1994) 7206–7222.
[29] A. Endou, N. Ohashi, K. Yoshizawa, S. Takami, M. Kubo, A. Miyamoto, J. Phys. [64] E.K. Parks, B.J. Winter, T.D. Klots, S.J. Riley, J. Chem. Phys. 94 (1991) 1882–1902.
Chem. B 104 (2000) 5110–5117. [65] Q.M. Ma, Z. Xie, Y. Liu, Y.C. Li, Chin. Phys. Lett. 24 (2007) 1908–1911.
[30] C.L. Dufaure, J. Roques, C. Mijoule, E. Sicila, N. Russo, V. Alexiev, T. Mineva, J. [66] Z. Xie, Q.M. Ma, Y. Liu, Y.C. Li, Phys. Lett. A 342 (2005) 459–467.
Mol. Catal. A: Chem 341 (2011) 28–34. [67] S.E. Apsel, J.W. Emmert, J. Deng, L.A. Bloomfield, Phys. Rev. Lett. 76 (1996)
[31] X. Ding, Z. Li, J. Yang, J.G. Hou, Q. Zhu, J. Chem. Phys. 121 (2004) 2558–2562. 1441–1444.
[32] V.E. Matulis, D.M. Palagin, A.S. Mazheika, O.A. Ivashkevich, Comp. Theor. Chem [68] M. Moseler, H. Häkkinen, R.N. Barnett, U. Landman, Phys. Rev. Lett. 86 (2001)
963 (2011) 422–426. 2545–2548.
[33] J. Wang, G. Fan, F. Li, RSC Adv. 2 (2012) 9976–9985. [69] B.V. Reddy, S.N. Khanna, B.I. Dunlap, Phys. Rev. Lett. 70 (1993) 3323–3326.
[34] J.W. Medlin, M.D. Allendorf, J. Phys. Chem. B 107 (2003) 217–223. [70] W. Tang, E. Sanville, G. Henkelman, J. Phys.: Condens. Matter 21 (2009)
[35] A.N. Andriotis, M. Menon, G.E. Froundakis, J.E. Lowther, Chem. Phys. Lett. 301 084204–084207.
(1999) 503–508.
[36] F. Baletto, R. Ferrando, Rev. Mod. Phys. 77 (2005) 371–423.
[37] P. Calaminici, J. Chem. Phys. 128 (2008) 164317/1-4.

You might also like