You are on page 1of 11

pubs.acs.

org/JPCC Article

A DFT Study of the Ammonia Decomposition Mechanism on the


Electronically Modified Fe3N Surface by Doping Molybdenum
Diksha Praveen Pathak, Anwesh Kumar, Balathanigaimani Marriyappan Sivagnanam,* A. S. K. Sinha,*
and J. Karthikeyan*
Cite This: J. Phys. Chem. C 2023, 127, 16442−16452 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Ammonia is a potential carbon-free carrier for the on-site delivery of hydrogen and is
critical in the development of a hydrogen economy. Designing a noble metal-free catalyst for the low-
Downloaded via SICHUAN UNIV on November 14, 2023 at 04:28:38 (UTC).

temperature decomposition of NH3 is a challenging task in this route. We conducted first-principles


studies to have a comprehensive understanding of the ammonia decomposition mechanism on the
molybdenum-doped iron nitride (Fe3N) surface. The results indicate that when Mo is doped on the
surface of Fe3N(111) it donates electrons to the surface and alters the overall electronic structure of the
catalyst. The activation energy for the intermediate steps of ammonia decomposition is substantially
reduced on the Mo-doped Fe3N surface compared to undoped Fe3N. NH3 adsorbs preferably on Mo sites
over Mo-doped Fe3N(111). However, subsequent intermediate species NH2 and NH prefer Fe sites for
adsorption. The activation barrier for the recombination and desorption of nitrogen adatoms is highest
compared to the other elementary steps in the breakdown of NH3 on Mo-doped Fe3N(111) and therefore
limits the overall rate of the decomposition reaction of ammonia. The Bader charge, density of state
(DOS), and crystal orbital Hamiltonian population (COHP) calculations elucidate the electronic
properties of the catalyst.

1. INTRODUCTION mechanism of ammonia decomposition typically begins with


Hydrogen is a promising alternative to fossil fuels in a the adsorption of NH3 on the surface of a catalyst, followed by
sustainable, carbon-free economy. However, finding a safe and the sequential removal of hydrogen resulting in the formation of
economical technology for the storage and on-site generation of N and H adatoms, which later recombine and then desorb from
hydrogen is a major challenge. For this purpose, ammonia is the catalyst surface as N2 and H2.9−13
proposed as a viable hydrogen vector for on-site hydrogen The rate-governing step for ammonia breakdown is still
ambiguous. However, regardless of the catalyst material used,
production due to the large gravimetric hydrogen density (17.8
there are two possible rate-limiting steps for the decomposition
wt %), easy storage, and well-developed transmission technol-
process of ammonia: N−H bond cleavage and recombination−
ogy.1,2 In the process of producing hydrogen from ammonia, the
desorption of N2 from the catalyst surface.12 One of the earliest
N−H bonds in the ammonia molecule have to be broken,
studies of NH3 decomposition on iron by Brauner et al. reported
followed by N−N bond formation and desorption. However,
N2 desorption as the rate-deciding step.14 In 1966, Takezawa et
without a catalyst, this molecular decomposition reaction is
al. studied the temperature dependence of the NH3 decom-
sluggish and occurs at very high temperatures. It is believed that
position mechanism over the iron catalyst. They found that at
without a catalyst ammonia decomposition cannot be
lower temperatures N2 desorption controls the overall rate while
accomplished below 1100 °C. However, the degree of
the dehydrogenation of any of the adsorbed NHx (x = 1−3)
degradation is dependent on the type of catalyst with which
fragments determines the rate at higher temperatures.15 Wang et
ammonia interacts.3 Several noble- and non-noble-metal-based
al. also confirmed in their studies of ammonia decomposition
catalysts have been studied for these reactions; among them, Ru-
that strongly bonded N adatoms control the reaction rate at
metal-based catalysts were found to be the most efficient for
lower temperatures.16 A theoretical study on ammonia
ammonia decomposition to produce hydrogen. However, the
decomposition over Fe(111) reported the dehydrogenation of
high cost and scarcity of Ru limit its utilization on a large scale.
Hence, it is important to find a nonprecious-metal-based catalyst
that can decompose the ammonia at low temperatures to make Received: June 22, 2023
the process of hydrogen production economically viable.4−8 It is Revised: July 30, 2023
critical to have detailed insight into the ammonia decomposition Published: August 9, 2023
mechanism for developing a less expensive catalyst that can
break down ammonia at low temperatures and understand why
ruthenium has better activity than other elements. The

© 2023 American Chemical Society https://doi.org/10.1021/acs.jpcc.3c04207


16442 J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

NH to N as the rate-determining step.17 NH3 breakdown over the fundamentals of the catalytic decomposition of ammonia on
Co(111), Fe(110), and Ni(111) surfaces has shown the Mo-doped Fe3N(111), explain how doping improves the
recombination reaction of N adatoms to be more energy costly catalytic activity, and provide insight into the development of
than any of the dehydrogenation steps.18 an efficient noble-metal-free catalyst for low-temperature
It has been proposed that transition-metal nitrides with ammonia decomposition.
properties like those of noble metals may efficiently operate as
catalysts in the processes of hydrodenitrogenation, Fischer− 2. COMPUTATIONAL METHODS
Tropsch synthesis, hydrodesulfurization, and the synthesis and The atomic structure and electronic structure of catalyst surfaces
breakdown of ammonia.19−22 Transition-metal nitrides have and various intermediate products of NH3 decomposition
high mechanical, thermal, and chemical stability, and they can reactions are studied using plane wave-assisted density func-
transfer electrons from the nitrogen atom.23 The nitrogen in tional theory calculations (DFT) as implemented in the Vienna
metal nitrides, in general, is present in the interstitial spaces Ab Initio Simulation Package.36−38 A plane-wave energy cutoff
between the large metal atoms, altering the surface’s electronic of 500 eV is fixed in all of the calculations. The electron−ion
characteristics. This lowers the energy of activation for ammonia potentials are defined using PAW pseudopotentials, and the
decomposition on nitride surfaces.24 Also, the surface (lattice) electron−electron exchange and correlation energy are
nitrogen in nitrides participates in the rate-determining N2 corrected using the Perdew, Burke, and Ernzerhof (PBE)
desorption step.25 The studies conducted on the decomposition functional39 described within generalized gradient approxima-
mechanism of ammonia on Mo2N and MoN surfaces have tions. The VASP computed data was further analyzed using
reported that the recombinative desorption process limits VASPKIT, and the crystal structures were illustrated using
overall ammonia decomposition.26,27 The rate of the catalytic VESTA.40,41 All of the atoms in the bulk Fe3N structure are
breakdown of ammonia into iron nitrides increases with an relaxed without any constraints, while for the surface models, the
increase in the concentration of ammonia. As compared to clean lattice parameters of bulk structures are fixed and the atoms of
iron, the activation energy for ammonia decomposition is adsorbed molecules are relaxed completely. For the electronic
approximately 2 times lower on iron nitrides.28 The lattice and ionic relaxations, tight convergence criteria of 10−6 and 0.01
nitrogen atoms in transition-metal nitrides assist the catalytic eV/Å, respectively, are fixed in all of the calculations. The
formation and breakdown of NH3 by following a Mars−van Brillouin zone was sampled using 8 × 8 × 1 Monkhrost-Pact k-
Krevelen mechanism.29−32 Iron nitride (Fe3N) has been mesh.42 A 15 Å vacuum space in the z direction was used. The
reported to be a promising catalyst in ammonia breakdown surface energy per Fe3N unit per unit area (Es) of various surface
and a potential alternative for scarce and costly noble-metal- slabs is defined as
based catalysts. Schnepp and co-workers reported hydrogen
generation via the ammonia breakdown reaction on the surface [E(surface) m × E(bulk Fe3N)]
of the Fe3N catalyst at a gas hourly space velocity of 15 000 cm3 Es =
[2A] (1)
gcat−1 h−1. These results demonstrate that ammonia decom-
position begins at 500 °C and reaches 100% decomposition at where E(surface) is the total energy of the surface slab, E(bulk‑Fe3N) is d

800 °C. The catalyst’s stability at 600 °C improves gradually the total energy per Fe3N formula unit of bulk Fe3N, A is the
over the observation time.33 Further investigations on the surface area of the slab, and m is the number of Fe3N units
surface reactions as well as the recombination and desorption of present in this slab. The adsorption energy Eads values were
adsorbed N atoms are required to have a full understanding of obtained by utilizing the given equation
the decomposition process of ammonia over Fe3N catalysts.
In this study, we have conducted a comprehensive theoretical Eads = E(ads *+surface) E(ads*) E(surface) (2)
investigation of the NH3 decomposition mechanism on the Mo-
doped Fe3N(111) catalyst surface using plane wave-based where E(ads*) and E(surface) are the total energies of the gaseous
density functional theory calculations. The decomposition of adsorbate and the pristine surface, respectively, and the
NH3 over Fe3N occurs at very high temperatures, which limits its combined adsorbate/surface system is represented as
utilization for decomposing ammonia. To overcome this E(ads*+surface). The reaction energy (Erxn) is equivalent to the
limitation, we suggest that molybdenum doping in Fe3N as energy of the product minus the energy of the reactant. The
among transition-metal nitrides Mo2N is considered to be best formation energy of Mo-doped Fe3N(111) by the substitution
and equivalent to platinum in terms of its catalytic activity for of Mo on pristine Fe3N(111) is calculated using the given
ammonia decomposition.34 The presence of electron-donating formula
species improves the catalytic activity by balancing the charge of formation energy = {E(Mo Fe3N) + E(Fe)} {E(Fe3N) + E(Mo)}
the intermediate steps of ammonia breakdown.12,35 Our study (3)
shows that Mo transfers the electron to the surface and acts as
the most stable site for NH3 adsorption. We utilized DFT where E(Mo‑Fe3N) and E(Fe3N) are the total energies of Mo-doped
d d

calculations to determine the adsorption sites and configurations Fe3N(111) and pristine Fe3N(111), respectively, E(Fe) is the
of all of the reactive intermediates that are involved in the energy of a single Fe which will be substituted by Mo on pristine
process of decomposing ammonia over Mo-doped Fe3N(111). Fe3N(111), and E(Mo) is the energy of a single Mo obtained from
The role of the lattice nitrogen atom in the ammonia either the Mo dimer (Mo-poor limit) or Mo bulk (Mo-rich
decomposition is included. We also calculated the adsorption limit).
energies, reaction energies, and activation barriers for the We used the climbing image nudged elastic band model (CI-
sequential dehydrogenation steps of NH3 and associative N NEB) to calculate the activation barriers and transition states for
desorption. The electronic effect of doping Mo on Fe3N has the ammonia breakdown process on the Mo-doped Fe3N-
been demonstrated by the density of states and the COHP (111).43 The density of states analysis was performed to examine
analysis. The objectives of the current work are to understand the electronic structure of the catalyst surface. We evaluated
16443 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 1. Optimized structures of (a) Fe3N; (b) cleaved Fe3N(111), where Fe1 and Fe2 represent iron coordinated with one and two nitrogens,
respectively, B and H represent bridging and hollow sites, respectively, on Fe3N(111); and (c) Mo-doped Fe3N(111), where B1, B2, and B3 are the
bridging sites and H1, H2, and H3 are the hollow sites on Mo-doped Fe3N(111). (d) Formation energy plot of Mo-doped Fe3N(111) (orange = Fe,
blue = N, and cyan = Mo).

crystal orbital Hamiltonian population (COHP) observations XRD peak.33 Hence, we selected the (111) surface for further
using LOBSTER.44,45 studies of the ammonia decomposition reaction. The optimized
structure of the (111) slab is shown in Figure 1(b), and there are
3. RESULTS AND DISCUSSION two types of iron present on the surface, that is, Fe1 (iron
attached with one nitrogen atom) and Fe2 (iron attached with
3.1. Optimized Structure and Charge-Transfer Study two nitrogen atoms). B is the bridging site between two Fe1
of the Catalyst Surface. The optimized structure of pristine atoms, and H is the hollow site formed by one Fe2 and two Fe1
Fe3N is shown in Figure 1(a), and nitrogen atoms are present in atoms. Furthermore, the Fe3N(111) surface is doped with Mo,
one-third of the octahedral voids formed by Fe atoms with an and Mo prefers to substitute the Fe2 atom on the Fe3N(111).
ABAB...-type stacking, leaving 2/3 of the voids empty. To check The optimized structure of Mo-doped Fe3N(111) is shown in
the catalytic activity on different crystalline surfaces, the Fe3N Figure 1(c). There are mainly two kinds of adsorption sites
bulk structure was cleaved along primary crystal planes such as available on this doped surface, namely, bridging sites B1, B2,
(100), (001), (110), and (111). The surface energies of these and B3 and hollow sites H1 and H2. The bridging site B1 is
planes are calculated using the formula given in the Computa- present between Fe1 and Fe2, B2 is located above Fe1 and Mo,
tional Methods section, and the values are reported in Table 1. and B3 is between Fe2 and Mo. The hollow site H1 is
Among these planes, the (111) surface is the most stable, which coordinated with three Fe1 atoms, H2 is formed between two
is in accordance with a previous experimental investigation on Fe1 atoms and one Mo atom, and H3 is formed between Fe1,
the Fe3N catalyst, where the (111) plane has the most intense Fe2, and Mo atoms.
We calculated the formation energy for Mo-doped Fe3N(111)
Table 1. Surface Energy of the Different Planes of Fe3N to confirm its stability. Figure 1(d) shows the plot between the
formation energy of Mo-doped Fe3N(111) and the Mo chemical
Plane Surface Energy/Unit Area (eV/Å2/Fe3N)
potential. The formation energy is negative in the region with a
111 0.21 Mo-poor limit and becomes almost zero in the region of a Mo-
100 0.23 rich limit. This explains that during the formation of Mo-doped
001 0.25 Fe3N(111) by the reaction of Mo and pristine Fe3N(111), when
110 0.26 there is a small amount of Mo present in the atmosphere (i.e., if
16444 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 2. Differential density plots of Mo-doped Fe3N(111). Isosurface levels (a) ±0.004 e−/Å3 and (b) ±0.00008 e−/Å3. Pink represents the charge
buildup region, and green represents the charge depletion region remaining Color code: similar to that in Figure 1.

Figure 3. Optimized adsorption configurations for NHx (x = 0−3) and a hydrogen atom on the Fe3N(111) surface. Here, “v” represents the vertical
adsorption geometry of N2, and “p” represents the parallel adsorption geometry of N2. All bond lengths are in Å (white = H; remaining color code is
similar to that in Figure 1).

the Mo comes from a Mo dimer), it can easily substitute the Fe that formation of Mo-doped Fe3N is less favorable in a Mo-rich
from pristine Fe3N(111) and form a stable Mo-doped Fe3N. environment.
However, if the atmosphere is Mo-rich or Mo comes from Mo- In order to elucidate the effect of Mo doping on the electronic
bulk, then the formation energy becomes almost zero, indicating structure of the (111) surface, we have performed the worse
16445 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 4. Optimized adsorption configurations for NHx (x = 0−3) and a hydrogen atom on the Mo-doped Fe3N(111) surface. Here, “v” represents the
vertical adsorption geometry of N2, and “p” represents the parallel adsorption geometry of N2. All bond lengths are in Å (white = h; remaining color
code is similar to that in Figure 3).

charge calculations. The calculated Bader charge revealed that Table 2. Adsorption Characteristics of the Optimized
the doped molybdenum atom on the catalyst surface transfers Reaction Intermediates in Ammonia Decomposition on
electrons to the surface, which is clearly evident from the before Fe3N(111)
and after doping Bader charge values of Mo which are 6e− and
Adsorbate Adsorption Sites Ead (eV) dN−M (Å) dN−H (Å)
4.7e−, respectively. Furthermore, to confirm the same, the
NH3 Top(Fe2) −0.56 2.13 1.02
differential charge density plots are demonstrated in Figure 2. In
NH2 B(Fe1, Fe1) −3.18 1.95, 2.00 1.02
the surface plot, the electron-rich region is covered with a pink
NH H(1 Fe2 and 2 Fe1) −5.00 1.83−1.9 1.02
isosurface, and the electron-deficient region is covered with a
N H(1 Fe2 and 2 Fe1) −6.13 1.76−1.82
green isosurface. It is clear from the isosurface plot that Mo is
N2(v) Top(Fe2) −0.39 1.81
transferring electron density to the nearby Fe atoms while it itself N2(p) Top(Fe2) 0.34 2.07, 2.08
is becoming electron-deficient. Similar to the earlier studies of H H(1 Fe2 and 2 Fe1) −0.69
charge transfer between Ru and CNT in the Ru/CNT system,
this charge transfer may aid in better adsorption and in lowering Table 3. Adsorption Characteristics of the Optimized
activation barriers on the surface of the catalyst, hence Reaction Intermediates in Ammonia Decomposition on Mo-
enhancing the overall catalytic activity.46−48 Doped Fe3N(111)
3.2. Preferred Adsorption Site of Reaction Intermedi-
ates. Adsorption energy calculations are performed to obtain Ead dN−H
Adsorbate Adsorption Sites (eV) dN−M (Å) (Å)
the most preferred adsorption configurations of the reactive
intermediates in the ammonia decomposition process. NHx (x = NH3 Top(Mo) −1.11 2.26 1.02
0−3) species and the hydrogen atom are adsorbed at various NH2 B1(Fe1, Fe2) −3.22 1.93, 1.99 1.02
NH H1(3 Fe1) −4.96 1.83−1.9 1.02
positions over Fe3N(111) and Mo-doped Fe3N(111). The
N H2(2 Fe1 and 1 Mo) −6.3 1.82−1.9
optimized geometries of preferred configurations on Fe3N(111)
N2(v) Top(Mo) −0.85 2.06
and Mo-doped Fe3N(111) are represented in Figures 3 and 4,
N2(p) Top(Mo) + B2(Fe1, Mo) −0.86 1.83−2.22
respectively, and their corresponding adsorption energies and
H Hollow(H2) −0.77
bond distances are summarized in Tables 2 and 3.
On an optimized Fe3N(111) surface, the NH3 adsorbs
preferably on the Fe2 site with an adsorption energy of −0.56 eV (Figure 3 ii(a) and ii(b)) with Eads = −3.18 eV and Erxn = 0.01
(Figure 3 i(a) and i(b)). However, the other reactive eV. The subsequent dehydrogenated species NH prefers the
intermediates NH2, NH, and N prefer the Fe1 site. NH2 is hollow site H formed by one Fe2 and two Fe1 atoms (Figure 3
adsorbed at a bridging position B between two iron atoms iii(a) and iii(b)), where Eads = −5 eV and Erxn = 0.37 eV. After
16446 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 5. Optimized adsorption geometries of Nlat−NHx (x = 0−2) on the Mo-doped Fe3N(111) surface. The upper and lower panels show top and
side views, respectively. (The color code is similar to that in Figure 4.)

the removal of H from NH, the nitrogen atoms remain adsorbed distances are 1.82 and 1.90 Å, and the Mo−N bond length is
at the hollow site H (Figure 3 iv(a) and iv(b)) with Eads = −6.13 1.91 Å (Figure 4 iv(a) and iv(b)). Clearly, the adsorption energy
eV and Erxn = 0.70 eV. N2 adsorbs in a perpendicular increases as the number of H atoms decreases from NHx (x = 0−
configuration with an adsorption energy value of 0.34 eV, 3) species since the total number of metal−nitrogen
suggesting that N2 does not favor the perpendicular config- coordinations increases and the metal−nitrogen bond distance
uration (Figure 3 v(a) and v(b)). The N2 species is preferably gets shorter.51 The adsorbed N atoms then combine to form N2.
adsorbed in a vertical configuration on top of Fe2 with an Molecular N2 may adsorb in either a vertical or a perpendicular
adsorption energy of −0.39 eV (Figure 3 vi(a) and vi(b)). The configuration on the catalyst surface. N2 adsorbs in vertical
H atom is also adsorbed at hollow site H with Eads = −0.69 eV configurations with an adsorption energy of −0.85 eV on the Mo
(Figure 3 vii(a) and vii(b)). top with a Mo−N bond length of 2.06 Å and a N−N bond length
On the surface of Mo-doped Fe3N(111), NH3 adsorbs very of 1.13 Å, which is almost equivalent to that of N2(g) (Figure 4
weakly on top of Fe1 (Eads = −0.10 eV) and does not adsorb over v(a) and v(b)). However, N 2 favors a perpendicular
Fe2. Ammonia preferentially adsorbs over Mo with an configuration with the adsorption energy equivalent to −0.86
adsorption energy value of −1.11 eV which is in accordance eV (Figure 4 vi(a) and vi(b)). Here the N2 molecule adsorbs in
with one previously reported DFT study on the Mo2N(111) such a way that both N atoms are adsorbed on top of Mo but one
surface having approximately similar NH3 adsorption energy.26 N is in a bridging position between Mo and Fe atoms with an
As evident from the adsorption energy values, the binding M−N bond length in the range of 1.83−2.22 Å and a N−N bond
affinity of NH3 on the catalyst surface is almost doubled by length of 1.2 Å, which is marginally longer than that of N2 in its
doping Mo on the surface of undoped Fe3N. The adsorbed NH3 gaseous form. Some other possible adsorption configurations of
has a Mo−N bond length of 2.26 Å, and the N−H bond distance the intermediates in the NH3 decomposition reaction have been
is 1.02−1.03 Å, similar to that of free ammonia (Figure 4 i(a) represented in Figure S2, and their corresponding adsorption
and i(b)).49 Here, Mo helps in the activation of the N−H bond energy values are summarized in Table S1. The H atoms that are
of NH3 to initiate the decomposition reaction. The N−H bond generated during the dehydrogenation steps in the NH3
activation of ammonia has been further analyzed with the help of decomposition process prefer to get adsorbed over a hollow
a crystal orbital Hamiltonian population (COHP) analysis in site constituting two Fe atoms and one Mo atom with an
section 3.4. In the next step of the decomposition process, first adsorption energy of −0.77 eV (Figure 4 vii(a) and vii(b)).
the deprotonation of NH3 takes place, leaving NH2 on the Another stable site for H adsorption is on the bridge between
surface. NH2 adsorbs on a bridging site (B1) between two iron two iron atoms, having an adsorption energy of −0.59 eV
atoms with an adsorption energy of −3.22 eV and Erxn = 0.22 eV. (Figure 4 viii(a) and viii(b)).
Here, the bridged NH2 has N−Fe bond distances of 1.93 and Reactive intermediates generated during the catalytic break-
1.99 Å, and the N−H bond distance is 1.02 Å (Figure 4 ii(a) and down of ammonia can also adsorb at the lattice nitrogen (Nlat)
ii(b)). In the subsequent step, after the second deprotonation, sites. After the dehydrogenation of ammonia at the Mo site, the
the NH will be generated on the surface. NH prefers a hollow resultant NH2 may move to the nearby lattice nitrogen atom and
site (H1) for adsorption which is formed by three irons with an form Nlat−NH2 (Figure 5 ia and ib). The adsorption energy for
adsorption energy value of −4.96 eV, and the Fe−N bond this attachment of NH2 at Nlat is −0.82 eV, which is much lower
lengths are 1.83, 1.85, and 1.9 Å. The N−H bond distance is 1.02 than that at its other three adsorption sites (B1, B2, and B3) on
Å, as shown in Figure 4 iii(a) and iii(b). In the last step of NH3 the catalyst surface (Table S1). Later, the dehydrogenation of
breakdown, NH is deprotonated, and an N atom is formed on Nlat−NH2 will generate Nlat−NH over the surface of the catalyst.
the catalyst surface. This N atom is preferentially adsorbed over Transit of the NH molecule adsorbed at the hollow site H1
the hollow space H3 coordinated with two Fe atoms and one Mo (Figure 4 iiia and iiib) toward the neighboring Nlat may also
atom with an adsorption energy value of −6.30 eV, which is generate Nlat−NH. Afterward, the NH migrates and gets
lower than those of both Fe(100) (Eads = −7.35 eV) and adsorbed in the hollow site formed by Fe, Mo, and Nlat (Figure
Mo2N(111) (Eads = −7.17 eV) and equivalent to that of 5 iia and iib) with an adsorption energy value equivalent to
Mo2N(100) (Eads = −6.29 eV).26,50 Here the Fe−N bond −3.16 eV and Erxn = 1.91 eV. Further dehydrogenation will result
16447 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

in the formation of Nlat−N. The nitrogen atom formed after the activation barrier is 0.14 eV, which is significantly lower than in
dehydrogenation of adsorbed NH is settled at a hollow site the Fe3N case, and it is exothermic by −0.3 eV. As represented in
coordinated with Nlat, Fe, and Mo (Figure 5 iiia and iiib). The Figure 6(b), in the first transition state (TS1) NH2 is adsorbed
adsorption energy for this step is −4.47 eV, and the reaction over the Mo top and H settles between Mo and the nearby Fe
energy is endothermic by 2.01 eV. This elucidates that the atom. Here the bond distance between N and H is 1.86 Å. After
formation of Nlat−N on the catalyst surface is challenging and the first dehydrogenation, NH2 adsorbed at Fe1 on Fe3N(111)
energy-intensive. is further dehydrogenated into NH and H, and the
3.3. Decomposition Reactions of Ammonia. We corresponding transition state for the same is shown in Figure
investigated the reaction pathway of ammonia decomposition 6(c). Here, both the NH and H are present at the bridging
on Fe3N(111) and Mo-doped Fe3N(111) and calculated the position between two irons, and the N−H bond length is 1.02 Å.
activation barriers for each step involved in the decomposition The activation barrier for this second dehydrogenation step on
process. Figure 6 illustrates the potential energy profile for the Fe3N(111) is 0.50 eV and is minutely exothermic by −0.01 eV.
The NH2 adsorbed at Mo on Mo-doped Fe3N(111) moves to its
most stable adsorption site, which is the bridging site B1 with a
diffusion barrier of 0.60 eV, and H migrates toward hollow site
H2. The second dehydrogenation starts with NH2 being
stabilized at the bridging position between two iron atoms and
is then dissociated into NH which settles over the hollow site
(H1) and hydrogen at the hollow site (H2) interacting with two
irons and one molybdenum. This disintegration reaction has an
activation barrier of 0.33 eV and is exothermic by −0.62 eV. In
the second transition state, NH adsorbs among three iron atoms;
meanwhile, H settles over one of the nearby iron atoms, forming
the hollow site (H1) with an N−H bond length of 1.34 Å
(Figure 6(d)). During the last step of dehydrogenation, NH is
present at the hollow site (H) on Fe3N(111) and decomposes to
N, which remains intact at the hollow site, and H migrates to the
bridging position. Figure 6(e) represents the third transition
state corresponding to the last dehydrogenation step on
Fe3N(111) here; the N is present at the void among three
iron atoms, and H is transitioning to the bridging position. The
activation barrier for this step is 0.58 eV and endothermic by 0.3
eV. On Mo-doped Fe3N, the NH adsorbed at the hollow site is
further dissociated into N and H. The N stays at the hollow site,
Figure 6. Schematic representation of the reaction pathway for and H moves to the bridging position between two iron atoms.
ammonia decomposition on Fe3N(111) and Mo-doped Fe3N(111) via The energy barrier for this pathway is 0.77 eV, which is the
Mo and Fe atoms. Here, TS = transition state and N* = adsorbed highest among all of the dehydrogenation steps and is slightly
nitrogen. The energies are calculated using the pristine surface and two endothermic by 0.02 eV. For the third transition state, N is
NH3 molecules as reference. The color code is similar to that in Figure
adsorbed in the hollow space among three iron atoms, and H
4.
attaches to the nearby iron atom (Figure 6(f)). Then the N
moves to its most stable adsorption configuration at the hollow
sequential dehydrogenation processes of ammonia on Fe3N- site (H2), and H is stabilized on the hollow site (H3)
(111) and Mo-doped Fe3N(111) including the optimized coordinated with the molybdenum atom and two iron atoms.
configurations of the transition states. On Fe3N(111), the Figure 7 depicts the reaction paths of N2* production on Mo-
decomposition process begins with the adsorption of NH3 on doped Fe3N(111), along with the transition states. For the
top of Fe2. Afterward, the adsorbed ammonia is dissociated into activation barrier calculation, we placed two nitrogen atoms
NH2, which remains adsorbed at Fe2, and the H is adsorbed on adjacent to each other and then optimized the configuration.
nearby surface nitrogen. The adsorbed NH2 then moves from The activation energy for this pathway is 2.23 eV, which is
the Fe2 to the Fe1 site, and H moves to the nearby hollow site approximately 0.7 eV higher as compared to that of Ru13@
formed by three iron atoms. The activation barrier for this CNT.46 However, when compared to the N2 formation barrier
dissociation is 1.82 eV, which is significantly high, suggesting on Mo2N(100) [Ebarrier = 3.59 eV (2N*-Diff + Diff-TS1)] it is
that the first dehydrogenation on Fe3N(111) is energy-extensive 1.36 eV less and on Mo2N(111) [Ebarrier = 2.76 ev (2N*-Diff +
and unfavorable. The transition state corresponding to this first Diff-TS1)] it is lower by 0.53 eV.26 The activation barrier for the
dehydrogenation step TS1 is represented in Figure 6(a) where N2* formation is much higher as compared to that of other
NH2 is adsorbed between two irons with an N−H bond distance dehydrogenation steps involved in the ammonia breakdown
of 1.04 Å, and hydrogen is adsorbed at the nearby nitrogen. For over Mo-doped Fe3N(111) and limits the overall reaction.
the Mo-doped Fe3N case, NH3 is preferentially adsorbed on the 3.4. Density of States and COHP Analysis. The effect of
Mo site, which is consistent with the earlier theoretical Mo doping on the electronic structure of the catalyst and its role
investigations of the preferential adsorption of ammonia at the during the NH3 decomposition reaction are clearly explained by
atop site on transition metals.5,18,26,51−54 Furthermore, the the density of states and d-band center analysis. We investigated
adsorbed NH3 is dehydrogenated and the dissociated H atom the d-band center energies of the d orbitals for the two
migrates to a bridging position between two iron atoms, while adsorption sites of ammonia on the catalyst surface, Mo and Fe1,
NH2 remains intact at the Mo site. For this pathway, the to examine their electronic structure. The d-band method by
16448 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

ammonia, the d bands are stretched from −10 to 3.5 eV, which
means that the d bands are moving away from the Fermi level
after the donation of a lone pair of electrons from ammonia to
the empty d orbitals of molybdenum, indicating a strong Mo−
NH3 interaction. However, for the Fe1 site, the d bands are a bit
shifted toward the Fermi level after NH3 adsorption from −8 to
6 eV, resulting in weak Fe−NH3 interaction and low adsorption
energy as perceived in our calculations. For Mo, both the up and
down spin states are present near the Fermi level, so ammonia
can easily donate its lone pair of electrons to the available up and
down d orbitals of Mo and will be strongly adsorbed. However,
for Fe1 only down spin states are available near the Fermi level,
which will make it difficult for ammonia to donate its lone pair,
resulting in a weaker Fe1−NH3 interaction (Figure 8). From
Table S1, we can observe that the adsorption energies are on the
Figure 7. Potential energy diagram of N2 formation on Mo-doped order of Eads,Mo > Eads,Fe1, which is consistent with the availability
Fe3N(111) including the transition state. The color code is similar to of up-spin states of molybdenum and iron near the Fermi level.
that in Figure 2. The COHP plots for NH3 adsorbed at Mo and Fe1 sites are
represented in Figure 9(a) and (b), respectively. It is clear that in
Hammer and Norskov is used to determine the projected
density of states.55 NH3 has a lone pair of electrons having both
up and down spins. Figure 8 shows the PDOS plots of both
molybdenum and iron sites before and after ammonia
adsorption. Before the Mo−NH3 binding, Mo shows d bands
in the range of −8 to 3 eV, whereas after coupling with the

Figure 9. Crystal orbital Hamiltonian population (COHP) plots of (a)


N*-Mo and (b) N*-Fe1 (N* is the N of adsorbed NH3, UP = up spin,
DW = down spin, and εf = Fermi level).

the case of N*−Fe1, antibonding states are slightly more


occupied near the Fermi level while bonding states are more
occupied in the case of N*−Mo. Here, N* is the N of NH3
adsorbed on the surface. More occupied antibonding states
make the molecule unstable, while more occupied bonding
states favor the stability of the molecule. Thus, the adsorption of
NH3 on the Mo site will be more stable and stronger as
compared to that on the Fe1 site. The DOS and COHP analyses
elucidate the strong adsorption of NH3 at Mo as compared to
the iron sites.
The arrangement of Fe d and N p states on the Fe3N(111)
surface (Figure 10(a)) was significantly skewed by the overlap of
the Fe d and N p states after Mo doping (Figure 10(b)). The
substantial transfer of the charge from Mo to the surface Fe and
N, which was shown in the Bader charge calculation, is also
reasonably supported by this observation. The iron d states and
nitrogen p states show a small overlap in Fe3N(111), although
after Mo doping, a significant overlap between the Fe d and N p
states is clearly visible as most of the N p states which were
Figure 8. d-orbital projected density of states for Fe1, Mo (before NH3 earlier present between −2 and −4 eV (Figure 10(a)) are now
adsorption) and Fe1*, Mo* (after NH3 adsorption). Up and down shifted to between −6 and −8 eV (Figure 10 (b)). The Mo d and
arrows represent up and down spin states, and the Fermi level is N p states also show a notable interaction with each other. This
represented by a dotted line. may be a sign of hybridization and bonding between Mo and the
16449 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

analysis, density of states, and COHP calculations clearly explain


the preferable adsorption of NH3 at the Mo site as observed in
this study.
In Figure 11, the COHP plots for the N−H bond of (a)
isolated gaseous NH3 and (b) NH3 adsorbed on the Mo top
have been plotted. For the N−H bond of isolated NH3(g),
antibonding peaks do not occur near the Fermi level, suggesting
a strong N−H interaction. However, for N−H bond of NH3
adsorbed on the Mo site, the occurrence of some of the
antibonding peaks near the Fermi level suggests the destabiliza-
tion and activation of the N−H bond.57

4. CONCLUSIONS
We theoretically investigated the Mo-doped Fe3N(111) surface
for NH3 decomposition. Our results showed that doping
molybdenum in the iron nitride increases its catalytic efficiency.
The presence of Mo on the catalyst surface offers two main
benefits: (i) NH3 has a strong binding tendency for Mo (Eads =
−1.11 eV on Mo-doped Fe3N) as compared to Fe (Eads = −0.56
eV on Fe3N), which is crucial for the reaction to begin, and (ii)
Mo donates electrons to the catalyst surface that helps in the
rate-controlling step of N2 recombination and desorption. The
activation barrier is highest for the nitrogen recombination step,
and it limits the overall ammonia decomposition reaction.
However, the comparatively lower activation barrier (Ebarrier =
2.23 eV for nitrogen recombination) and adsorption energy
(Eads = −6.30 eV for nitrogen) values observed in this study
indicate the enhanced potential of the Mo-doped Fe3N catalyst
compared to those of other well-known catalysts such as
Figure 10. DOS of all of the Mo(d), Fe(d), and N(p) orbitals of (a) Fe(100) and Mo2N for NH3 breakdown.26,50 This may be due to
Fe3N(111) and (b) Mo-doped Fe3N(111). Here the curly bracket from the presence of two different active metal sites: Mo, which
−6 eV to −8 eV represents the regions of small overlap between Fe(d) strongly adsorbs NH3 while NH2 and NH intermediate species
and N(p), before Mo doping shown in (a) and a significant overlap after are adsorbed at iron sites, and N, which prefers combined Fe−
Mo doping shown in (b). Up and down arrows represent up and down Mo sites. Similar to our observations, an experimental study also
spin states, and the Fermi level is represented by a dotted line. demonstrated that the combined iron−molybdenum catalyst
Fe3Mo3N has better activity than Mo2N for the ammonia
Fe3N(111) surface. The significant interaction among Mo, Fe, decomposition reaction due to the presence of a binary Fe−Mo
and N may further enhance the strong Mo−NH3 interaction as site.25 The preferable adsorption of ammonia at the Mo site is
well as increase the adsorption energy.56 The presence of a more clearly explained by the DOS and COHP calculations. Bader
positive charge on Mo observed in the charge-transfer charge calculations and charge-transfer analysis offer insight into
calculation (Figure 2) also suggests the strong adsorption of the electronic modification of the catalyst surface due to Mo
NH3 at the Mo site compared to iron. Thus, charge-transfer doping. Hence, it is proposed that Mo-doped Fe3N(111) may

Figure 11. Crystal orbital Hamiltonian population (COHP) plots of (a) N−H of NH3(g) and (b) N−H of NH3* at Mo (NH3* is the adsorbed NH3,
UP = up spin, DW = down spin, and εf = Fermi level).

16450 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

act as a sustainable and efficient catalyst for carbon-free H2 (5) Novell-Leruth, G.; Valcárcel, A.; Pérez-Ramírez, J.; Ricart, J. M.
generation through ammonia decomposition. Ammonia Dehydrogenation over Platinum-Group Metal Surfaces.
Structure, Stability, and Reactivity of Adsorbed NHx Species. J. Phys.
■ ASSOCIATED CONTENT
* Supporting Information

Chem. C 2007, 111 (2), 860−868.
(6) Yin, S. F.; Xu, B. Q.; Zhou, X. P.; Au, C. T. A mini-review on
ammonia decomposition catalysts for on-site generation of hydrogen
The Supporting Information is available free of charge at for fuel cell applications. Appl. Catal. A Gen. 2004, 277 (1−2), 1−9.
https://pubs.acs.org/doi/10.1021/acs.jpcc.3c04207. (7) Klerke, A.; Christensen, C. H.; Nørskov, J. K.; Vegge, T. Ammonia
Optimized structure of different planes of the catalyst for hydrogen storage: Challenges and opportunities. J. Mater. Chem.
surface; adsorption sites of reaction intermediates other 2008, 18 (20), 2304−2310.
than the most stable sites; PDOS plot of Mo-doped (8) Yu, P.; Wu, H.; Guo, J.; Wang, P.; Chang, F.; Gao, W.; Zhang, W.;
Liu, L.; Chen, P. Effect of BaNH, CaNH, Mg3N2 on the activity of Co
Fe3N(111); and zero point energy-corrected adsorption
in NH3 decomposition catalysis. J. Energy Chem. 2020, 46, 16−21.
energy values of the reactive intermediates (PDF) (9) Ganley, J. C.; Thomas, F. S.; Seebauer, E. G.; Masel, R. I. A Priori

■ AUTHOR INFORMATION
Corresponding Authors
Catalytic Activity Correlations: The Difficult Case of Hydrogen
Production from Ammonia. Catal. Lett. 2004, 96 (3), 117−122.
(10) Yin, S.-F.; Zhang, Q.-H.; Xu, B.-Q.; Zhu, W.-X.; Ng, C.-F.; Au, C.-
Balathanigaimani Marriyappan Sivagnanam − Department of T. Investigation on the catalysis of COx-free hydrogen generation from
ammonia. J. Catal. 2004, 224 (2), 384−396.
Chemical and Biochemical Engineering, Rajiv Gandhi Institute
(11) García-García, F. R.; Ma, Y. H.; Rodríguez-Ramos, I.; Guerrero-
of Petroleum Technology, Amethi 229304 Uttar Pradesh, Ruiz, A. High purity hydrogen production by low temperature catalytic
India; orcid.org/0000-0001-5862-6215; ammonia decomposition in a multifunctional membrane reactor. Catal.
Email: msbala@rgipt.ac.in Commun. 2008, 9 (3), 482−486.
A. S. K. Sinha − Department of Chemical and Biochemical (12) Schüth, F.; Palkovits, R.; Schlögl, R.; Su, D. S. Ammonia as a
Engineering, Rajiv Gandhi Institute of Petroleum Technology, possible element in an energy infrastructure: Catalysts for ammonia
Amethi 229304 Uttar Pradesh, India; Email: asksinha@ decomposition. Energy Environ. Sci. 2012, 5 (4), 6278−6289.
rgipt.ac.in (13) Lamb, K. E.; Dolan, M. D.; Kennedy, D. F. Ammonia for
J. Karthikeyan − Department of Sciences & Humanities, Rajiv hydrogen storage; A review of catalytic ammonia decomposition and
Gandhi Institute of Petroleum Technology, Amethi 229304 hydrogen separation and purification. Int. J. Hydrogen Energy. 2019, 44
Uttar Pradesh, India; Department of Physics, National (7), 3580−3593.
(14) Brunauer, S.; Love, K. S.; Keenan, R. G. Adsorption of Nitrogen
Institute of Technology, Durgapur 713209 West Bengal,
and the Mechanism of Ammonia Decomposition Over Iron Catalysts. J.
India; orcid.org/0000-0002-1781-8357; Am. Chem. Soc. 1942, 64 (4), 751−758.
Email: karthikeyan.jeyakumar@phy.nitdgp.ac.in (15) Takezawa, N.; Toyoshima, I. The Change of the Rate-
Determining Step of the Ammonia Decomposition over an Ammonia
Authors
Synthetic Iron Catalyst. J. Phys. Chem. 1966, 70 (2), 594−595.
Diksha Praveen Pathak − Department of Chemical and (16) Wang, L.; Zhao, Y.; Liu, C.; Gong, W.; Guo, H. Plasma driven
Biochemical Engineering, Rajiv Gandhi Institute of Petroleum ammonia decomposition on a Fe-catalyst: eliminating surface nitrogen
Technology, Amethi 229304 Uttar Pradesh, India; poisoning. Chem. Commun. 2013, 49 (36), 3787−3789.
orcid.org/0000-0002-1321-0926 (17) Lin, R.-J.; Li, F.-Y.; Chen, H.-L. Computational Investigation on
Anwesh Kumar − Department of Chemical and Biochemical Adsorption and Dissociation of the NH3Molecule on the Fe(111)
Engineering, Rajiv Gandhi Institute of Petroleum Technology, Surface. J. Phys. Chem. C 2011, 115 (2), 521−528.
Amethi 229304 Uttar Pradesh, India (18) Duan, X.; Ji, J.; Qian, G.; Fan, C.; Zhu, Y.; Zhou, X.; Chen, D.;
Yuan, W. Ammonia decomposition on Fe(1 1 0), Co(1 1 1) and Ni(1 1
Complete contact information is available at: 1) surfaces: A density functional theory study. J. Mol. Catal. A Chem.
https://pubs.acs.org/10.1021/acs.jpcc.3c04207 2012, 357, 81−86.
(19) Nagai, M. Transition-metal nitrides for hydrotreating catalyst�
Notes Synthesis, surface properties, and reactivities. Appl. Catal. A-general -
The authors declare no competing financial interest. APPL CATAL A-GEN. 2007, 322, 178−190.

■ ACKNOWLEDGMENTS
This research is supported by the Department of Science and
(20) Hargreaves, J. S. J. Nitrides as ammonia synthesis catalysts and as
potential nitrogen transfer reagents. Appl. Petrochemical Res. 2014, 4
(1), 3−10.
Technology, India, through a DST Inspire Faculty Award Grant (21) Oyama, S. T. Kinetics of ammonia decomposition on vanadium
nitride. J. Catal. 1992, 133 (2), 358−369.
(grant no. DST/INSPIRE/04/2019/000283), and all authors (22) Schlögl, R. Catalytic Synthesis of Ammonia�A “Never-Ending
are grateful to the Center for Computing and Information Story”? Angew. Chemie Int. Ed. 2003, 42 (18), 2004−2008.
Sciences RGIPT-Jais for high-performance computing. (23) Dongil, A. B. Recent progress on transition metal nitrides

■ REFERENCES
(1) Green, L., Jr. An ammonia energy vector for the hydrogen
nanoparticles as heterogeneous catalysts. Nanomaterials. 2019, 9 (8),
1111.
(24) Mukherjee, S.; Devaguptapu, S. V.; Sviripa, A.; Lund, C. R. F.;
economy. Int. J. Hydrogen Energy. 1982, 7 (4), 355−359. Wu, G. Low-temperature ammonia decomposition catalysts for
(2) Crabtree, G. W.; Dresselhaus, M. S.; Buchanan, M. V. Thoughts on hydrogen generation. Appl. Catal. B Environ. 2018, 226, 162−181.
Starting the Hydrogen Economy. Phys. Today. 2005, 58 (6), 15. (25) Srifa, A.; Okura, K.; Okanishi, T.; Muroyama, H.; Matsui, T.;
(3) Perman, E. P.; Atkinson, G. A. S.; Ramsay, W. The decomposition Eguchi, K. COx -free hydrogen production via ammonia decomposition
of ammonia by heat. Proc. R Soc. London. 1905, 74 (497−506), 110− over molybdenum nitride-based catalysts. Catal. Sci. Technol. 2016, 6
117. (20), 7495−7504.
(4) Bell, T. E.; Torrente-Murciano, L. H2 Production via Ammonia (26) Zhao, J.; Cui, C.; Wang, H.; Han, J.; Zhu, X.; Ge, Q. Insights into
Decomposition Using Non-Noble Metal Catalysts: A Review. Top the Mechanism of Ammonia Decomposition on Molybdenum Nitrides
Catal. 2016, 59 (15), 1438−1457. Based on DFT Studies. J. Phys. Chem. C 2019, 123 (1), 554−564.

16451 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(27) Huo, L.; Han, X.; Zhang, L.; Liu, B.; Gao, R.; Cao, B.; Wang, W.; (48) Zheng, W.; Zhang, J.; Zhu, B.; Blume, R.; Zhang, Y.; Schlögl, R.;
Jia, C.; Liu, K.; Liu, J.; Zhang, J. Spatial confinement and electron Schüth, F.; Su, D. Structure−Function Correlations for Ru/CNT in the
transfer moderating MoN bond strength for superior ammonia Catalytic Decomposition of Ammonia. ChemSusChem. 2010, 3 (2),
decomposition catalysis. Appl. Catal. B Environ. 2021, 294, No. 120254. 226−230.
(28) Arabczyk, W.; Zamlynny, J. Study of the ammonia decom- (49) Lide, D. R.; Frederikse, F. P. R. CRC Handbook of Chemistry and
position over iron catalysts. Catal. Lett. 1999, 60 (3), 167−171. Physics; CRC Press: New York, 1996; p 4.
(29) Johnson, D.; Hunter, B.; Christie, J.; King, C.; Kelley, E.; Djire, A. (50) Yeo, S. C.; Han, S. S.; Lee, H. M. Mechanistic investigation of the
Ti2N nitride MXene evokes the Mars-van Krevelen mechanism to catalytic decomposition of ammonia (NH3) on an Fe(100) surface: A
achieve high selectivity for nitrogen reduction reaction. Sci. Rep. 2022, DFT study. J. Phys. Chem. C 2014, 118 (10), 5309−5316.
12 (1), 657. (51) Duan, X.; Qian, G.; Fan, C.; Zhu, Y.; Zhou, X.; Chen, D.; Yuan,
(30) Qiao, Z.; Johnson, D.; Djire, A. Challenges and opportunities for W. First-principles calculations of ammonia decomposition on Ni(110)
nitrogen reduction to ammonia on transitional metal nitrides via Mars- surface. Surf. Sci. 2012, 606, 549−553.
van Krevelen mechanism. Cell Reports Phys. Sci. 2021, 2 (5), (52) Chen, W.; Ermanoski, I.; Madey, T. E. Decomposition of
No. 100438. Ammonia and Hydrogen on Ir Surfaces: Structure Sensitivity and
(31) Laassiri, S.; Zeinalipour-Yazdi, C. D.; Catlow, C. R. A.; Nanometer-Scale Size Effects. J. Am. Chem. Soc. 2005, 127 (14), 5014−
Hargreaves, J. S. J. Nitrogen transfer properties in tantalum nitride 5015.
based materials. Catal. Today. 2017, 286, 147−154. (53) Xiao, X.-Z.; Cao, Y.-L.; Cai, Y.-Y. Decomposition of NH3 on
(32) Zeinalipour-Yazdi, C. D.; Hargreaves, J. S. J.; Catlow, C. R. A. Ir(110): A first-principle study. Surf. Sci. 2011, 605 (7), 802−807.
Nitrogen Activation in a Mars−van Krevelen Mechanism for Ammonia (54) Huang, W.; Lai, W.; Xie, D. First-principles study of
decomposition of NH3 on Ir(100). Surf. Sci. 2008, 602, 1288−1294.
Synthesis on Co3Mo3N. J. Phys. Chem. C 2015, 119 (51), 28368−
(55) Hammer, B.; Norskov, J. K. Why gold is the noblest of all the
28376.
metals. Nature. 1995, 376 (6537), 238−240.
(33) Schnepp, Z.; Thomas, M.; Glatzel, S.; Schlichte, K.; Palkovits, R.;
(56) Jung, J.-E.; Geatches, D.; Lee, K.; Aboud, S.; Brown, G. E.;
Giordano, C. One pot route to sponge-like Fe3N nanostructures. J.
Wilcox, J. First-Principles Investigation of Mercury Adsorption on the
Mater. Chem. 2011, 21 (44), 17760−17764. α-Fe2O3(11̅02) Surface. J. Phys. Chem. C 2015, 119 (47), 26512−
(34) Tagliazucca, V.; Schlichte, K.; Schüth, F.; Weidenthaler, C. 26518.
Molybdenum-based catalysts for the decomposition of ammonia: In (57) Wang, Y.; Yannello, V.; Graterol, J.; Zhang, H.; Long, Y.; Shatruk,
situ X-ray diffraction studies, microstructure, and catalytic properties. J. M. Theoretical and Experimental Insights into the Effects of Zn Doping
Catal. 2013, 305, 277−289. on the Magnetic and Magnetocaloric Properties of MnCoGe. Chem.
(35) Dahl, S.; Logadottir, A.; Jacobsen, C. J. H.; Nørskov, J. K. Mater. 2020, 32 (15), 6721−6729.
Electronic factors in catalysis: the volcano curve and the effect of
promotion in catalytic ammonia synthesis. Appl. Catal. A Gen. 2001,
222 (1), 19−29.
(36) Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid
metals. Phys. Rev. B 1993, 47 (1), 558.
(37) Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy
calculations for metals and semiconductors using a plane-wave basis set.
Comput. Mater. Sci. 1996, 6 (1), 15−50.
(38) Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio
total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996,
54 (16), 11169−11186.
(39) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient
approximation made simple. Phys. Rev. Lett. 1996, 77 (18), 3865.
(40) Wang, V.; Xu, N.; Liu, J. C.; Tang, G.; Geng, W. VASPKIT: A
user-friendly interface facilitating high-throughput computing and
analysis using VASP code. Comput. Phys. Commun. 2021, 267,
No. 108033.
(41) Momma, K.; Izumi, F. VESTA 3 for three-dimensional
visualization of crystal, volumetric and morphology data. J. Appl.
Crystallogr. 2011, 44 (6), 1272−1276.
(42) Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone
integrations. Phys. Rev. B 1976, 13 (12), 5188−5192.
(43) Henkelman, G.; Uberuaga, B. P.; Jónsson, H. A climbing image
nudged elastic band method for finding saddle points and minimum
energy paths. J. Chem. Phys. 2000, 113 (22), 9901−9904.
(44) Deringer, V. L.; Tchougréeff, A. L.; Dronskowski, R. Crystal
Orbital Hamilton Population (COHP) Analysis As Projected from
Plane-Wave Basis Sets. J. Phys. Chem. A 2011, 115 (21), 5461−5466.
(45) Maintz, S.; Deringer, V. L.; Tchougréeff, A. L.; Dronskowski, R.
LOBSTER: A tool to extract chemical bonding from plane-wave based
DFT. J. Comput. Chem. 2016, 37 (11), 1030−1035.
(46) Zhou, S.; Lin, S.; Guo, H. First-Principles Insights into Ammonia
Decomposition Catalyzed by Ru Clusters Anchored on Carbon
Nanotubes: Size Dependence and Interfacial Effects. J. Phys. Chem. C
2018, 122 (16), 9091−9100.
(47) Li, L.; Zhu, Z. H.; Yan, Z. F.; Lu, G. Q.; Rintoul, L. Catalytic
ammonia decomposition over Ru/carbon catalysts: The importance of
the structure of carbon support. Appl. Catal. A Gen. 2007, 320, 166−
172.

16452 https://doi.org/10.1021/acs.jpcc.3c04207
J. Phys. Chem. C 2023, 127, 16442−16452

You might also like