You are on page 1of 9

Journal of Catalysis 383 (2020) 322–330

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

DFT and microkinetic comparison of Pt, Pd and Rh-catalyzed ammonia


oxidation
Hanyu Ma a, William F. Schneider a,b,⇑
a
Department of Chemical and Biomolecular Engineering, University of Notre Dame, Notre Dame, IN 46556, United States
b
Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame, IN 46556, United States

a r t i c l e i n f o a b s t r a c t

Article history: Ammonia oxidation is the heart of the Ostwald process and is important in emissions control. Catalytic
Received 26 November 2019 behaviors are a function of conditions and are observed to vary across the platinum group metals (PGMs)
Revised 7 January 2020 Pt, Pd, Rh. Here, we combine density functional theory computations and microkinetic modeling to
Accepted 29 January 2020
rationalize these dependencies. We compute reactions over model (2 1 1) and (1 1 1) surfaces of
Available online 20 February 2020
PGMs. Binding energies are similar on Pd and Pt and generally greater on Rh, while activation energies
vary across all metals. Rates on (2 1 1) surfaces are greater than (1 1 1) surfaces. The stepped Pt is most
Keywords:
active and stepped Rh most selective to N2 at ammonia slip conditions, while at Ostwald process condi-
Ammonia oxidation
Platinum group metals
tions, stepped Pd is most active and stepped Pt most selective to NO. Degree of rate and selectivity control
Density functional theory analysis provides insights into the reactions limiting performance of PGMs. Both activation barriers and
Microkinetic modeling surface coverages influence rates and selectivities.
Structure sensitivity Ó 2020 Elsevier Inc. All rights reserved.

1. Introduction Pt, Pd and Rh are the most studied platinum group metals
(PGMs) for ammonia oxidation [7,2,8–10]. In the low-
Ammonia oxidation is catalyzed by platinum-based materials to temperature regime, reaction rates are greatest for Pt while N2
effect both ammonia slip control and the Ostwald process [1]: selectivity is greatest over Rh [11,12]. In the high-temperature
regime, rates are greatest over Pd, while selectivity to NO is great-
NH3 ðgÞ þ 0:75O2 ðgÞ ! 0:5N2 ðgÞ þ 1:5H2 OðgÞ ð1aÞ est for Pt [13,14]. The microscopic mechanisms that control the
rate and selectivity trends remain unclear.
NH3 ðgÞ þ 1:25O2 ðgÞ ! NOðgÞ þ 1:5H2 OðgÞ ð1bÞ
Chmielarz et al. measured ammonia conversion and product
NH3 ðgÞ þ O2 ðgÞ ! 0:5N2 OðgÞ þ 1:5H2 OðgÞ ð1cÞ selectivity over oxide-supported Pt, Pd and Rh particles in a plug
flow reactor (PFR) at temperatures relevant to ammonia slip con-
Selectivity to the thermodynamic N2 product over N2O and NO is a trol [11]. Independent of oxide support, NH3 conversion reached
key performance metric in ammonia slip applications for environ- 100% at the lowest temperature over Pt, followed by Pd and Rh.
mental protection, in which the goal is to oxidize NH3 with O2 to The N2 selectivity, however, followed the reverse order of
limit NH3 release from diesel vehicles or industrial point sources Rh > Pd > Pt, across temperatures from 373 to 773 K. Selectivities
to the environment. Pt nanoparticles are the most common cata- over Pt and Pd were similar. Li and Armor reported similar exper-
lysts [2], and selectivities of current materials need to be improved iments over Al2O3-supported Pt, Pd and Rh catalysts [12]. Activities
upon to meet increasingly stringent regulations to protect the envi- were greatest and N2 selectivity least over the Pt catalyst. Activities
ronment [3,4]. NO is the desired product of NH3 oxidation in the were lower and selectivities to N2 were greater over both Pd and
Ostwald process, and a Pt-based gauze is the most common catalyst Rh. Pignet and Schmidt observed NH3 oxidation over Pt, Pd, and
[5]. While selectivity to NO is generally high, N2O, a greenhouse gas Rh wires in a continuous stirred-tank reactor (CSTR) operated from
and stratospheric ozone depleter, is a common secondary product. 500 to 1673 K and total pressures of 0.1–1 Torr [14]. At tempera-
Nitric acid plants are the largest source of N2O emissions in the tures relevant to ammonia slip control, reaction rates followed
chemical industry [6]. Pt > Pd > Rh and N2 selectivities tracked in the opposite order. At
higher temperatures relevant to the Ostwald process, rates fol-
⇑ Corresponding author at: Department of Chemical and Biomolecular Engineer- lowed Pd > Pt > Rh and NO selectivities followed Pt > Pd > Rh.
ing, University of Notre Dame, Notre Dame, IN 46556, United State. Pérez-Ramírez et al. used Temporal Analysis of Products (TAP) to
E-mail address: wschneider@nd.edu (W.F. Schneider).

https://doi.org/10.1016/j.jcat.2020.01.029
0021-9517/Ó 2020 Elsevier Inc. All rights reserved.
H. Ma, W.F. Schneider / Journal of Catalysis 383 (2020) 322–330 323

measure ammonia oxidation rates over Pt, Pd and Rh wires at reactions, where a 3  3 supercell was used. The bottom two
1073 K [13]. The ammonia conversion rates follow Pd > Pt  Rh (1 1 1) layers were fixed to bulk metal locations and the top three
and the NO selectivity followed Pt > Pd > Rh. Thus, relative rates (1 1 1) layers were allowed to relax in all optimizations. We sam-
and selectivities differ across metals but are the same whether pled the first Brillouin zone with a C-centered 5  5  1
the metal catalysts are supported or unsupported. Monkhorst-Pack k-point mesh [31,32]. We modeled the (1 1 1)
NH3 oxidation reaction pathways have been evaluated using surface with five-layer-thick slab in a 2  2 supercell. The bottom
supercell density functional theory (DFT) models of Pt(2 1 1) two layers were fixed and the top three metal layers were allowed
[15,16], Pt(1 1 1) [15,17], Pt(100) [18] and Rh(1 1 1) [19] surface to relax. The first Brillouin zone was sampled with a C-centered
facets, and in the case of Pt, the results used to parametrize 7  7  1 Monkhorst-Pack k-point mesh. Both the (2 1 1) and
microkinetic models [15,18,20]. A consistent set of DFT and (1 1 1) slabs were separated by a 14Å vacuum gap.
microkinetic results across metal and surface structure has not We calculated the rate constants of non-activated adsorption
been reported. DFT calculations report that barriers to oxidatively steps using the Hertz-Knudsen equation. We then calculated the
activate ammonia are similar and barriers to form products are adsorption equilibrium constants and estimated the desorption
lower at Pt(2 1 1) step models than on Pt(1 1 1) terraces [15]. rate constants [33]. In the Hertz-Knudsen equation (Eq. (2)), a is
NH3 activation and product desorption barriers are reported to a sticking coefficient, A is the area of one free site, m is the molec-
be lower still on Pt(100) than either Pt(1 1 1) or (2 1 1) [18]. These ular weight of a gas molecule, kB is the Boltzmann constant and T is
results all suggest a strong structure sensitivity to catalytic activity. the absolute temperature [34]. Previous comparisons show that
However, microkinetic models parametrized from the Pt(1 1 1) and that computed rate and selectivity trends are insensitive to details
Pt(2 1 1) DFT results report that selectivity shifts from N2 to NO of these sticking coefficients [15] for ammonia oxidation on Pt. We
with increasing temperature, independent of facet, that these take sticking coefficient here to be unity, as an upper limit on
selectivities are controlled by the relative coverages of adsorbed adsorption rates [33].
N and O, and that while selectivities are structure-insensitive,
aA
absolute rates on step sites exceed terraces at low temperatures, kads ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð2Þ
where product desorption is rate-controlling, and converge at high 2pmkB T
temperature, where ammonia activation becomes rate-controlling. We calculated the adsorption equilibrium constants with Eq. (3).
A similar DFT-computed reaction network has been reported over The enthalpies of adsorption are taken as the zero-point-energy-
Rh(1 1 1) [19]. In general, ammonia activation and product forma- corrected DFT energies. We used the entropies of gas molecules
tion barriers are greater over Rh(1 1 1) than Pt(1 1 1). Correspond- from NIST-JANAF thermochemical tables [35] and estimated the
ing microkinetic results are not available. entropies of adsorbates using harmonic oscillator (HO) model. For
Here, we report intrinsic reaction and activation energies of NH3 and H2O, we exclude the molecular rotational modes parallel
reactions relevant to ammonia oxidation using supercell DFT and to surfaces from the HO model. These modes have frequencies near
(2 1 1) and (1 1 1) model surfaces of Pd and Rh and contrast against 50 cm1 and their contributions to the overall entropies are esti-
Pt results. While Pt and Pd are similar, adsorption and activation mated to be small with the hindered rotor model. However, these
energies are systematically greater over Rh. We report microki- rotational entropies are severely overestimated by the HO model
netic models parametrized against these DFT results and compare [36].
degrees of rate and selectivity control across all metals and two  
surface structures. In general, predicted relative rates and selectiv- DEDFT þ DEZPE  T DS ðTÞ
KðTÞ ¼ exp  ð3Þ
ities are consistent with experimental observation: Pt is both most kB T
active and least selective to N2 at NH3 slip conditions, and Rh is We calculated the rate constants of desorption processes with Eq.
selective to N2 at all conditions. While relative surface coverages (4) to enforce thermodynamic consistency, where P is the standard
appear to control selectivity on Pt and Pd, selectivity on Rh arises pressure of 1 bar:
from the significantly lower barrier for N2 formation than NO.
These results highlight the interplay between NH3 activation and kads P
kdes ¼ ð4Þ
product release kinetics that influence both rates and selectivities. K
Reaction rate constants (k) involving activated chemisorption
2. Computational details and surface reactions were calculated using harmonic transition
state theory (hTST):
 
Plane-wave, periodic supercell DFT calculations were per- kB T qTS Ea
formed with the Vienna ab initio Simulation Package (VASP) k¼ exp  ð5Þ
h qIS kB T
[21,22]. Ion core states were treated with the projector augmented
wave (PAW) method [23,24] and electron exchange and correlation Here h is the Planck constant and Ea is the zero-point-corrected
were treated within the generalized gradient approximation (PBE- energy difference between the transition state and the adsorbed
GGA) [25]. We included plane-waves to a cutoff energy of 400 eV reactants at infinite separation. We assume qqTSIS ¼ 1 for surface reac-
and converged the forces to 0.05 eV/Å. Transition states tions; we show that this simplification has negligible influence on
were located by the Climbing-Image Nudged Elastic Band method computed microkinetic rates and selectivities when compare
(CI-NEB) [26] and then refined using the dimer method [27]. We against models parametrized within a harmonic oscillator model
further confirmed the transition states by performing vibrational [15]. The steady-state surface coverages, rates and selectivities were
analysis, and identifying one imaginary mode along the reaction solved assuming fixed pressures of gaseous species with a mean-
coordinate. field microkinetic model [37]. The results are equivalent to kinetics
The lattice constants of Pd and Rh are 3.941 and 3.825 Å, respec- of catalysts in a continuous stirred-tank reactor with infinite small
tively, within the model used here, as determined by fitting to the residence time. The differential equations for the coverage of sur-
Birch-Murnaghan equation of state [28,29]. These values agree face species i are written as:
well with the experimental values of 3.950 and 3.820 Å [30]. We @hi X
modeled the (2 1 1) surfaces using five layers of close packed ¼ cij rj ð6Þ
@t
(1 1 1) atoms in a 3  2 supercell, except for O2 dissociation j
324 H. Ma, W.F. Schneider / Journal of Catalysis 383 (2020) 322–330

and 3.1. Adsorption


Y c
Y c
Y c Y c
r j ¼ kþj hþj ij Pþj ij  kj hjij Pjij ð7Þ
The most favorable binding sites of surface intermediates and
þj þj j j
their corresponding geometries are comparable on the three met-
where cij is the coefficient for the stoichiometry of species i in reac- als (Table S1, Fig. S1 and Fig. S2). In general, NH3 and H2O prefer
tion j and r j is the net rate of surface reaction j. The forward and atop sites, NH2 and OH prefer bridge sites, and NH, N and O prefer
backward reactions are represented as þj and j, respectively. kj hollow sites for both (2 1 1) and (1 1 1) surfaces. The site prefer-
is the rate constant and hj and P j are the coverages and pressures ences of NO and N2O are different on (2 1 1) and (1 1 1) surfaces,
of different species involved in reaction j, respectively. The coverage but the preference across PGMs are unvaried. For intermediates
is conserved for a surface: that bind on hollow sites, there is no obvious trend on the prefer-
X ence of fcc or hcp sites across different metals and surfaces. Bind-
h þ hi ¼ 1 ð8Þ ing energies are summarized in Table S1 and results on all surfaces
i
are plotted against results on Pt [15] in Fig. 1. Binding energies are
where h is the coverage of free sites. We solved the ordinary differ- greatest in magnitude on Rh and similar on Pd and Pt, independent
ential equations (ODE) at steady state, where @h@t
i
¼ 0, using a method of facet. The exceptions are NH3 and H2O, which bind similarly on
that automatically switches between nonstiff (Adams) and stiff all surfaces, and NO, which binds more strongly on Pd than Pt.
(BDF) solvers, as implemented in scipy.integrate.odeint in Python. Binding site preferences and binding energies are consistent with
The steady-state coverages can also be solved from a system of alge- previous reports [43–45].
P
braic equations ( j cij rj ¼ 0). We further solved the system of alge-
braic equations using the coverages solved from ODE as initial 3.2. Ammonia activation and product formation
guesses to confirm the steady state coverages. We used Newton’s
method, as implemented in mpmath.findroot in Python. The similar adsorption site preferences suggest similar reaction
pathways on the three metals, and as shown in Fig. S3-S14, com-
3. Results and discussion puted transition state structures are similar on all metals. In sum-
mary, the minimum energy path (MEP) is a convolution of H2N-H
We included ammonia oxidation pathways suggested by both bond scission followed by NH2 migration from an atop to bridge
prior experimental and theoretical work, including NHx site, and the highest point along this MEP corresponds with the lat-
(x ¼ 0; 1; 2; 3) activation over metal, O*, and OH* sites, where * ter migration process. H-transfer corresponds to the highest barrier
indicates a surface-bound species. N2, NO, and N2O are assumed for NH2 and NH activation, similar to observations on Pt(2 1 1) [16],
to be formed from N* combination with N*, O* and NO*. Possible Pt(1 1 1) [17], Rh(1 1 1) [19], Au(1 1 1) [41] and Cu(1 1 1) [42]. On
routes from NH3 adsorption to product generation are shown in the (2 1 1) surfaces of the three metals, oxidative NH3 and NH acti-
Scheme 8. Experimentally, surface-bound NHx (x ¼ 0; 1; 2; 3) and vation is preferred at the (2 1 1) step while NH2 activation occurs
OHy (y ¼ 0; 1) have been observed on Pt(1 1 1) under ultrahigh vac- on the terrace. N2 and NO formation occur on the step of (2 1 1) and
uum conditions after dosing O and NH3 [38]. Surface-bound NHx N2O formation occurs on the terrace, consistent with a previous
(x ¼ 0; 1; 2; 3) and O have been identified on Ir(1 1 1) with X-ray report [46]. Fig. 2 compares reaction profiles on the (2 1 1) facets
photoelectron spectroscopy (XPS) following dosing of O and NH3 of the three metals. NHx activations on O* sites have smaller reac-
[39]. Similarly, NHx and O were observed on stepped Pt(533) with tion and activation energies than on OH* sites, except for the NH2
in situ XPS at low pressures [40]. Recently, Dann et al. observed N activation barriers. Activation barriers for the reactions occurring
and O insertion into supported metallic Pd at low and high T on the terrace—NH2 activation and N2O formation—are greater
respectively with operando X-ray absorption fine structure spec- than those occurring at the (2 1 1) step.
troscopy, suggesting high surface coverage of N at low T and O at Fig. 3 reports corresponding results on the (1 1 1) facet. Unlike
high T [10]. Theoretically, the reaction pathways were evaluated the (2 1 1) surfaces, ammonia activations on OH* site of (1 1 1) sur-
with DFT computation on Pt(2 1 1) [15,16], Pt(1 1 1) [15,17], Pt faces have smaller reaction and activation energies than on O sites.
(1 0 0) [18], Rh(1 1 1) [19], Au(1 1 1) [41] and Cu(1 1 1) [42]. These reaction energies are related to reaction (9) in Table S2,
Microkinetic models on Pt(2 1 1), Pt(1 1 1), and Pt(1 0 0) including OH* + OH* M O* + H2O*. As a result, the energy difference between
the aforementioned pathways capture the essential rate and selec- NH3* + OH* M NH2* + H2O* and NH3* + O* M NH2* + OH* is positive
tivity trends of experimental observations [15,18,20]. We therefore on (2 1 1) surfaces and negative on (1 1 1) surfaces, for instance.
performed plane-wave, supercell DFT computation within the gen- N2* and NO* formation barriers on (1 1 1) surfaces are much
eralized gradient approximation on the reactions shown in greater than those on (2 1 1) surfaces, suggesting lower N2 and
Scheme 8 for Pd(2 1 1), Pd(1 1 1), Rh(2 1 1) and Rh(1 1 1). We NO production rates on the former than the latter.
include previously reported results using the same methods for Activation barriers are plotted against Pt results in Fig. 4 to
Pt(2 1 1) and Pt(1 1 1) for comparison [15]. illustrate variations with metal. Ammonia activation barriers are

Scheme 8. Schematic representation of possible reaction pathways from NH3 adsorption to oxidation product generation. Each segment represents an elementary step, and
widths are arbitrary. Light pink, light blue, and light gray segments denote successive N–H cleavage steps across metal, O*, and OH* sites, respectively. Orange, pink, and blue
segments represent pathways leading to NO, N2O, and N2, respectively.
H. Ma, W.F. Schneider / Journal of Catalysis 383 (2020) 322–330 325

Fig. 1. Binding energies of (a) NHx and (b) oxygen-containing species on Pt(2 1 1), Pd(2 1 1) and Rh(2 1 1) plotted against those on Pt(2 1 1). Binding energies of (c) NHx and
(d) oxygen-containing species on Pt(1 1 1), Pd(1 1 1) and Rh(1 1 1) plotted against those on Pt(1 1 1).

Fig. 2. Energy diagrams for O*- and OH*-assisted ammonia activation on (a) Pt(2 1 1), (b) Pd(2 1 1) and (c) Rh(2 1 1). Energy diagrams for N2, NO and N2O formation on (d) Pt
(2 1 1), (e) Pd(2 1 1) and (f) Rh(2 1 1).

generally greater on Rh than Pd and Pt. Activation barriers are sim- DFT reaction and activation energies to predict the steady-state
ilar on (1 1 1) facets of Pt and Pd and lower on Pd(2 1 1) than Pt reaction rates and selectivities at temperatures relevant to both
(2 1 1). Barriers to form products are greater on Rh than on Pt ammonia slip control and the Ostwald process [37]. We assume
and Pd. These differences suggest differences in performance in adsorbate diffusion is much faster than surface reactions, consis-
ammonia oxidation that we explore next. tent with convention [47]. We first computed rates at zero ammo-
nia conversion to capture the intrinsic activities of the surfaces. We
3.3. Rate and selectivity then parametrized a continuous stirred-tank reactor (CSTR) to
determine reaction rates and selectivities as a function of conver-
To explore the kinetic consequences of these differences, we sion. Rate parameters are in Table S2-S5. The evolution of most
parametrized a mean-field microkinetic model with the computed abundant surface intermediate (MASI) with T at the same reaction
326 H. Ma, W.F. Schneider / Journal of Catalysis 383 (2020) 322–330

Fig. 3. Energy diagrams for O*- and OH*-assisted ammonia activation on (a) Pt(1 1 1), (b) Pd(1 1 1) and (c) Rh(1 1 1). Energy diagrams for N2, NO and N2O formation on (d) Pt
(1 1 1), (e) Pd(1 1 1) and (f) Rh(1 1 1).

Fig. 4. Activation barriers of (a) ammonia oxidation and (b) product formation on Pt(2 1 1), Pd(2 1 1) and Rh(2 1 1) plotted against those on Pt(2 1 1). Activation barriers of (c)
ammonia oxidation and (d) product formation on Pt(1 1 1), Pd(1 1 1) and Rh(1 1 1) plotted against those on Pt(1 1 1). The activation barriers for product formation are
calculated as A* + B* ? AB(g) + 2*.

conditions of Fig. 5 is shown in Fig. S15. The MASIs on Pt(2 1 1) and on terrace surfaces are similar across the three metals: N* domi-
Pd(2 1 1) are comparable, evolving from N* to NO* to O* with nates at low T, transitioning to O* at high T.
increasing T. Unlike Pt(2 1 1) and Pd(2 1 1), NO* is not observed Fig. 5(a) and (b) report computed rates as Arrhenius plots at
on Rh(2 1 1) and the MASI changes from N* to O* with increasing zero conversion. TOFs are greater on stepped than terraced facets
T. The absence of NO* is consistent with its low decomposition at low and intermediate temperatures across all on the metals. Ter-
energy and barrier to decompose to N* and O*. The MASI profile raced surfaces have a more pronounced temperature dependence,
H. Ma, W.F. Schneider / Journal of Catalysis 383 (2020) 322–330 327

Fig. 5. (a) Reaction rate against reciprocal temperature on Pt(2 1 1), Pd(2 1 1) and Rh(2 1 1). (b) Reaction rate against reciprocal temperature on Pt(1 1 1), Pd(1 1 1) and Rh
(1 1 1). (c) Selectivities to reaction products on Pt(2 1 1), Pd(2 1 1) and Rh(2 1 1). (d) Selectivities to reaction products on Pt(1 1 1), Pd(1 1 1) and Rh(1 1 1). Pressures of NH3, O2
and H2O are 0.001, 0.02 and 0.05 bar, respectively.

and thus rates begin to converge at the highest temperature limits. selectivities than Pd at high conversions of NH3. Significant N2O
Relative activity trends across metal are both facet and selectivity is observed at high conversion on Pt. Rh shows high
temperature-dependent. Stepped Pt rates are greatest at the lower N2 selectivities at different conversions. Selectivity trends are also
temperatures characteristic of ammonia slip catalysis, and stepped consistent with the predicted orders at zero conversion.
Pd rates exceed Pt at Ostwald conditions. Fig. 5(c) and (d) compare The predicted rate and selectivity trends compare well with a
corresponding product selectivities, defined as the relative product number of experimental observations. In plug-flow reactor exper-
formation rates, si : iments at conditions representative of NH3 slip control, NH3
conversions are greatest and N2 selectivities least on oxide-
ri
si ¼ P ð9Þ supported Pt nanoparticle catalysts [11]. Conversions are dimin-
j rj ished and N2 selectivities greater on Pd and Rh. These results are
consistent with TOF and selectivity in Fig. 5(a) and (c). The rate
where i represents a product and j indexes over the three products. and selectivity orders are in general the same on different oxide
The N2 selectivity decreases and NO selectivity increases with supports. In the high-temperature regime of ammonia oxidation,
increasing T across all models, consistent with experimental obser- bulk PGM wires were tested in both continuous stirred-tank reac-
vations. The temperatures at which products change from N2 to NO, tors (CSTR) and temporal analysis of products (TAP) reactors to
and the propensity to generate N2O in the transition region, vary understand NO production for the Ostwald process [13,14]. Pd
with both metal and facet. The step surfaces evolve from N2 to shows the highest rate and Pt shows the highest NO selectivity.
NO selectivity in the order T Pt < T Pd  T Rh , with only Pt predicted Rh was the least active and selective to NO. Moreover, TAP exper-
to produce N2O. While Pt results are relatively insensitive to facet, iments showed Pd and Rh have lower selectivity to N2O than Pt
transition temperatures are substantially decreased on Rh(1 1 1) [13]. These observations are consistent with the prediction from
and Pd(1 1 1), and the latter in particular is predicted to produce the simulated kinetics at zero conversions and in CSTR. The relative
a substantial fraction of N2O at intermediate temperatures. activities predicted from the terraced surface computations do not
To understand the sensitivity of these predictions to NH3 con- correspond with observations over supported nanoparticles or
version, we applied the (2 1 1) parameters to a continuous bulk wires discussed above. Further, predicted TOFs at steps
stirred-tank reactor (SI, Page 19) with the inlet conditions the same greatly exceed those at terraces. Thus, observed activities more
as those in the zero conversion model (Fig. 5), varying the reactor likely represent contributions from step-like than terrace-like sites,
residence time at fixed number of active sites per volume. Results especially at low and intermediate T.
are reported in Fig. S16 and Fig. S17. The conversion of NH3
increases from 0 to 100% by increasing the residence time. The
light-off residence time, s, follows sPt < sPd < sRh at 500 K and 3.4. Degree of rate and selectivity control and reaction flow pathways
sPd < sPt < sRh at 1000 K. Both trends are consistent with the pre-
dicted rate trends at zero conversion. At 500 K, all three surfaces To identify the elementary steps most responsible for the over-
are selective to N2 across different conversions, except that Pt all reaction rates, we performed degree of rate control (DRC) across
(2 1 1) shows decrease of N2 and increase of N2O selectivities when PGMs [48]. We define rate sensitivity to specific reactions in Eq.
conversion approaches unity. At 1000 K, NO selectivities on Pt (10), where X RC;i is calculated by changing the forward (kþi ) and
and Pd drop with NH3 conversion. Pt shows much higher NO backward (ki ) rate constants of step i simultaneously by 1% while
328 H. Ma, W.F. Schneider / Journal of Catalysis 383 (2020) 322–330

holding constant the rate constants of other steps, kj , and the equi-
librium constant, K i , and monitoring the change of the overall rate
(r) with respect to the reference rate, r0 .
 
ki @r ki;0 r  r 0 r  r0
X RC;i ¼   ¼ ð10Þ
r @ki kj–i ;K i r0 ki  ki;0 0:01r 0

Because the predicted TOFs on stepped surfaces are higher than


those on terraced surfaces, and the rate and selectivity trends of
the stepped surfaces match those in experiments, we compare
degree of rate control as a function of T for Pt, Pd and Rh stepped
surfaces in Fig. 6. The DRC values sum to 1, as expected [48]. We
compare the DRC values for different reactions vs T to understand
the dependence of rate on T and metals. Different reactions control
the overall rate on the three metals. At low T, multiple reactions
control the reaction rate on stepped Pt and Pd surfaces. On stepped
Rh, N2 formation is the single rate-determining step. The higher
barrier of N2 formation on stepped Rh surface results in its low
activity at low T. At high T, the reaction rate is controlled by NH3*
activation by O* on all three metals. The reaction rate order, i.e.
Pd > Pt > Rh, is therefore determined by the barrier order of
NH3* + O*.
Similar to the activity, ammonia oxidation selectivity has a
metal and temperature dependence. The temperature dependence
is similar on different PGMs, but selectivity transitions from N2 to
NO at different temperatures on the three metals. To identify the
elementary steps responsible for this selectivity behavior, we com-
puted degree of N2 selectivity control (DSC) (Eq. (11)) by replacing
rate with N2 selectivity in Eq. (10) [48].
  Fig. 7. Steady-state net rates of surface reactions on Pt(2 1 1), Pd(2 1 1) and Rh
ki @sN2 ki;0 sN2  sN2 ;0 sN2  sN2 ;0 (2 1 1) at 500 and 1000 K, presented as a Sankey diagram, in which each segment
X SC;i ¼   ¼ ð11Þ
sN 2 @ki kj–i ;K i sN2 ;0 ki  ki;0 0:01sN2 ;0 indicates an elementary step and the corresponding segment width indicates net
rate normalized to the overall ammonia oxidation rate. The partial pressures of NH3,
O2 and H2O are 0.001, 0.02 and 0.05 bar, respectively.
As expected, the DSC values sum to 0 [48]. DSC as a function of T are
again similar on stepped Pt and Pd surfaces: NH3* activation by O*
and N2 formation contributes positively to N2 selectivity while NO surface N and O positively and therefore improves the N2 formation
formation and O2 adsorption contributes negatively to N2 selectiv- and depresses NO formation. O2 adsorption, however, contributes to
ity. NH3* activation by O* contributes to the relative abundance of the relative abundance of surface N and O negatively and decreases

Fig. 6. Degree of rate control on (a) Pt(2 1 1), (b) Pd(2 1 1) and (c) Rh(2 1 1). Degree of N2 selectivity control on (d) Pt(2 1 1), (e) Pd(2 1 1) and (f) Rh(2 1 1). The partial
pressures of NH3, O2 and H2O are 0.001, 0.02 and 0.05 bar, respectively.
H. Ma, W.F. Schneider / Journal of Catalysis 383 (2020) 322–330 329

the N2 selectivity. On stepped Rh, reactions start to show observable Data statement
DSC values only at high T, indicating its high N2 selectivity at low
and intermediate T is not controlled by any single elementary reac- Raw data and Python codes are available at https://doi.org/10.
tion. Rh shows much higher N2 selectivity over a wide range of T 5281/zenodo.3620943.
than both Pt and Pd. The high formation barrier of NO from N*
and O* on stepped Rh prevents the formation of NO and increases Declaration of Competing Interest
the N2 selectivity.
To compare the relative contribution of each reaction step to The authors declare that they have no known competing finan-
the overall rate, we report the net rate of each reaction step nor- cial interests or personal relationships that could have appeared
malized to the overall rate in the form of Sankey diagrams [49]. to influence the work reported in this paper.
Unlike Scheme 8, where the segment widths are arbitrary, segment
widths in Fig. 7 indicate the contribution of each elementary reac-
Acknowledgment
tion to the overall ammonia oxidation rate. NH3 decomposition
shows negligible contribution to the overall rate compared to
This research was supported by the U.S. Department of Energy,
NH3 activation by O* and OH*. The net rate of the first step of
Office of Science, Office of Basic Energy Sciences, under Award DE-
ammonia activation is always dominated by the O* pathway. At
FG02-06ER15830. The computing resources and technical support
low T, OH* pathway contributes more than O* pathway for NH2*
for this work were provided by the Notre Dame Center for Research
and NH activation. At high T, the contribution of OH* pathway
Computing and National Energy Research Scientific Computing
decreases.
Center, a DOE Office of Science User Facility supported by the Office
The reaction fluxes are generally comparable but there are still
of Science of the U.S. Department of Energy under Contract No. DE-
differences among PGMs. At 500 K, the contribution of OH* path-
AC02-05CH11231. We thank Craig Waitt for comments and
way to the oxidation of NH* follows Rh > Pd > Pt. At 1000 K, OH*
discussions.
pathway shows observable contribution to the oxidation of NH2
on Pd while the contribution is negligible on the other two metals.
The Sankey diagrams show the unity N2 selectivity at 500 K on all Appendix A. Supplementary material
three metals. At 1000 K, high NO selectivity is observed on Pt and
Pd while Rh shows high N2 selectivity. Supplementary data associated with this article can be found, in
the online version, at https://doi.org/10.1016/j.jcat.2020.01.029.

4. Conclusions References

Both rates and selectivities are predicted to be functions of plat- [1] J.G. Chen, R.M. Crooks, L.C. Seefeldt, K.L. Bren, R.M. Bullock, M.Y. Darensbourg,
P.L. Holland, B. Hoffman, M.J. Janik, A.K. Jones, M.G. Kanatzidis, P. King, K.M.
inum group metals. DFT and microkinetic models show that the Lancaster, S.V. Lymar, P. Pfromm, W.F. Schneider, R.R. Schrock, Beyond fossil
NH3 oxidation rate is lower and selectivity to N2 is greater over fuel-driven nitrogen transformations, Science 360 (6391) (2018) eaar6611,
Rh and Pd than over Pt. The choice of PGMs metals for ammonia https://doi.org/10.1126/science.aar6611.
[2] L. Chmielarz, M. Jabłońska, Advances in selective catalytic oxidation of
slip catalysis is therefore a trade-off between rate and selectivity. ammonia to dinitrogen: a review, RSC Adv. 5 (54) (2015) 43408–43431,
Currently, a layer of selective catalytic reduction (SCR) catalyst is https://doi.org/10.1039/C5RA03218K.
often placed on top of supported Pt in ammonia slip catalysis [3] M. Jabłońska, R. Palkovits, Copper based catalysts for the selective ammonia
oxidation into nitrogen and water vapour-recent trends and open challenges,
[50]. The zeolite-based SCR catalysts reduce NO with NH3 to N2 Appl. Catal., B 181 (2016) 332–351, https://doi.org/10.1016/j.
to compensate the low N2 selectivity of Pt [51]. The dual layer cat- apcatb.2015.07.017.
alyst increases the N2 selectivity but reduces the overall rate [52]. [4] P.S. Dhillon, M.P. Harold, D. Wang, A. Kumar, S. Joshi, Hydrothermal aging of Pt/
Al2O3 monolith: washcoat morphology degradation effects studied using
Combining Pt with Rh or Pd may achieve similar performance of ammonia and propylene oxidation, Catal. Today 320 (2019) 20–29, https://doi.
the dual layer catalyst for ammonia slip control. For the Ostwald org/10.1016/j.cattod.2017.12.023.
process, NO selectivity is the highest on Pt. However, a small per- [5] C.H. Bartholomew, R.J. Farrauto, Fundamentals of Industrial Catalytic
Processes, second ed., John Wiley & Sons, Hoboken, 2006, doi: 10.1002/
centage of N2O is still generated on Pt. Models predict N2O forma-
9780471730071.
tion to be negligible on Pd(2 1 1) and Rh(2 1 1), suggesting Pd and [6] J. Pérez-Ramırez, F. Kapteijn, K. Schöffel, J. Moulijn, Formation and control of
Rh may be beneficial to reduce the reaction selectivity to N2O. In N2O in nitric acid production: where do we stand today?, Appl Catal., B 44 (2)
the Ostwald process, Rh-doped Pt gauze needs to be reinstalled (2003) 117–151, https://doi.org/10.1016/S0926-3373(03)00026-2.
[7] N.I. Il’chenko, Catalytic oxidation of ammonia, Russ. Chem. Rev. 45 (12) (1976)
periodically because platinum is lost from the formation of PtO2 1119–1134, https://doi.org/10.1070/RC1976v045n12ABEH002765.
vapor. After the loss of Pt, increased formation of unwanted nitro- [8] Y.F. Zeng, R. Imbihl, Structure sensitivity of ammonia oxidation over platinum,
gen is observed. We show here Rh selectively catalyzes ammonia J. Catal. 261 (2) (2009) 129–136, https://doi.org/10.1016/j.jcat.2008.05.032.
[9] Y. Yao, Z. Huang, P. Xie, S.D. Lacey, R.J. Jacob, H. Xie, F. Chen, A. Nie, T. Pu, M.
oxidation to N2 [53]. The depletion of Pt and enrichment of Rh Rehwoldt, D. Yu, M.R. Zachariah, C. Wang, R. Shahbazian-Yassar, J. Li, L. Hu,
likely leads to the decrease of NO selectivity during operation. Carbothermal shock synthesis of high-entropy-alloy nanoparticles, Science
Surface coverage plays a significant role in determining product 359 (6383) (2018) 1489–1494, https://doi.org/10.1126/science.aan5412.
[10] E.K. Dann, E.K. Gibson, R.H. Blackmore, C.R.A. Catlow, P. Collier, A. Chutia, T.E.
selectivity when formation energies of products are comparable. Erden, C. Hardacre, A. Kroner, M. Nachtegaal, et al., Structural selectivity of
When significant differences among product formation energies supported Pd nanoparticles for catalytic NH3 oxidation resolved using
exist, screening of high-selectivity materials may benefit from combined operando spectroscopy, Nature Catal. 2 (2) (2019) 157, https://doi.
org/10.1038/s41929-018-0213-3.
the comparison of product formation energies. Except for Rh [11] L. Chmielarz, M. Jabłońska, A. Strumiński, Z. Piwowarska, A. Wegrzyn, S.
(2 1 1), the selectivity is highly dependent on the surface coverage Witkowski, M. Michalik, Selective catalytic oxidation of ammonia to nitrogen
on both (2 1 1) and (1 1 1) surfaces of PGMs: NO selectivity over Mg-Al, Cu-Mg-Al and Fe-Mg-Al mixed metal oxides doped with noble
metals, Appl. Catal., B 130–131 (2013) 152–162, https://doi.org/10.1016/j.
increases with surface O but decreases with surface N. On Rh
apcatb.2012.11.004.
(2 1 1), however, high N2 selectivity is observed even at high O cov- [12] Y. Li, J.N. Armor, Selective NH3 oxidation to N2 in a wet stream, Appl. Catal., B
erage, because the N2(g) formation barrier is significant less than 13 (2) (1997) 131–139, https://doi.org/10.1016/S0926-3373(96)00098-7.
the NO(g) barrier. It is therefore promising to search for catalysts [13] J. Pérez-Ramírez, E.V. Kondratenko, G. Novell-Leruth, J.M. Ricart, Mechanism of
ammonia oxidation over PGM (Pt, Pd, Rh) wires by temporal analysis of
with high NO(g) formation barrier and low N2(g) formation barrier products and density functional theory, J. Catal. 261 (2) (2009) 217–223,
to increase the N2 selectivity for ammonia slip control. https://doi.org/10.1016/j.jcat.2008.11.018.
330 H. Ma, W.F. Schneider / Journal of Catalysis 383 (2020) 322–330

[14] T. Pignet, L. Schmidt, Kinetics of NH3 oxidation on Pt, Rh, and Pd, J. Catal. 40 (2) [36] L.H. Sprowl, C.T. Campbell, L. Arnadottir, Hindered translator and hindered
(1975) 212–225, https://doi.org/10.1016/0021-9517(75)90249-3. rotor models for adsorbates: partition functions and entropies, J. Phys. Chem. C
[15] H. Ma, W.F. Schneider, Structure- and temperature-dependence of Pt- 120 (18) (2016) 9719–9731, https://doi.org/10.1021/acs.jpcc.5b11616.
catalyzed ammonia oxidation rates and selectivities, ACS Catal. 9 (3) (2019) [37] L.C. Grabow, Computational Catalyst Screening, in: A. Asthagiri, M.J. Janik
2407–2414, https://doi.org/10.1021/acscatal.8b04251. (Eds.), Computational Catalysis, Royal Society of Chemistry, London, 2013, pp.
[16] R. Imbihl, A. Scheibe, Y. Zeng, S. Günther, R. Kraehnert, V. Kondratenko, M. Baerns, 1–58.
W. Offermans, A. Jansen, R. Van Santen, Catalytic ammonia oxidation on platinum: [38] W.D. Mieher, W. Ho, Thermally activated oxidation of NHs on Pt(1 1 1):
mechanism and catalyst restructuring at high and low pressure, Phys. Chem. intermediate species and reaction mechanisms, Surf. Sci. 322 (1–3) (1995)
Chem. Phys. 9 (27) (2007) 3522–3540, https://doi.org/10.1039/b700866j. 151–167, https://doi.org/10.1016/0039-6028(95)90026-8.
[17] W. Offermans, A. Jansen, R. van Santen, Ammonia activation on platinum 111: [39] C.J. Weststrate, J.W. Bakker, A.C. Gluhoi, W. Ludwig, B.E. Nieuwenhuys,
a density functional theory study, Surf. Sci. 600 (9) (2006) 1714–1734, https:// Ammonia oxidation on Ir (1 1 1): why ir is more selective to N2 than Pt,
doi.org/10.1016/j.susc.2006.01.031. Catal. Today 154 (1) (2010) 46–52, https://doi.org/10.1016/
[18] G. Novell-Leruth, J.M. Ricart, J. Pérez-Ramírez, Pt(100)-Catalyzed ammonia j.cattod.2010.03.049.
oxidation studied by DFT: mechanism and microkinetics, J. Phys. Chem. C 112 [40] S. Gunther, A. Scheibe, H. Bluhm, M. Haevecker, E. Kleimenov, A. Knop-Gericke,
(35) (2008) 13554–13562, https://doi.org/10.1021/jp802489y. R. Schlogl, R. Imbihl, In situ X-ray photoelectron spectroscopy of catalytic
[19] C. Popa, R.A. Van Santen, A.P.J. Jansen, Density-functional theory study of NHx ammonia oxidation over a Pt (533) surface, J. Phys. Chem. C 112 (39) (2008)
oxidation and reverse reactions on the Rh(1 1 1) surface, J. Phys. Chem. C 111 15382–15393, https://doi.org/10.1021/jp803264v.
(27) (2007) 9839–9852, https://doi.org/10.1021/jp071072g. [41] N. López, M. García-Mota, J. Gómez-Díaz, NH3 oxidation on oxygen-
[20] M. Rafti, J.L. Vicente, A. Albesa, A. Scheibe, R. Imbihl, Modeling ammonia precovered Au(1 1 1): a density functional theory study on selectivity, J.
oxidation over a Pt (533) surface, Surf. Sci. 606 (1–2) (2012) 12–20, https://doi. Phys. Chem. C 112 (1) (2008) 247–252, https://doi.org/10.1021/jp077205f.
org/10.1016/j.susc.2011.08.014. [42] Z. Jiang, P. Qin, T. Fang, Mechanism of ammonia decomposition on clean and
[21] G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy calculations for oxygen-covered Cu (1 1 1) surface: a DFT study, Chem. Phys. 445 (2014) 59–
metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci 6 67, https://doi.org/10.1016/j.chemphys.2014.10.018.
(1) (1996) 15–50, https://doi.org/10.1016/0927-0256(96)00008-0. [43] G. Novell-Leruth, A. Valcárcel, J. Pérez-Ramírez, J.M. Ricart, Ammonia
[22] G. Kresse, J. Furthmüller, Efficient iterative schemes for ab initio total-energy dehydrogenation over platinum-group metal surfaces. structure, stability,
calculations using a plane-wave basis set, Phys. Rev. B 54 (1996) 11169– and reactivity of adsorbed NHx species, J. Phys. Chem. C 111 (2) (2007) 860–
11186, https://doi.org/10.1103/PhysRevB.54.11169. 868, https://doi.org/10.1021/jp064742b.
[23] P.E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50 (24) (1994) [44] H. Falsig, J. Shen, T.S. Khan, W. Guo, G. Jones, S. Dahl, T. Bligaard, On the
17953–17979, https://doi.org/10.1103/PhysRevB.50.17953. structure sensitivity of direct NO decomposition over low-index transition
[24] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector metal facets, Top. Catal. 57 (1–4) (2014) 80–88, https://doi.org/10.1007/
augmented-wave method, Phys. Rev. B 59 (3) (1999) 1758–1775, https://doi. s11244-013-0164-5.
org/10.1103/PhysRevB.59.1758. [45] J.P. Clay, J.P. Greeley, F.H. Ribeiro, W.N. Delgass, W.F. Schneider, DFT
[25] J. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made comparison of intrinsic WGS kinetics over Pd and Pt, J. Catal. 320 (1) (2014)
simple, Phys. Rev. Lett. 77 (18) (1996) 3865–3868, https://doi.org/10.1103/ 106–117, https://doi.org/10.1016/j.jcat.2014.09.026.
PhysRevLett. 77.3865. [46] Z.-P. Liu, S.J. Jenkins, D.A. King, Step-enhanced selectivity of NO reduction on
[26] G. Henkelman, B.P. Uberuaga, H. Jónsson, Climbing image nudged elastic band platinum-group metals, J. Am. Chem. Soc. 125 (48) (2003) 14660–14661,
method for finding saddle points and minimum energy paths, J. Chem. Phys. https://doi.org/10.1021/ja0372208.
113 (22) (2000) 9901–9904, https://doi.org/10.1063/1.1329672. [47] J.K. Nørskov, F. Studt, F. Abild-Pedersen, T. Bligaard, Fundamental Concepts in
[27] G. Henkelman, H. Jónsson, A dimer method for finding saddle points on high Heterogeneous Catalysis, John Wiley & Sons, 2014.
dimensional potential surfaces using only first derivatives, J. Chem. Phys. 111 [48] C.T. Campbell, The degree of rate control: a powerful tool for catalysis
(15) (1999) 7010–7022, https://doi.org/10.1063/1.480097. research, ACS Catal. 7 (4) (2017) 2770–2779, https://doi.org/10.1021/
[28] F. Birch, Finite elastic strain of cubic crystals, Phys. Rev. 71 (11) (1947) 809– acscatal.7b00115.
824, https://doi.org/10.1103/PhysRev. 71.809. [49] H. Ma, S. Li, H. Wang, W.F. Schneider, Water-mediated reduction of aqueous N-
[29] F.D. Murnaghan, The compressibility of media under extreme pressures, Proc. nitrosodimethylamine with Pd, Environ. Sci. Technol. 53 (13) (2019) 7551–
Natl. Acad. Sci. USA 30 (9) (1944) 244–247, https://doi.org/10.1073/ 7563, https://doi.org/10.1021/acs.est.9b01425.
pnas.30.9.244. [50] S. Shrestha, M.P. Harold, K. Kamasamudram, Experimental and modeling study
[30] A.W. Hull, X-ray crystal analysis of thirteen common metals, Phys. Rev. 17 (5) of selective ammonia oxidation on multi-functional washcoated monolith
(1921) 571–588, https://doi.org/10.1103/PhysRev. 17.571. catalysts, Chem. Eng. J. 278 (2015) 24–35, https://doi.org/10.1016/j.
[31] H. Monkhorst, J. Pack, Special points for brillouin zone integrations, Phys. Rev. cej.2015.01.015.
B 13 (12) (1976) 5188–5192, https://doi.org/10.1103/PhysRevB.13.5188. [51] C. Paolucci, J.R. Di Iorio, F.H. Ribeiro, R. Gounder, W.F. Schneider, Catalysis
[32] W. Offermans, A. Jansen, R. Van Santen, G. Novell-Leruth, J. Ricart, J. Perez- science of NOx selective catalytic reduction with ammonia over Cu-SSZ-13 and
Ramirez, Ammonia dissociation on Pt {100}, Pt {111}, and Pt {211}: a Cu-SAPO-34, in: Advances in Catalysis, vol. 59, 2016, pp. 1–4. doi: 10.1016/bs.
comparative density functional theory study, J. Phys. Chem. C 111 (47) acat.2016.10.002.
(2007) 17551–17557, https://doi.org/10.1021/jp073083f. [52] A. Scheuer, W. Hauptmann, A. Drochner, J. Gieshoff, H. Vogel, M. Votsmeier,
[33] J. Dumesic, The Microkinetics of Heterogeneous Catalysis, ACS Professional Dual layer automotive ammonia oxidation catalysts: experiments and
Reference Book, American Chemical Society, Washington, 1993. computer simulation, Appl. Catal., B 111–112 (2012) 445–455, https://doi.
[34] L. Foppa, M.-C. Silaghi, K. Larmier, A. Comas-Vives, Intrinsic reactivity of Ni, Pd org/10.1016/j.apcatb.2011.10.032.
and Pt surfaces in dry reforming and competitive reactions: insights from first [53] D. Waller, M.S. Grønvold, N. Sahli, Ammonia Oxidation Catalyst for the
principles calculations and microkinetic modeling simulations, J. Catal. 343 Production of Nitric Acid Based on Metal Doped Yttrium Ortho Cobaltate, US
(2016) 196–207, https://doi.org/10.1016/j.jcat.2016.02.030. Patent 9,403,681 (Aug. 2 2016).
[35] M.W. Chase Jr., NIST-JANAF Thermochemical Tables, Journal of Physical and
Chemical Reference Data Monograph No.9, National Institute of Standards and
Technology, New York, 1998.

You might also like