transition metal doping of pd(111) for the no + co ... · transition metal doping of...

11
Transition metal doping of Pd(1 1 1) for the NO+CO reaction Citation for published version (APA): Zhang, L., Filot, I. A. W., Su, Y., Liu, J., & Hensen, E. J. M. (2018). Transition metal doping of Pd(1 1 1) for the NO+CO reaction. Journal of Catalysis, 363, 154-163. https://doi.org/10.1016/j.jcat.2018.04.025 Document license: CC BY DOI: 10.1016/j.jcat.2018.04.025 Document status and date: Published: 01/07/2018 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 07. Sep. 2020

Upload: others

Post on 18-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

Transition metal doping of Pd(1 1 1) for the NO+CO reaction

Citation for published version (APA):Zhang, L., Filot, I. A. W., Su, Y., Liu, J., & Hensen, E. J. M. (2018). Transition metal doping of Pd(1 1 1) for theNO+CO reaction. Journal of Catalysis, 363, 154-163. https://doi.org/10.1016/j.jcat.2018.04.025

Document license:CC BY

DOI:10.1016/j.jcat.2018.04.025

Document status and date:Published: 01/07/2018

Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:www.tue.nl/taverne

Take down policyIf you believe that this document breaches copyright please contact us at:[email protected] details and we will investigate your claim.

Download date: 07. Sep. 2020

Page 2: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

Journal of Catalysis 363 (2018) 154–163

Contents lists available at ScienceDirect

Journal of Catalysis

journal homepage: www.elsevier .com/locate / jcat

Transition metal doping of Pd(111) for the NO + CO reaction

https://doi.org/10.1016/j.jcat.2018.04.0250021-9517/� 2018 The Author(s). Published by Elsevier Inc.This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

⇑ Corresponding author.E-mail address: [email protected] (E.J.M. Hensen).

Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun Liu, Emiel J.M. Hensen ⇑Laboratory of Inorganic Materials Chemistry, Schuit Institute of Catalysis, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands

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

Article history:Received 4 March 2018Revised 19 April 2018Accepted 21 April 2018Available online 9 May 2018

Keywords:DFT calculationNO + CO reactionTransition metal dopingPd(111)

The replacement of platinum group metals by non-noble metals has attracted significant attention in thefield of three-way catalysis. Here, we use DFT calculations to comprehensively study NO reduction by COand CO oxidation on Pd(111) and transition metal doped Pd(111). Whilst direct NO dissociation is verydifficult on metallic Pd(111), doping with transition metals can substantially lower the reaction barrierfor NO dissociation. The lowest barrier is predicted for Ti-doped Pd(111). An electronic structure analysisshows that the low barrier is due to the strong adsorption of N and O on surface sites involving Ti atoms.It relates to strong hybridization of the N and O orbitals with the half-filled d-band of the metallic surface.At the same time, the anti-bonding states are shifted above the Fermi level, which further strengthens theadsorption of N and O. A Brønsted-Evans-Polanyi relation for NO dissociation on TM-doped Pd(111)surfaces is identified. The complete reaction pathway for N2, N2O and CO2 formation on Pd(111) andTi-doped Pd(111) was considered. Besides more facile NO dissociation, the energy barrier for CO oxida-tion is decreased for the Ti-doped surface. Microkinetics simulations confirm that the activity and selec-tivity for NO reduction and CO oxidation are drastically improved after Ti doping. Our findings indicatethat doping of Pd with non-noble metal can further improve the performance of three-way catalysts.

� 2018 The Author(s). Published by Elsevier Inc. This is an open access article under the CC BY license(http://creativecommons.org/licenses/by/4.0/).

1. Introduction

Nitric oxide (NO), carbon monoxide (CO) and unburnt hydrocar-bons are major pollutants emitted from automobile exhaust. Thesetoxic gases are harmful to human health and the environment.Therefore, the catalytic removal of these exhaust gases has becomean important technology [1,2]. Current three-way catalysts (TWCs)consist of typical platinum-group metals, such as Pt, Pd and Rh,which promote NO reduction and CO and hydrocarbon oxidationreactions [3]. Previous reports showed that the activity of thesemetals for the NO dissociation differs significantly. In general, Rhand Pt nanoparticles are the more active components for NO reduc-tion in TWCs. Especially Rh-based catalysts can convert NOx intoN2 with high activity and selectivity at relatively low temperature[4,5]. Early DFT computations showed that the energy barriers ofNO dissociation on Rh(100) and Rh(111) are 0.48 eV and 1.53eV, respectively. These barriers are much lower than the barrierof 2.44 eV reported for Pd(111) [6]. Eichler and Hafner predicteda barrier of 1.21 eV for NO dissociation on Pt(100) as the rate-limiting step in the NO + CO reaction [7].

While for a long time the combination of Pt and Rh has been uti-lized to simultaneously reduce NO and oxidize CO [8–11], the high

price of these noble metals has driven research to employ cheapertransition metals (TM). The lower price has been a driver to replacePt by Pd in TWCs, although recently Pt and Pd prices are compara-ble Pd-based catalysts have gained wider spread interest from theacademic community in recent years [12–15]. The possibility ofalloying a noble metal like Pd with cheaper TMs has also beenexplored. Lopez and Nørskov investigated synergetic effects in COadsorption on Cu-doped Pd(111) alloys and found that differencesin adsorption energies can be correlated to changes in the elec-tronic structure [16]. Yang and co-workers studied NO adsorptionand dissociation on neutral and charged TM-doped Pd clusters[17]. Sautet’s group investigated the structure sensitivity of NO dis-sociation on Pd surfaces and computed NO dissociation barriers of2.44 eV and 1.63 eV for Pd(111) and Pd(100), respectively. NO dis-sociation on the stepped Pd(511) surface also involved a relativelyhigh barrier of 1.54 eV [6]. The higher barriers for Pd compared toRh are consistent with the preference to use Rh for achieving goodNO reduction activity in practical TWCs.

Selective catalytic reduction (SCR) of NO requires a reductantsuch as CO, NH3, H2 or CxHy [18,19]. In the context of TWC, NOreduction with CO or H2 has been extensively studied[9,10,20,21]. Paredis et al. explored the evolution of the structureand oxidation state of Pd nanoparticles supported on ZrO2 duringNO reduction by H2 [22]. Liu et al. reported that NO dissociationby H2 on Pd(111) and their results indicated that the N–O bond

Page 3: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163 155

scission can be facilitated via H-assisted reaction pathway, withthe activation barrier of 1.58 eV. [23]. Molecular beam studies havealso been used to study the NO + CO reaction on Pd(111) [24].Goodman et al. explored the structure sensitivity of the NO + COreaction on Pd(100) and Pd(111) [25]. They observed that Pd(111) was about five times more active than Pd(100) for NO +CO reaction. In general, it can be stated that Pd is not active enoughfor NO dissociation, even in the presence of H2. This relates to theweak activation of NO on Pd surface and, possibly, the poisoningeffect of N and O atoms produced by NO dissociation [26]. Recentreports provide new directions to the field of TWC [27,28]. Hamand co-authors reported that small Pd ensembles in AuPd alloysfacilitate CO oxidation [29]. Cheng et al. found that the activity ofAuPd alloys for CO oxidation can be tuned by changing the compo-sition [30]. These results suggest that the CO oxidation activity ofPd can be improved by TM doping. A comprehensive overview ofthe effect of TM doping on the NO + CO reaction is however lacking.

In this work, we use density functional theory (DFT) calcula-tions combined with microkinetics simulations of the modelTWC NO + CO reaction to investigate the influence of TM dopingof Pd(111). We use the Pd(111) surface as a reference as it isthe dominant surface of Pd nanoparticle catalysts [31,32]. We firstinvestigate the adsorption of atomic and molecular species on Pd(111) and Ti-doped Pd(111), followed by a mechanistic study ofdirect NO dissociation on various TM-doped Pd(111). A strongBrønsted-Evans-Polanyi (BEP) correlation of the NO dissociationbarrier as function of the N and O adsorption energies will be dis-cussed. We computed the complete potential energy diagram forthe NO + CO reaction on Pd(111) and Ti-doped Pd(111). Theresults show that Ti doping not only improves NO dissociationbut also facilitates N atom recombination as well as the oxidationof CO with the O atom originating from NO dissociation. Microki-netics simulations revealed that the activity and N2 selectivity ofPd in the NO + CO reaction can be drastically improved by Tidoping.

2. Computational methods

2.1. DFT calculations

We performed spin-polarized DFT calculations by using the pro-jector augmented wave (PAW) [33] method as implemented in theVienna Ab Initio Simulation Package (VASP) [34,35]. The Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional [36] wasused. The cut-off energy for the plane-wave basis was set to 400eV. Partial occupancies were determined by the first-orderMethfessel-Paxton scheme with a smearing width of 0.2 eV. Theoptimized lattice constant of bulk Pd was 3.95 Å, which agrees wellwith previous results [23,37]. In order to model the Pd(111) sur-face, we constructed a 2 � 2 unit cell with five atomic layers, givinga slab thickness of 9.15 Å. The top two layers were relaxed and thebottom three layers were frozen to the configuration of the bulk. Toevaluate the influence of lateral interactions between periodicimages, we calculated for some elementary reaction steps the tran-sition states in a larger 3 � 3 unit cell. The energy difference ofcomputed activation barriers between the 2 � 2 and 3 � 3 unitcells was negligible (i.e., less than 0.01 eV/atom). To avoid spuriousinteractions of adsorbates between neighbouring super cells, a vac-uum thickness of 12 Å was used. A Monkhorst-Pack mesh with a k-point sampling of 5 � 5 � 1 was used for the Brillouin zone inte-gration. For the doped Pd(111) model, a Pd atom in the top layerwas substituted by a transition metal (TM = Ti, Pt, Fe, Au, Ag, Cu,Ni), resulting in a surface doping content of 25%. The doping ofPd(111) in this way with a TM atom is indicated by TM-Pd(111). To investigate the stability of these TM-Pd(111) surfaces,

we calculated the exchange energy of the doped surfaces (seeFig. S1). It was thus found that the doped TM atoms are stronglybound in the Pd metal surface. It is noteworthy that PdTi alloy witha controllable bimetallic ratio was able to fabricated via a milddealloying process [38]. Besides, recent studies show that Pd-based alloys were stable [39,40]. For the Ti and Fe atom, aHubbard-like term describing the on-site Coulombic interactionswas introduced and set to Ueff = 4.5 eV and 4 eV, respectively. Thisapproach was previously reported to provide a better descriptionof localized states for strongly correlated system [41,42]. We alsotested the effect of U on Ti-doped and Fe-doped system and foundthat the U term has limited influence on energy difference and acti-vation barrier. The systems were assumed to be converged whenthe Hellmann-Feynman forces were less than 0.05 eV/Å. To studythe reaction mechanism, we calculated the location and energyof transition states by the climbing-image nudged elastic band(CI-NEB) method [43,44].

Adsorption energies are computed by

Eads ¼ Em þ surf � ðEsurf þ EmÞ ð1Þwhere Em þ surf , Esurf and Em are the total energies of the adsorbedsystem, the empty surface and the corresponding gas phase species,respectively. Repulsion was also considered in our study, which isdefined as the difference between the total energy of the co-adsorbed species and their corresponding configuration at infiniteseparation on the catalytic surface.

2.2. Microkinetics modelling

Using the stable and transition states identified for the exploredreaction mechanisms, we carried out microkinetics simulations todetermine the reaction rate and the product distribution. For sur-face reactions, the computed activation energies are used to esti-mate the forward and backward rate constant using the Eyringequation:

k ¼ kbTh

QTS

Qe� EakbT ð2Þ

Herein k is the reaction rate constant, kb and h the Boltzmannand Planck’s constants, respectively, T the temperature (in K),and Ea the electronic activation energy (in J). QTS and Q refer tothe partition functions of the transition state and the ground state,respectively. As an approximation, we assumed that all vibrationalpartition functions equal unity. This leads to a pre-factor for all sur-face elementary reaction steps of �1013 s�1.

For adsorption reactions, we assumed that the molecule losesone of its translational degrees of freedom with respect to thegas phase. Therefore, the rate for molecular adsorption was definedas:

kads ¼ PA0ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2pmkbT

p S ð3Þ

where P is the partial pressure of the adsorbate in the gas phase,A0the surface area of the adsorption site, and m and S the mass ofthe adsorbate and its sticking coefficient, respectively.

For desorption, we assumed that the activated complex has twotranslational and three rotational degrees of freedom. Accordingly,the rate of molecular desorption is defined as [45]

kdes ¼ kbT3

h3

A0ð2pmkbÞrhrot

e�Edes

kbT ð4Þ

where r indicates the symmetry number, h refers to the character-istic temperature for rotation, and Edes is the desorption energy.

The details for the microkinetics simulations have beendescribed in our previous work [46,47] and is briefly discussed

Page 4: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

156 L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163

here for clarity. For all surface reaction intermediates, the differen-tial equations were constructed by using the rate constants of theelementary reaction steps. For each of the M components involvedin the reaction network, a single differential equation is written as:

ri ¼XNj¼1

kjm ji

YMk¼1

ckmjk

!ð5Þ

where kj is the elementary reaction rate constant, m ji refers to the

stoichiometric coefficient of component i in elementary reactionstep k and ck is the concentration of component k on the reactionsurface.

To identify the elementary steps that control the overall reac-tion rate of the NO + CO reaction, Campbell’s degree of rate control(DRC) analysis [48–50] was used. For a specific elementary step i,the degree of rate control coefficientvRC;i is determined by

vRC;i ¼kir

@r@ki

� �kj–i ;Ki

¼ @lnr@lnki

� �kj–i ;Ki

ð6Þ

In the above equation, r indicates the overall reaction rate andki, and Ki represent the forward rate and the equilibrium constantsfor step i, respectively.

The DFT-based microkinetics calculations were performedusing the MKMCXX program [51]. The overall conversion rates ofthe NO + CO reaction, steady-state coverages and product distribu-tion were computed as a function of temperature by integrating

Table 1Adsorption energetics and geometries of surface intermediates relevant to the NO + CO re

Reaction species Configurations Bond lengths (Å)

Pd(111)

NO fcc (N-end) d(Pd–N) = 2.04NO1 hcp (N-end) d(Pd–N) = 2.04N2 Bridge d(Pd–N) = 2.12N2O Bridge d(Pd–N) = 2.07N fcc d(Pd–N) = 1.90N1 hcp d(Pd–N) = 1.91O fcc d(Pd–O) = 2.00O1 hcp d(Pd–O) = 2.00CO fcc d(Pd–C) = 2.12CO2 fcc d(Pd–C) = 2.13

Fig. 1. Potential energy diagram and associated structures for

the ordinary differential equations with respect to time using thebackward differentiation formula method [52–54].

3. Results and discussion

3.1. Atomic and molecular species adsorption on Pd(111) and Ti-Pd(111)

We investigated the adsorption properties of all relevant inter-mediates in the NO + CO reaction for Pd(111) and Ti-Pd(111) sur-face models. Table 1 summarizes the adsorption energies andgeometries of the relevant reaction intermediates. For Pd(111),NO will bind at fcc and hcp sites (at the N-terminus) with adsorp-tion energies of �2.19 eV and �2.17 eV, respectively. The Pd–Nbond length is 2.04 Å for both surfaces. The vibrational frequencyof adsorbed NO molecule is 1584 cm�1 at the fcc site with an N–O bond length of 1.21 Å. These values are in good agreement withliterature data [23,55]. NO adsorption on Ti-Pd(111) was slightlyless favourable with values of �2.03 eV and �1.95 eV on fcc andhcp sites, respectively. Correspondingly, the Ti-N bond lengthsare somewhat longer: 2.22 Å for the fcc site and 2.12 Å for thehcp site. The corresponding stretching frequency for the NO mole-cule is around 1533 cm�1 at the fcc site of Ti-Pd(111). Clearly, dop-ing the Pd(111) surface with Ti results in a red-shift of the N–Ostretching frequency, indicative of a weakening of the N–O bondwith respect to NO adsorbed on Pd(111) and consistent with the

action on Pd(111) and Ti-Pd(111).

Eads (eV)

Ti-Pd(111) Pd(111) Ti-Pd(111)

d(Ti-N) = 2.22 �2.19 �2.03d(Ti-N) = 2.12 �2.17 �1.95d(Ti-N) = 2.16 �0.08 �0.23d(Ti-N) = 2.06 �0.19 �0.07d(Ti-N) = 1.91 �4.73 �4.75d(Ti-N) = 1.90 �4.63 �4.67d(Ti-O) = 1.85 �4.55 �5.73d(Ti-O) = 1.85 �4.35 �5.65d(Pd–C) = 2.05 �1.63 �1.57d(Pd–C) = 2.06 �0.03 �0.71

the direct dissociation of NO on Pd(111) and Ti-Pd(111).

Page 5: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163 157

increased N–O bond length of 1.22 Å. This result suggests that Tidoping will result in a more facile NO bond dissociation. N2 bindsweakly at the bridge site of Pd(111) (Eads = �0.08 eV), which is inagreement with previous studies of Huai et al., who calculated anadsorption energy of �0.17 eV [23]. Similar to N2 adsorption, theN2O molecule has a small adsorption energy of �0.19 eV. Our cal-culated adsorption energy is very close to the previous work of Weiand co-workers, who reported an adsorption energy of �0.14 eV[56]. CO prefers to adsorb at an fcc site on oxygen pre-coveredPd(111) and Ti-Pd(111). The adsorption energies are �1.63 eV

Fig. 2. (a) Partial density of states (PDOS) analysis of N and O in the initial, transition andthe PDOS of N, and the panels on the right to the PDOS of O. (b) PDOS analysis of N and Opanels on the left represent to the PDOS of N, and the panels on the right to the PDOS o

and �1.57 eV, respectively. The Pd–C bond is 2.12 Å on the Pd(111) surface, which is slightly longer than Ti-Pd(111) (2.05 Å).

Atomic N adsorbs strongly on the metal surfaces. Its adsorptionenergy, with respect to gaseous N, is �4.73 eV and �4.63 eV at fccand hcp sites, respectively. The Pd–N bond length is �1.90 Å. ForTi-Pd(111), N adsorbs even more strongly with energies of�4.75 eV and �4.67 eV, respectively for fcc and hcp sites. For Oadsorption, a similar trend was found. For Pd(111), an O atomadsorbs with energies of �4.55 eV and �4.35 eV on fcc and hcpsites, respectively. For Ti-Pd(111), O adsorption is much more

final states for the NO dissociation on Pd(111). The panels on the left represent toin the initial, transition and final states for the NO dissociation on Ti-Pd(111). Thef O.

Page 6: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

Fig. 3. COHP bonding analysis. Pd–N and Ti–N interactions in initial states, transition states and final states. The upper panels show the situation in Pd(111), whereas thelower panels correspond to the Ti-Pd(111). The Fermi levels were set to the energy zero. Bonding state interactions to the right and antibonding state interactions to the left.Palladium is shown in dark cyan, the Palladium atom involved in Pd–N interaction is yellow, titanium in cyan, nitrogen in blue, and oxygen in red.

158 L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163

exothermic with �5.73 eV for the fcc site and �5.65 eV for the hcpsite. Clearly, the impact of Ti doping is greater for O adsorptionthan for N adsorption.

3.2. Direct NO dissociation

NO dissociation is an important elementary reaction step incar exhaust clean-up catalysis, as it is a critical step in the overallreaction to reduce NO into N2. We choose NO adsorbed in itsmost stable adsorption mode as the initial state (IS) for directNO dissociation, which is in a threefold coordination on a fcc site.In the transition state (TS), the N and O atoms migrate to adja-cent bridge sites. In the final state (FS), the N and O atoms arebound in adjacent fcc sites of Pd(111). The reaction barrier forthis process is 2.32 eV (Fig. 1) and the reaction is endothermicby 1.12 eV. This calculated energy barrier is very close to previ-ous result [6]. Exploring the same pathway on Ti-Pd(111)resulted in an activation barrier of only 0.65 eV. It is importantto mention that this barrier is much lower than previouslyobtained barriers for NO dissociation on open Pd(100) andstepped Pd(511) surfaces [6].

In order to understand the origin of the enhanced activitytowards direct NO dissociation upon Ti doping, the electronicstructure of the surfaces was analysed in more detailed by a partialdensity of states (PDOS) and crystal orbital Hamilton population(COHP) analysis [57,58]. Fig. 2a shows the orbital-resolved PDOSfor N and O in the initial, transition and final states. The features

in the PDOS were identified by comparison with the PDOS of NOplaced in an empty simulation cell (Fig. S2). We will first discussthe PDOS of NO dissociation on Pd(111) and then contrast theinsights with corresponding data for Ti-Pd(111).

In Fig. 2a, it can be seen that in the initial state the 5r, 1p and4rmolecular orbitals of NO are respectively located around �7.76eV, �8.01 eV and �13.34 eV with respect to the Fermi level. In freeNO, the 2p orbital of NO is only partially filled, whereas in theadsorbed state it lies far below the Fermi-level (i.e., at �2.55 eV).This shows that there is back–donation from the filled d-orbitalsof the Pd metal to the partially filled 2p molecular orbital. In thetransition state, the N and O atoms are only weakly bound to eachother, as can be seen from the disappearance of the 5r and 1pmolecular orbitals.

Instead, two p bands appear between�5.02 eV and�7.11 eV forthe N and O atom, respectively. These p bands correspond to theatomic orbitals of N and O. The PDOS for the final state closelyresembles the PDOS of the transition state, indicative of a late tran-sition state. This is also apparent from the similar geometries of thetransition and final state. The late character of the transition stateis further confirmed by the linear coefficient of the BEP analysis(vide infra).

Fig. 2b depicts the PDOS of Ti-Pd(111). As compared to Pd(111), the 5r, 1p and 4r orbitals of molecular NO are shiftedtowards the Fermi level in the initial state. This is evident fromthe fact that Ti, a d2 metal, has fewer electrons than Pd, which isa d10 metal. The px and py orbitals corresponding to the 2p

Page 7: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163 159

molecular orbital are more delocalized, which indicates a strongerorbital overlap between NO and the d-band as well as a strongerback-donation from Ti-Pd(111) as compared to Pd(111). Thiscan be further rationalized by the stronger ability of Ti to acceptelectrons from the 5r molecular orbital and consequently todonate electrons in the 2p orbital. Similar to NO dissociation onPd(111), the PDOS shows that the 5r, 1p and 4r molecular orbi-tals disappear in the transition and final state.

To analyse the bonding nature of the different states, a COHPanalysis was conducted. In Fig. 3, the Pd–N interaction is shown.It can be seen that there are a number of anti-bonding states closeto the Fermi-level for the transition and final states for NO dissoci-ation on Pd(111). In comparison, these states are bonding in nat-ure for Ti-Pd(111). The COHP analysis for the interactionbetween the next-nearest neighbour Pd atom on the surface andadsorbed N is shown (Fig. S3). Herein, a similar trend is seen, indi-cating that the adsorption of N atom is strengthened by Ti doping.The same kind of analysis was conducted for the interaction of Pd

Fig. 5. (a) Brønsted–Evans–Polanyi relation for NO dissociation on transition metal surenergy barrier is linearly related to the sum of the N and O adsorption energies.

Fig. 4. Partial density of states (PDOS) analysis of the first atomic layer in the initial, trans(right panels).

with O, further confirming the observed trend (Fig. S4). For Pd(111), there are several anti-bonding states around the Fermi-level, whereas for the Ti-doped case, these Ti-O states are bonding.Again, this shows that substitution of a Pd surface atom with a Tiatom results in a stronger interaction between the metal surfaceand the N and O atoms and in turn this results in a lower activationbarrier for NO dissociation.

In Fig. 4, the averaged PDOS of the first atomic layer of the sur-face is shown. For Pd(111) in the initial state, the majority of the d-states lie below the Fermi level. For the corresponding Ti-Pd(111)case, a large number of d-states are located above the Fermi leveland are thus empty. These empty d-states are able to accept elec-trons and they will, via r-donation and p-back-donation, weakenthe N–O molecular bond. The DOS analyses are consistent withthe COHP analyses. Both indicate that Ti doping can effectivelymodulate the electronic structure of Pd metal surface, whichstrengthens the back-donation effect and facilitates the NOdissociation.

faces. The energy barrier is linearly related to the NO dissociation energy. (b) The

ition and final states for the NO dissociation on Pd(111) (left panels) and Ti-Pd(111)

Page 8: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

160 L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163

Next, we carried out similar transition state calculations for awider range of TM-doped Pd(111) surfaces. The activation barrierfor direct NO dissociation is plotted against the NO dissociationreaction energy in Fig. 5a. It emphasizes the existence of a BEPrelation for NO dissociation with a linear scaling parameter of a= 0.86. This high BEP value implies that NO dissociation occursvia a late transition state, which is common dissociation reactions[59]. The activation barrier depends in a similar manner on theatomic adsorption energy (Fig. 5b). This result is not surprising,as NO adsorption energies are relatively independent of the transi-tion metal dopant. Therefore, the reaction enthalpy is mainlydetermined by the differences in adsorption energies in the finalstate. The correlation in Fig. 5b can be used to explore other cata-lyst compositions without having to explicitly perform a transitionstate search.

The COHP analysis in Fig. 6 concerns the metal-N interaction,which is presented in descending order of the NO dissociation bar-rier. It clearly shows that with increasing energy of the anti-bonding states, the metal-atom binding strength increases andthe corresponding barrier for NO dissociation decreases. In otherwords, the anti-bonding states are shifted above the Fermi level,which strengthens the adsorption of the atoms and decreases thereaction barriers.

3.3. N2 and N2O formation

After NO dissociation, there are two competing pathways thatlead to either N2 or N2O. In the first pathway, two N atoms recom-bine to form N2. In the second one, a N atom reacts with anadsorbed NO molecule to N2O. In the following, we compare thesetwo reaction routes for N2 and N2O formation on Pd(111) and Ti-

Fig. 6. COHP bonding analysis. TM-N interactions in final states of NO dissociation oninteractions to the right and antibonding state interactions to the left.

Pd(111). In Fig. 7, the reaction energy diagram is shown. The bar-rier for N + N recombination to form N2 is 0.99 eV and the processis strongly exothermic by 2.17 eV.

Subsequently, N2 can readily desorb from the surface as itsadsorption energy is only 0.08 eV. The activation barrier for N2

formation on Ti-Pd(111) is only 0.55 eV; this reaction is exother-mic by 2.00 eV. It is noteworthy that this barrier is much lowerthan the N2 recombination barrier of 1.60 eV on Pd(100). N2

adsorbs slightly stronger on Ti-Pd(111) (0.23 eV) than on Pd(111).

The other product of NO reduction is N2O. In order for NO and Nto react, they need to migrate to adjacent fcc sites on the surface.This migration is energetically unfavourable and the migrationenergy with respect to NO and N infinitely far apart is 0.90 eVand 1.09 eV for Pd(111) and Ti-Pd(111), respectively. The energybarriers for the formation of N2O by reaction of N + NO on Pd(111) and Ti-Pd(111) are 1.00 eV and 0.77 eV, respectively(Fig. 7). N2O desorption is facile in both cases due to its weak bind-ing with the surface. The energy barrier for N2O formation is veryclose to the energy barrier for N2 formation on Pd(111). On theother hand, the energy barrier for N2O formation is much higherthan for N2 formation on Ti-Pd(111). Thus, we infer that Ti dopingof Pd will result in enhanced selectivity to N2 over N2O.

3.4. CO oxidation on oxygen pre-covered surfaces

Besides N2 and N2O formation, another important pathway isthe formation of CO2. Here, we explored the effect of Ti dopingon CO oxidation on oxygen pre-covered Pd(111). In our model(Fig. 8), the initial state is represented by co-adsorbed molecularCO and atomic O on Pd(111). The reaction proceeds through a

TM-doped Pd(111). The Fermi levels were set to the energy zero. Bonding state

Page 9: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

Fig. 8. Potential energy diagram and structures for CO oxidation on Pd(111) and Ti-Pd(111).

Fig. 7. Potential energy diagram and structures for the N2 and N2O formation mechanism on Pd(111) and Ti-Pd(111). Palladium is shown in darkcyan, titanium in cyan,nitrogen in blue, and oxygen in red.

L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163 161

transition state with a barrier of 0.92 eV. Herein, the distancebetween C and O atoms decreases from 2.80 Å to 1.88 Å. In the finalstate, adsorbed CO2 is formed which is only weakly bound on Pd(111). For oxygen pre-covered Ti-Pd(111), the adsorption energyof CO on the surface is similar to that on Pd(111). In contrast,the bond distance of the C and O atom decreases to 1.79 Å in thetransition state. Correspondingly, the energy barrier for CO oxida-tion is only 0.74 eV. Clearly, the C–O bond on Ti-Pd(111) is muchstronger as compared to that on Pd(111). As such, not only NOreduction benefits from Ti doping in Pd, but CO oxidation is alsofacilitated by it.

4. Microkinetics simulations

In order to gain a detailed insight into the impact of Ti dopingon the performance of Pd(111) in environmental catalysis, we car-ried out microkinetics simulations to compute reaction rates andproduct distributions. Here, we will discuss the reaction networkin more detail in terms of the composition of the adsorbed layerand the degree of rate control as a function of temperature.

Fig. 9 shows that the N2, N2O and CO2 conversion rates aremuch higher on Ti-Pd(111) than on Pd(111) at relatively low tem-peratures. The reason for this is the much lower barrier for directNO dissociation on Ti-Pd(111). Also, the activation barriers for

N2, N2O and CO2 formation on Ti-Pd(111) are lower than on Pd(111).

The steady-state surface coverages as a function of temperatureare shown in Fig. 10. It can be clearly seen that the Pd(111) is com-pletely covered by NO at low temperature. As NO dissociation pro-ceeds at intermediate temperatures, N2O is produced. At highertemperature, the fraction of free sites increases sharply until thesurface is nearly completely empty. For Ti-Pd(111), NO can alreadydissociate at room temperature, which leads to a high atomic Ncoverage at relatively low temperature (Fig. 10c). With increasingtemperature, N2 is the main product which leads to a decrease ofthe atomic N coverage at intermediate temperature. Furthermore,a small amount of atomic O is present at intermediate tempera-tures. Adsorbed CO can react with atomic O to form CO2, whichcan easily desorb.

The DRC analysis supports these trends. From Fig. 10b, it canbe seen that NO dissociation is the rate-determining step below750 K for Pd(111). With increasing temperature, both N2O as wellas CO2 formation become rate-controlling steps with a similarDRC coefficient (v = 0.5). For Ti-Pd(111), Fig. 10d shows thatNO dissociation controls the overall NO consumption rate at verylow temperatures (T < 400 K). With increasing temperature, N2

formation by atomic N recombination becomes the rate-determining step around 600 K.

Page 10: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

Fig. 10. (a) Surface coverage and (b) degree of rate control for the NO + CO reaction as a function of the temperature on Pd(111); (c) surface coverage and (d) degree of ratecontrol for the NO + CO reaction as a function of the temperature on Ti-Pd(111) (p = 1 atm, NO/CO ratio = 1).

Fig. 9. The product conversion rates ln(r) (r in mol s�1 site�1) on Pd(111) and Ti-Pd(111) (p = 1 atm, NO/CO ratio = 1) for: (a) N2, (b) N2O, (c) CO2.

162 L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163

This result is reflected in Fig. 9a, wherein it can be seen that therate of N2 formation is significantly higher for Ti-Pd(111) as com-pared with Pd(111). At the highest temperature on both surfaces,CO2 formation and N2 formation become rate-controlling steps asthe temperature is sufficiently high that all previous elementaryreaction steps pose no longer a kinetic barrier.

Fig. 11. Selectivity as a function of the temperature for the NO + CO reac

In terms of selectivity, N2 is formed on Ti-Pd(111) at a muchhigher rate than N2O. In other words, the pathway for NO reduc-tion to N2 is kinetically more favourable on Ti-Pd(111). This canbe clearly seen in Fig. 11, wherein the selectivity as a function oftemperature is depicted. At low temperature, N2O is the dominantproduct of NO reduction. However, the main production switches

tion on (a) Pd(111) and (b) Ti-Pd(111) (p = 1 atm, NO/CO ratio = 1).

Page 11: Transition metal doping of Pd(111) for the NO + CO ... · Transition metal doping of Pd(111) for the NO + CO reaction Long Zhang, Ivo A.W. Filot, Ya-Qiong Su, Jin-Xun

L. Zhang et al. / Journal of Catalysis 363 (2018) 154–163 163

to N2 at high temperatures. In Fig. S5, plots of the degree of selec-tivity control (DSC) for N2 and N2O as a function of temperaturesfor the Pd(111) suggest that elementary steps of N2 and N2O for-mation control the product selectivity at high temperature. ForTi-Pd(111), the DSC analysis for N2 and N2O formation indicatesthat the selectivity of N2 is determined by NO dissociation and Natom recombination, while the reaction of N + NO controls theselectivity of N2O at low temperatures (Fig. S6). Thus, the very highselectivity of N2 on the Ti-doped surface can be attributed to thelow energy barrier for NO dissociation and N atom recombination.

In summary, from our microkinetics simulations it is clear thatunder typical reaction conditions, the Ti-Pd(111) surface is muchmore active as well as selective towards N2 formation as comparedto pristine Pd(111). The dominant contributor to this enhancedactivity is the very low NO dissociation barrier on Ti-Pd(111) asis evident from the DRC analysis. Whereas this elementary reactionstep is rate-determining over a relatively large temperature rangefor Pd(111), it is only moderately rate-controlling for Ti-Pd(111).

5. Conclusions

We have investigated the thermodynamics and kinetics of allelementary reaction steps relevant to NO + CO reaction on pristineand Ti-doped Pd(111) using DFT calculations and used these datato predict reaction rate and product distribution in microkineticssimulations. Direct NO dissociation faces high activation barrierson Pd(100) and Pd(111). Ti-doping of Pd(111) results in a strongdecrease of the NO dissociation barrier to 0.65 eV. DOS and COHPanalyses show that Ti-doping changes the electronic structure suchthat N and O atoms binds strong, explaining the lower barrier forNO dissociation. This fact is further reflected by a BEP relationshowing that the transition state is late and stabilization of disso-ciated N and O thus facilitates a decrease of the reaction barrier.The mechanism of N2, N2O and CO2 formation is explored on Pd(111) and Ti-Pd(111). The energy barriers for the N + N and CO+ O association on the Ti-doped surface are also lower than onthe Pd(111) surface. Microkinetics simulations show that NOalready can be dissociated at room temperature on Ti-Pd(111)which leads to much higher overall reaction rates at low tempera-tures. In addition, the rate-determining step for NO + CO reaction isidentified. At low temperatures, the NO dissociation controls theoverall reaction rate on both unpromoted and Ti-doped Pdsurfaces. Importantly, our findings indicate that the activity andselectivity can be drastically improved by non-noble transitionmetal doping and point out a very promising way to design thePd-based three-way catalysts.

Acknowledgements

The authors acknowledge financial support from the EuropeanUnion’s Horizon 2020 research and innovation programme underGrant No. 686086 (Partial-PGMs). Access to supercomputing facil-ities was funded by The Netherlands Organization for ScientificResearch.

Appendix A. Supplementary material

Supplementary data associated with this article can be found, inthe online version, at https://doi.org/10.1016/j.jcat.2018.04.025.

References

[1] M. Iwamoto, H. Hamada, Catal. Today 10 (1991) 57–71.[2] A. Fritz, V. Pitchon, Appl. Catal. B 13 (1997) 1–25.

[3] P.S. Lambrou, P.G. Savva, J.L.G. Fierro, A.M. Efstathiou, Appl. Catal. B 76 (2007)375–385.

[4] K.C. Taylor, Cat. Rev. - Sci. Eng. 35 (1993) 457–481.[5] M. Shelef, G. Graham, Cat. Rev. - Sci. Eng. 36 (1994) 433–457.[6] D. Loffreda, D. Simon, P. Sautet, J. Catal. 213 (2003) 211–225.[7] A. Eichler, J. Hafner, J. Catal. 204 (2001) 118–128.[8] R.M. Wolf, J. Siera, F.C.M.J.M. van Delft, B.E. Nieuwenhuys, Faraday Discuss.

Chem. Soc. 87 (1989) 275–289.[9] Q. Zhang, L. Lv, J. Zhu, X. Wang, J. Wang, M. Shen, Catal, Sci. Technol. 3 (2013)

1069–1077.[10] Z.P. Liu, P. Hu, Top. Catal. 28 (2004) 71–78.[11] Y. Wang, R. Oord, D. van den Berg, B.M. Weckhuysen, M. Makkee,

ChemCatChem 9 (2017) 2935–2938.[12] S. Piccinin, M. Stamatakis, Top. Catal. 60 (2017) 141–151.[13] V. Celorrio, P.M. Quaino, E. Santos, J. Flórez-Montaño, J.J.L. Humphrey, O.

Guillén-Villafuerte, D. Plana, M.J. Lázaro, E. Pastor, D.J. Fermín, ACS Catal. 7(2017) 1673–1680.

[14] Y. Nishihata, J. Mizuki, T. Akao, H. Tanaka, M. Uenishi, M. Kimura, T. Okamoto,N. Hamada, Nature 418 (2002) 164–167.

[15] C.J. Zhang, P. Hu, J. Am. Chem. Soc. 123 (2001) 1166–1172.[16] N. Lopez, J.K. Nørskov, Surf. Sci. 477 (2001) 59–75.[17] Y. Gao, L.M. Zhang, C.C. Kong, Z.M. Yang, Y.M. Chen, Chem. Phys. Lett. 658

(2016) 7–11.[18] Y. Hu, K. Griffiths, P.R. Norton, Surf. Sci. 603 (2009) 1740–1750.[19] S. Roy, M.S. Hegde, G. Madras, Appl. Energy 86 (2009) 2283–2297.[20] C.A. Farberow, J.A. Dumesic, M. Mavrikakis, ACS Catal. 4 (2014) 3307–3319.[21] W.C. Ding, X.K. Gu, H.Y. Su, W.X. Li, J. Phys. Chem. C 118 (2014) 12216–12223.[22] K. Paredis, L.K. Ono, F. Behafarid, Z. Zhang, J.C. Yang, A.I. Frenkel, B.R. Cuenya, J.

Am. Chem. Soc. 133 (2011) 13455–13464.[23] L.Y. Huai, C.Z. He, H. Wang, H. Wen, W.C. Yi, J.Y. Liu, J. Catal. 322 (2015) 73–83.[24] K. Thirunavukkarasu, K. Thirumoorthy, J. Libuda, C.S. Gopinath, J. Phys. Chem.

B 109 (2005) 13272–13282.[25] S. Vesecky, D. Rainer, D. Goodman, J. Vac. Sci. Technol. A 14 (1996) 1457–1463.[26] J. Rempel, J. Greeley, L.B. Hansen, O.H. Nielsen, J.K. Nørskov, M. Mavrikakis, J.

Phys. Chem. C 113 (2009) 20623–20631.[27] T. Ward, L. Delannoy, R. Hahn, S. Kendell, C.J. Pursell, C. Louis, B.D. Chandler,

ACS Catal. 3 (2013) 2644–2653.[28] J.X. Liu, Z. Liu, I.A.W. Filot, Y.Q. Su, I. Tranca, E.J.M. Hensen, Catal, Sci. Technol. 7

(2017) 75–83.[29] H.C. Ham, J.A. Stephens, G.S. Hwang, J. Han, S.W. Nam, T.H. Lim, J. Phys. Chem.

Lett. 3 (2012) 566–570.[30] D. Cheng, H. Xu, A. Fortunelli, J. Catal. 314 (2014) 47–55.[31] K.H. Dostert, C.P. O’Brien, F. Ivars-Barceló, S. Schauermann, H.J. Freund, J. Am.

Chem. Soc. 137 (2015) 13496–13502.[32] T. Dellwig, G. Rupprechter, H. Unterhalt, H.J. Freund, Phys. Rev. Lett. 85 (2000)

776–779.[33] P.E. Blöchl, Phys. Rev. B: Condens. Matter 50 (1994) 17953–17979.[34] G. Kresse, J. Hafner, Phys. Rev. B: Condens. Matter 47 (1993) 558–561.[35] G. Kresse, J. Furthmüller, Phys. Rev. B: Condens. Matter 54 (1996) 11169–

11186.[36] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865–3868.[37] Z. Duan, G. Henkelman, ACS Catal. 4 (2014) 3435–3443.[38] Y. Liu, C. Xu, ChemSusChem 6 (2013) 78–84.[39] J. Greeley, I. Stephens, A. Bondarenko, T.P. Johansson, H.A. Hansen, T. Jaramillo,

J. Rossmeisl, I. Chorkendorff, J.K. Nørskov, Nat. Chem. 1 (2009) 552.[40] D. Chen, P. Sun, H. Liu, J. Yang, J. Mater. Chem. 5 (2017) 4421–4429.[41] B.J. Morgan, G.W. Watson, Phys. Rev. B 80 (2009) 233102.[42] Y.L. Lee, J. Kleis, J. Rossmeisl, D. Morgan, Phys. Rev. B 80 (2009) 224101.[43] G. Henkelman, B.P. Uberuaga, H. Jónsson, J. Chem. Phys. 113 (2000) 9901–

9904.[44] G. Henkelman, H. Jónsson, J. Chem. Phys. 113 (2000) 9978–9985.[45] A.P.J. Jansen, An Introduction to Kinetic Monte Carlo Simulations of Surface

Reactions, Springer, 2012.[46] I.A.W. Filot, R.A. van Santen, E.J.M. Hensen, Angew. Chem. Int. Ed. 53 (2014)

12746–12750.[47] I.A.W. Filot, R.J.P. Broos, J.P.M. van Rijn, G.J.H.A. van Heugten, R.A. van Santen,

E.J.M. Hensen, ACS Catal. 5 (2015) 5453–5467.[48] C.T. Campbell, Top. Catal. 1 (1994) 353–366.[49] C.T. Campbell, J. Catal. 204 (2001) 520–524.[50] C. Stegelmann, A. Andreasen, C.T. Campbell, J. Am. Chem. Soc. 131 (2009)

8077–8082.[51] I.A.W. Filot, B. Zijlstra, E.J.M. Hensen, MKMCXX, a C++ Program for

Constructing Microkinetic Models. http://www.mkmcxx.nl.[52] P.N. Brown, G.D. Byrne, A.C. Hindmarsh, J. SIAM., Sci. Comput. 10 (1989) 1038–

1051.[53] G.D. Byrne, A.C. Hindmarsh, A.C.M. Trans, Math. Softw. 1 (1975) 71–96.[54] G.D. Byrne, A.C. Hindmarsh, J. Comput. Phys. 70 (1987) 1–62.[55] D. Loffreda, D. Simon, P. Sautet, Chem. Phys. Lett. 291 (1998) 15–23.[56] X. Wei, X.F. Yang, A.Q. Wang, L. Li, X.Y. Liu, T. Zhang, C.Y. Mou, J. Li, J. Phys.

Chem. C 116 (2012) 6222–6232.[57] M. Wuttig, D. Lüsebrink, D. Wamwangi, W. Wełnic, M. Gilleßen, R.

Dronskowski, Nat. Mater. 6 (2007) 122–128.[58] W.V. Glassey, G.A. Papoian, R. Hoffmann, J. Chem. Phys. 111 (1999) 893–910.[59] R.A. van Santen, M. Neurock, S.G. Shetty, Chem. Rev. 110 (2010) 2005–2048.