Elsevier

Journal of Catalysis

Volume 300, April 2013, Pages 20-29
Journal of Catalysis

Structure–activity relationships in NH3-SCR over Cu-SSZ-13 as probed by reaction kinetics and EPR studies

https://doi.org/10.1016/j.jcat.2012.12.020Get rights and content

Abstract

Cu-SSZ-13 catalysts with various Cu loadings were prepared via aqueous solution ion-exchange. The hydrated samples were characterized with Electron Paramagnetic Resonance (EPR). Cu2+ ion coordination numbers were obtained by analyzing the hyperfine structures, while Cu–Cu distances were estimated from line broadening of the EPR features. By examining EPR and temperature-programmed reduction (TPR) results, two Cu2+ ion locations are suggested. Standard NH3-SCR, as well as non-selective NH3 oxidation reaction with O2, were carried out over these catalysts at high-space velocities. For the SCR reaction, intra-particle diffusion limitations are found. The kinetic data allow for reactant diffusivities to be estimated. However, clear structure–activity relationships for the SCR reaction cannot be derived due to this diffusion limitation. The slower NH3 oxidation reaction, on the other hand, is kinetically limited at low temperatures, and, therefore allows for a correlation between Cu2+ ion location and reaction kinetics to be made.

Graphical abstract

This study confirms two isolated Cu2+ ion locations for the Cu-SSZ-13 catalyst by TPR and EPR studies. Intra-particle diffusion limitation was identified for the NH3-SCR reaction via detailed analysis of the reaction kinetics data.

  1. Download : Download high-res image (93KB)
  2. Download : Download full-size image

Highlights

► This study further confirms the significance of isolated Cu ions in catalyzing NH3-SCR. ► Two Cu2+ ion monomer locations were suggested. ► Intra-particle diffusion limitation was realized in NH3-SCR. ► NH3 oxidation activity increases with increasing Cu loading.

Introduction

The abatement of environmentally harmful NOx compounds (NO, NO2, and N2O) emitted from mobile or stationary power sources remains a challenging task for the catalysis community. In particular, conventional three-way catalysts used in the exhaust after-treatment technologies of internal combustion engines prove ineffective when the engine is operated under highly oxidizing conditions. However, in order to achieve better fuel efficiency and decrease CO2 emission, operation under such fuel-lean conditions is a prerequisite for diesel engines and new-generation gasoline engines. The problem is daunting, since reduction chemistry (NOx to N2) has to be carried out under highly oxidizing conditions. Two elegant after-treatment techniques have been developed for reduction in NOx from lean burn engines; namely lean NOx traps (LNT) and selective catalytic reduction (SCR) with hydrocarbons (HC-SCR) or with ammonia (NH3-SCR). For the NH3-SCR technology, transition metal (in particular Fe and Cu) ion-exchanged zeolite catalysts have shown high activity and N2 selectivity.

The most extensive studies have been carried out on Cu2+ ion-exchanged ZSM-5 (Cu-ZSM-5) zeolites, first shown to exhibit high NO decomposition rates and NOx SCR activities in the 1980s [1], [2], [3], [4], [5], [6], [7]. Later, Cu2+ ion-exchanged beta zeolite (Cu-beta) was shown to have excellent activity in the SCR of NOx with NH3, and metal-exchanged beta zeolites are generally found to have greater hydrothermal stability than similar ZSM-5 catalysts [8]. Yet, none of these Cu-zeolite catalysts show durability sufficient for the automotive industry. In the year 2010, Cu2+-exchanged molecular sieves with Chabazite (CHA) structures, for example, Cu-SSZ-13 and Cu-SAPO-34, have been commercialized as NOx after-treatment catalysts in diesel-powered engines for transportation, due apparently to their much improved activity, selectivity, and durability [9], [10], [11], [12], [13], [14], [15], [16], [17], [18].

We and other researchers compared SCR activities and hydrothermal stabilities of Cu-SSZ-13 with other Cu2+ ion-exchanged zeolites (especially Cu-ZSM-5 and Cu-beta) [10], [13], [15]. Indeed, Cu-SSZ-13 shows comparable or even higher activity, higher N2 selectivity and most impressively, much higher hydrothermal stability even under harsh treatment conditions; for example, extended hydrothermal treatments at 1073 K. There appears to be general consensus that these unique properties of Cu-SSZ-13 are due to its small-pore structure [10], [13], [15]. SSZ-13 adopts the R3¯m space group with a hexagonal unit cell described via the following face symbol: [46·62] + [412·62·86]. The largest pore has an opening of 3.8 Å (8-membered ring). In contrast, ZSM-5 has medium size pore openings (∼5.5 Å, 10-membered ring), while beta has the largest pores (∼7 Å, 12-membered ring) [9]. This structural difference is significant in improving the stability of SSZ-13: (1) for zeolite materials, dealumination is the major cause of structural damage and activity loss. The dealumination product, Al(OH)3, has a kinetic diameter of ∼5.03 Å [15]. Thus, it is possible that aluminum cannot escape from the SSZ-13 pores during hydrothermal treatments; even if a dealumination reaction occurs it can insert back to the framework during cooling (i.e., reversible dealumination) to maintain the integrity of the zeolite structure [15]. (2) Medium- and large-pore zeolites adsorb large quantities of hydrocarbons at low temperatures. This is inevitable in vehicle emission treatments, since significant quantities of hydrocarbons are generated during cold-start. As the temperature rises, however, reaction heat generated via burning of these molecules can thermally destroy zeolites. Again, the small-pore SSZ-13 greatly eliminates this deactivation mechanism since hydrocarbons with kinetic diameters larger than methane (3.8 Å) have at most very limited diffusion into the pores [15].

Apart from the commercial success, detailed catalyst structures, reaction mechanisms, and structure–activity relationships are still lacking. For example, there are debates even regarding the location of Cu2+, the catalytically active centers, within the SSZ-13 framework. Lobo and coworkers applied Rietveld refinement for TR-XRD data of their dehydrated Cu-SSZ-13 catalyst and proposed that Cu2+ ions are located only in six-membered ring windows with a coordination number of 3 and average Cu–O distances of 2.2–2.3 Å [11]. In a more recent study, they used EXAFS to analyze Cu-SSZ-13 and concluded that, for a hydrated sample, four lattice oxygen atoms coordinate to the Cu2+ ion with an average Cu–O distance of 2.02 Å, whereas for a dehydrated sample, the coordination number reduces to 3 at an average distance of 1.93 Å [16], [19]. We, on the other hand, have recently concluded that the copper locations within the SSZ-13 framework are affected by two factors: copper loading and moisture [14]. Both temperature-programmed reduction and FTIR coupled with CO and NO titrations reveal that there are two different Cu moieties in Cu-SSZ-13 as a function of Cu loading. For dehydrated samples and at low Cu loadings (at an ion-exchange (IE) level of ∼20%), only a single type of Cu2+ species exists which is rather hard to reduce to Cu+ (at ∼653 K). However, as the Cu loading increases, another Cu species develops, and the reduction temperatures are much lower (∼503 K). Furthermore, under wet-reducing conditions, the reduction temperatures for both species are significantly lower and are found at ∼523 K and ∼483 K, respectively. These findings are explained in the following way: at low Cu loading and under dehydrated conditions, Cu2+ ions are located close to the face of the 6-membered rings via formation of Cu–O bonds with lattice oxygen as suggested by Lobo et al. [11], [16], [19]. When exposed to gases that strongly interact with Cu2+ ions (e.g., NH3), the Cu2+ ions are pulled slightly into the large cages [19]. However, as the Cu loadings increase, some Cu2+ ions also populate inside the large CHA cages next to 8-membered rings. Cu2+ ions next to 8-membered rings have easier access to hydrogen (as they are closer to the pore openings) perhaps explaining, in part, their lower reduction temperatures. Moreover, under wet-reducing conditions, the interaction between water and Cu2+ ions weakens interactions between Cu2+ ions and lattice oxygen, allowing both types of Cu2+ ions to move toward the large CHA cages and become more readily reducible. Note that this water-induced ion mobility is well known within the CHA cages [17], [20].

We note that understanding the Cu locations within SSZ-13 under wet conditions is more relevant to practical NH3-SCR since engine exhausts typically contain ∼10% moisture. In the current study, we use Electron Paramagnetic Resonance (EPR), a technique ideally suited for studying Cu2+ exchanged zeolites, to study hydrated Cu-SSZ-13 catalysts at various Cu loadings in order to gain further insights into their locations. Also, NH3-SCR and NH3 oxidation kinetics are investigated over these catalysts at high-space velocity conditions for the development of structure–activity relationships.

Section snippets

Experimental

SSZ-13 was synthesized using a procedure first developed by Zones [21] and recently slightly modified by others [9], [10], [11], [12], [14], [15], [16], [17]. The synthesis method is detailed in Ref. [11]. Briefly, following Na-SSZ-13 formation, an ion-exchange step was applied using excess amount of 0.1 M NH4NO3 at 353 K for 8 h to generate NH4-SSZ-13. ICP analysis of the calcined sample gave a Si/Al ratio of 6. For a hexagonal unit cell containing 36 tetrahedral (T) atoms and 72 lattice oxygen

Results

Fig. 1 presents EPR results of the Cu-SSZ-13 samples at room temperature. These samples were cooled to room temperature in air after calcination during which they adsorbed moisture and, therefore, are defined as hydrated samples. In these samples, Cu species should all be present as EPR active Cu2+ ions. Monotonic signal intensity increases with increasing Cu loading further confirm this. Rather unexpectedly, two (instead of one) features are found at high field at 3334 and 3407 G. The much

Nature of Cu species in Cu-SSZ-13

To understand structure–activity relationships in Cu-SSZ-13 catalyzed SCR reactions, the nature of Cu-containing species within the SSZ-13 framework is first discussed. Three types of Cu species are common for Cu-zeolites: charge-balancing extra-framework Cu2+ monomers, charge-balancing [Cu–O–Cu]2+ dimers and CuOx clusters [28]. Fickel and Lobo applied Rietveld refinement for TR-XRD data of their dehydrated Cu-SSZ-13 catalyst at an ion-exchange level of 70% (4.39 wt.% Cu, Si/Al = 6 and Cu/Al = 0.35)

Conclusions

  • (1)

    Cu-SSZ-13 powder catalysts with various ion-exchange levels were characterized with Electron Paramagnetic Resonance (EPR) and temperature-programmed reduction (TPR). At low ion-exchange levels (e.g. IE 23%), Cu2+ ions are far apart suggesting one Cu2+ ion within one hexagonal unit cell. In this case, hydrated Cu2+ ions are octahedrally coordinated at sub-ambient temperatures suggesting these are located within the large CHA cages yet still coordinated to lattice oxygen atoms of the 6-membered

Acknowledgments

The authors gratefully acknowledge the US Department of Energy (DOE), Office of Energy Efficiency and Renewable Energy/Vehicle Technologies Program for the support of this work. The research described in this paper was performed at the Environmental Molecular Sciences Laboratory (EMSL), a national scientific user facility sponsored by the DOE’s Office of Biological and Environmental Research and located at Pacific Northwest National Laboratory (PNNL). PNNL is operated for the US DOE by Battelle

References (44)

  • J.Y. Yan et al.

    J. Catal.

    (1996)
  • A. Grossale et al.

    J. Catal.

    (2008)
  • K. Rahkamaa-Tolonen et al.

    Catal. Today

    (2005)
  • T. Komatsu et al.

    J. Catal.

    (1994)
  • J.H. Kwak et al.

    J. Catal.

    (2010)
  • J.H. Kwak et al.

    J. Catal.

    (2012)
  • S.J. Schmieg et al.

    Catal. Today

    (2012)
  • D.W. Fickel et al.

    Appl. Catal. B: Environ.

    (2011)
  • J.-S. McEwen et al.

    Catal. Today

    (2012)
  • M. Calligaris et al.

    Zeolites

    (1982)
  • A.V. Kucherov et al.

    Zeolites

    (1985)
  • A. Sultana et al.

    Catal. Today

    (2011)
  • C. Torre-Abreu et al.

    Appl. Catal. B: Environ.

    (1997)
  • P.S. Metkar et al.

    Appl. Catal. B: Environ.

    (2011)
  • G. Centi et al.

    Appl. Catal. A: Gen.

    (1995)
  • A. Shichi et al.

    Appl. Catal. B: Environ.

    (2000)
  • P.S. Metkar et al.

    Chem. Eng. Sci.

    (2011)
  • E.Y. Choi et al.

    J. Catal.

    (1996)
  • J. Eng et al.

    J. Catal.

    (1997)
  • H.Y. Huang et al.

    Appl. Catal. A: Gen.

    (2002)
  • K. Kamasamudram et al.

    Catal. Today

    (2010)
  • D.M. Ruthven et al.

    Stud. Surf. Sci. Catal.

    (2001)
  • Cited by (434)

    View all citing articles on Scopus
    1

    Visiting graduate student from University of Washington.

    View full text