throbber
Journal of Catalysis 240 (2006) 47–57
`
`www.elsevier.com/locate/jcat
`
`Hydrothermal stability of CuZSM5 catalyst in reducing NO by NH3
`for the urea selective catalytic reduction process
`Joo-Hyoung Park a, Hye Jun Park a, Joon Hyun Baik a, In-Sik Nam a,∗
`Jong-Hwan Lee c, Byong K. Cho c, Se H. Oh c
`
`, Chae-Ho Shin b,
`
`a Department of Chemical Engineering/School of Environmental Science and Engineering, Pohang University of Science and Technology (POSTECH), Pohang,
`Korea
`b Department of Chemical Engineering, Chungbuk National University, Chungbuk, Korea
`c General Motors R&D Planning Center, Warren, MI, USA
`Received 6 September 2005; revised 30 December 2005; accepted 6 March 2006
`
`Available online 5 April 2006
`
`Abstract
`
`To confirm the hydrothermal stability of CuZSM5 for urea selective catalytic reduction (SCR), NO removal activity over a series of the catalyst
`containing various amounts of copper, ranging from 1 to 5 wt%, was examined before and after hydrothermal treatment under a simulated feed
`◦
`C. The degree of catalyst aging varies with respect to the copper content of
`gas stream containing 10% water at a temperature range of 600–800
`the catalyst and the aging temperature. The optimal copper content seems to be about 4 wt% (nearly 125% in terms of ion-exchange level) from
`the standpoint of catalyst aging. The catalysts before and after aging were characterized by XRD, 27Al-MAS-NMR, BET, XAFS, and ESR to gain
`insight into the sintering mechanism of CuZSM5 for the urea SCR process. A slight alteration of the catalyst structure was observed by XRD and
`BET analysis on catalyst aging. The copper ions on the surface of CuZSM5 catalysts on aging remained as either isolated Cu2+
`and/or an oxide
`form of copper, as confirmed by EXAFS. However, ESR spectra of the catalysts clearly revealed four distinct local structures of Cu2+
`species
`in the framework of ZSM5 in the form of a square pyramidal site, a square planar site, and two unresolved distorted sites. Migration of Cu2+
`ions from the square pyramidal and/or the square planar sites of the isolated cupric ions to the two unresolved sites occurs during hydrothermal
`treatment. This causes an alteration of the population of isolated cupric ions on the square pyramidal and/or planar sites and is responsible for the
`hydrothermal stability of the CuZSM5 catalyst in the urea SCR process.
`© 2006 Elsevier Inc. All rights reserved.
`
`Keywords: Urea SCR; Deactivation; Hydrothermal stability; ESR; XAFS; CuZSM5
`
`1. Introduction
`
`The selective catalytic reduction (SCR) of NOx with urea is
`a recognized well-developed technology to remove NOx from
`light- and heavy-duty diesel engines [1,2]. For the actual diesel
`engine exhaust system, the H2O content generated from the
`combustion of diesel fuel with a high carbon number is dis-
`tinctive in the gas stream, and the development of a hot spot
`in the catalytic converter from the sudden burning of locally
`collected particulate can be expected. Consequently, achieving
`
`* Corresponding author. Fax: +82 54 279 8299.
`E-mail address: isnam@postech.ac.kr (I.-S. Nam).
`
`0021-9517/$ – see front matter © 2006 Elsevier Inc. All rights reserved.
`doi:10.1016/j.jcat.2006.03.001
`
`hydrothermal stability of the catalyst is a critical issue in the
`commercial application of urea SCR technology to the exhaust
`stream from diesel engines [3].
`CuZSM5 is one of the most promising catalysts for the SCR
`◦
`C,
`of NOx by urea at an exhaust temperature of around 150
`particularly for light-duty diesel engines [3,4]. For the decom-
`position of urea, 1 mol of urea is thermally decomposed and
`then easily hydrolyzed on the catalyst surface to produce 2 mol
`of ammonia and 1 mol of carbon dioxide, leading to the practi-
`cally equivalent overall reaction to NH3 SCR of NO [3–7].
`Numerous studies have been conducted regarding the na-
`ture of Cu on the surface of zeolite, which is generally recog-
`nized as an active reaction site for the present reaction system,
`
`Exhibit 2024.001
`
`

`
`48
`
`J.-H. Park et al. / Journal of Catalysis 240 (2006) 47–57
`
`and its reaction mechanism through redox chemistry [8–14].
`It has been commonly accepted that isolated Cu2+
`and Cu–
`O–Cu dimer species on the catalyst surface play key roles in
`the NH3 SCR reaction. Mizumoto et al. [12] prepared Cu-Y
`zeolite containing mainly isolated Cu2+
`to suggest a Langmuir–
`Hinshelwood-type reaction mechanism by the reaction between
`strongly adsorbed NH3 and weakly adsorbed NO on isolated
`copper ions over the catalyst surface. Choi et al. also confirmed
`a similar mechanism by temperature-programmed desorption
`experiments of NO and NH3 over Cu-exchanged Mordenite-
`type zeolite catalyst [13]. Komatsu et al. [14] studied NH3 SCR
`for a series of Cu zeolites with varying copper exchange lev-
`els and Si/Al ratios and concluded that the active reaction site
`of the CuZSM5 catalyst for NH3 SCR is copper dimer formed
`on the catalyst surface. The active reaction site on the surface
`of Cu-exchanged zeolite has been also investigated for the de-
`composition of NO and HC (hydrocarbon) SCR by lumines-
`cence, Fourier transform infrared spectroscopy, electron spin
`resonance (ESR), and X-ray absorption fine structure (XAFS)
`studies [15–23]. Iwamoto et al. proposed oxygen-bridged bin-
`uclear [Cu–O–Cu]2+
`in CuZSM5 for NO decomposition [15],
`and Sárkány et al. reported experimental evidence indicating
`that the complex forms all at once [16]. Ddeˇcek et al. examined
`the four distinctive sites of Cu ion coordination over various
`zeolite structures, including MFI, MOR, FER, BEA, and FAU,
`+
`emission
`using a multispectroscopic approach including Cu
`spectra, in situ infrared spectroscopy for Cu2+
`-adsorbing NO,
`and ESR of Cu2+
`[17].
`But little literature exists on the inactivation of the active re-
`action sites for NH3 SCR, although the hydrothermal stability
`of zeolite has long been considered a necessary issue to resolve
`for the commercial application of zeolite-type catalyst to auto-
`motive engines from the standpoint of alteration of the active
`reaction sites, including isolated Cu2+
`, and Cu2+
`dimer during
`the reaction. It has been commonly observed that the deacti-
`vation of SCR activity over CuZSM5 catalyst is due mainly
`to the degradation of zeolite support and/or the formation of
`Cu-aluminate by dealumination, the decrease in active reaction
`sites by the transformation of Cu2+
`to CuO, the redistribution
`of the reaction sites through the migration of Cu2+
`, or a com-
`bination of these mechanisms [19,20,24–28]. Furthermore, the
`optimal copper content on the catalyst surface has been little
`examined from the standpoint of catalyst aging for commercial
`application.
`The purpose of the present study was to optimize the copper
`content on the surface of ZSM5 catalyst from the standpoint
`of catalyst aging by evaluating the hydrothermal stability of a
`series of CuZSM5 catalysts. To accomplish this goal, various
`complementary spectroscopic techniques were used and their
`results compared. In particular, XAFS and ESR studies were
`carried out to investigate the nature of copper ions on the sur-
`face of CuZSM5 catalyst, such as the oxidation states of copper
`and the coordination structure of Cu2+
`in the framework of
`zeolite on the catalyst sintering with respect to aging temper-
`ature.
`
`Table 1
`Physicochemical properties of the catalysts employed in the present study
`
`Cu
`content
`(wt%)
`1.95
`
`Ion exchange
`level
`(%)
`61
`
`2.93
`
`92
`
`3.87
`
`124
`
`4.73
`
`150
`
`BET
`(m2/g)
`
`400
`
`372
`363
`
`387
`
`357
`321
`
`383
`355
`324
`
`357
`337
`294
`191
`
`Micropore
`surface area
`(m2/g)
`377
`
`331
`321
`
`362
`
`319
`303
`
`354
`328
`285
`
`323
`301
`260
`182
`
`Sample
`
`CuZSM5-61-fresh
`CuZSM5-61-600
`CuZSM5-61-700
`CuZSM5-61-800
`
`CuZSM5-92-fresh
`CuZSM5-92-600
`CuZSM5-92-700
`CuZSM5-92-800
`
`CuZSM5-124-fresh
`CuZSM5-124-600
`CuZSM5-124-700
`CuZSM5-124-800
`
`CuZSM5-150-fresh
`CuZSM5-150-600
`CuZSM5-150-700
`CuZSM5-150-800
`
`2. Experimental
`
`2.1. Catalyst preparation
`type ZSM5 (Si/Al ratio = 14), obtained from Tosoh
`+
`NH4
`Corp., was used as a parent catalyst for preparing the series of
`catalysts in the present study. The ZSM5 was exchanged with a
`0.01 M (CH3CO2)2Cu·H2O (Aldrich, 98%) solution to obtain
`a Cu-based ZSM5 at room temperature for 5 h [3,4]. The room
`temperature for exchanging copper ions into ZSM5 catalyst was
`used to avoid the formation of copper oxides on the catalyst
`surface during the course of the ion exchange [4]. This was fol-
`◦
`◦
`C for 12 h and calcining at 500
`C in
`lowed by drying at 110
`air for 5 h. The copper content was controlled by repeating the
`ion-exchange procedure.
`To investigate the hydrothermal stability of CuZSM5, the
`catalyst samples were sintered under a simulated feed gas
`stream containing 10% H2O in air balance with a flow rate
`◦
`C for 24 h. Cu-based
`of 500 cc/min at 600, 700, and 800
`ZSM5 catalysts containing a various copper loadings with re-
`spect to the aging temperatures were obtained and designated as
`CuZSM5-x-y, where x and y represent the percentage degree
`of Cu2+
`ion exchange, and the aging temperature, respectively.
`Table 1 lists the physicochemical properties of the catalysts pre-
`pared in this study.
`
`2.2. Reactor system and experimental procedure
`
`For urea SCR technology, urea is thermally decomposed into
`one mol of ammonia and one mol of isocyanic acid initially,
`and the isocyanic acid formed by the thermal decomposition
`of urea can readily undergo hydrolysis on the catalyst surface
`to produce another mol of ammonia [7]. Thus, the complete
`decomposition of 1 mol of urea produces 2 mol of ammonia
`and 1 mol of carbon dioxide [1,5–7],
`H2N–CO–NH2 + H2O → 2NH3 + CO2.
`
`(1)
`
`Exhibit 2024.002
`
`

`
`J.-H. Park et al. / Journal of Catalysis 240 (2006) 47–57
`
`49
`
`N2-filled ionization chambers. CuO, Cu2O, and Cu metal were
`used as reference compounds.
`The data analyses for EXAFS were carried out by a standard
`procedure. The inherent background in the data was removed
`by fitting a polynomial to the edge region and then extrapolat-
`ing through the entire spectrum from which it was subtracted.
`The resulting spectra, μ(E), were normalized to an edge jump
`of unity for comparing the X-ray absorption near-edge struc-
`tures (XANES) directly. The EXAFS function, χ (E), was ob-
`tained from χ (E) = {μ(E) − μ0(E)}/μ0(E) [29]. The result-
`ing EXAFS spectra were k2-weighted to compensate for the
`attenuation of EXAFS amplitude at high k and then Fourier
`−1 (cid:2) k (cid:2) 11.5 Å
`−1 with
`transformed in the range of 2.0 Å
`−1. To determine the
`a Kaiser–Bessel function of dk = 1 Å
`structural parameters, nonlinear least squares curve fitting was
`performed in the range of R (cid:2) ∼2 Å corresponding to the dis-
`tance to the Cu–O in the central-atom phase-corrected Fourier-
`transformed (FT) spectra, using the UWXAFS and IFEFFIT
`packages according to the following EXAFS formula [30]:
`(cid:4)
`(cid:4)
`χ (k) = −S2
`Ni
`Fi (k) exp
`kR2
`(cid:4)
`i
`2kRi + φi (k)
`× sin
`where the backscattering amplitude, Fi (k), the total phase shift,
`φi (k), and the photoelectron mean free path, λ(k), were theo-
`retically calculated for all scattering paths, including multiple
`ones by a curved wave ab initio EXAFS code FEFF8 [31].
`The ESR spectra were obtained at room temperature in the
`X-band using a JEOL model JES-TE300 spectrometer. The
`ESR signals of Cu2+
`were recorded in the field region from
`2000 to 4000 G with a sweep time of 10 min. A coaxial quartz
`cell was placed in the ESR cavity and connected with stainless
`steel capillaries of the flow system by Teflon ferrules. The sam-
`◦
`C in pure O2, evacuated, and sealed off
`ples were heated at 500
`in situ. To provide the maximum accuracy of the ESR signals,
`the packing height of the ampoule was constant in all cases,
`with the center of the sample positioned in the middle of the
`ESR cavity. The ESR signal from DPPH (g(cid:6) = 2.0036) was
`used as an internal standard. The Origin and Excel programs for
`Windows were used for the baseline correction, analysis of the
`fourfold hyperfine lines, and double-integration of the recorded
`ESR spectra.
`
`,
`
`(2)
`
`(cid:3)−2σ 2
`i k2
`
`(cid:3)−2Ri /λ(k)
`
`exp
`
`(cid:2) i
`
`(cid:3)
`
`o
`
`3. Results and discussion
`
`3.1. Effect of aging on NO removal activity
`
`The NO removal activity for a series of CuZSM5 catalysts
`◦
`C, where catalyst de-
`before and after aging, particularly at 700
`activation is specifically evident, is given in Fig. 1. The figure
`shows a bell-shaped activity profile with respect to the reac-
`tion temperature, typical for NH3 SCR catalyst maintaining a
`◦
`wide operating temperature window (200–400
`C) containing
`NO conversion >90%. In a reaction temperature range higher
`◦
`C, NO conversion gradually decreases mainly due to
`than 400
`
`This leads to the practically equivalent overall reaction of NO to
`NH3 SCR. The identical deNOx performance of the CuZSM5
`catalyst by urea SCR to that by NH3 SCR was observed under
`the specific reaction conditions [3]. Therefore, an NH3 SCR
`test to confirm the hydrothermal stability of the series of the
`CuZSM5 catalysts for urea SCR was performed for experimen-
`tal convenience.
`The activity of the CuZSM5 catalysts for NO reduction by
`NH3 was examined in a fixed-bed flow reactor, typically con-
`taining ca. 1 g of the catalyst was sieved to a mesh size of 20/30
`to minimize the mass transfer limitations of the catalyst. Be-
`fore the evaluation of activity, the catalyst was pretreated in situ
`with a total flow of 3300 cc/min containing 79% N2 and 21%
`◦
`C, and then cooled to room temperature. A reaction
`O2 at 500
`gas mixture consisting of 500 ppm NO, 500 ppm NH3, 5% O2,
`and 10% H2O in N2 balance was fed into the reactor system
`through Brooks mass flow controllers (model 5850E). A total
`flow rate of 3300 cc/min (corresponding to GHSV = 100,000
`−1) was mainly used for the catalyst activity testing. The NO
`h
`concentration was analyzed by an on-line chemiluminescence
`NO–NOx analyzer (Thermo Electron, model 42H). The details
`of the reactor system have been given elsewhere [3,7].
`
`2.3. Catalyst characterization
`
`BET surface areas were measured by Micromeritics ASAP
`2010 sorption analyzer with a static volumetric technique,
`based on the amount of N2 adsorbed at liquid N2 temperature.
`◦
`C in vacuum for 5 h before
`The samples were degassed at 200
`the adsorption measurements. BET surface area and micropore
`volume were calculated by the t-plot method. The micropore
`surface area was obtained by subtracting the external surface
`area from the total surface area.
`Powder XRD patterns for the catalysts were observed by us-
`ing an M18XHF X-ray diffractometer with Ni-filtered Cu-Kα
`radiation (λ ∼ 1.54184 Å). Data were collected in angles rang-
`◦
`◦
`◦
`to 90
`with a step size of 0.02
`and a step time of
`ing from 5
`10 s under continuous sample rotation during the scan.
`The 27Al magic-angle spinning nuclear magnetic resonance
`(MAS-NMR) spectra were obtained on a Bruker AVANCE 500
`spectrometer at an 27Al frequency of 130.325 MHz in 4-mm
`rotors at a spinning rate of 10.0 kHz. The spectra were obtained
`with the acquisition of ca. 10,000 pulse transients, which were
`reported with a π/4 rad pulse length of 5.00 µs and a recycle
`delay of 1.0 s.
`Extended X-ray absorption fine structure (EXAFS) spectro-
`scopic measurements were performed with the synchrotron ra-
`diation by using the EXAFS facility installed at beamline 3C1
`in the Pohang Accelerator Laboratory. The ring was operated
`at 2.5 GeV with 200 mA electron current and 1% coupling.
`The spectra were measured with an Si(111) channel-cut mono-
`chromator with an energy resolution of E/E = 2× 10
`−4 that
`remained constant at the Cu K-edge (8979 eV).
`The samples spread uniformly between adhesive tapes to ob-
`tain an optimal absorption jump. All of the data were recorded
`in a transmission mode at room temperature, and the intensities
`of the incident and transmitted beams were measured in 100%
`
`Exhibit 2024.003
`
`

`
`50
`
`J.-H. Park et al. / Journal of Catalysis 240 (2006) 47–57
`
`ation of the chemical state of copper on the catalyst surface
`[19,20,24–28].
`
`3.2. Confirming the structure of the catalyst support, ZSM5,
`by 27Al MAS-NMR and powder XRD
`
`27Al MAS-NMR spectra were examined for the CuZSM5-
`150 catalyst before and after hydrothermal treatment with the
`simulated aging gas stream containing 10% water, as shown in
`Fig. 4. No shoulder due to extra-framework AlO4 units was de-
`tected even for CuZSM5-150-800, except for that assigned to
`tetrahedral Al species. It may be direct evidence for the main-
`tenance of the catalyst structure without dealumination from
`the framework of zeolite due to hydrothermal treatment even
`◦
`at 800
`C. Identical NMR results were observed for all of the
`catalysts before and after aging. However, the intensity of the
`Al tetrahedral peak decreased, as also shown in Fig. 4, proba-
`bly due to the paramagnetic field of the Cu ions at their ion-
`exchange positions in the framework of zeolite structure. This
`characteristic of the Cu ions also produces an intensified and ex-
`tended side band of the NMR spectra, which may be originated
`from the anisotropic dipolar interaction between the electron
`spin of the Cu ions and the nuclei spin of the Al in the zeolite
`framework [33].
`Fig. 5 shows the typical powder XRD patterns for the series
`of the catalysts prepared before and after hydrothermal treat-
`ment. The relative intensities of X-ray peaks decrease due to
`the amount of Cu loadings on the catalyst surface and the aging
`temperature, particularly for underexchanged CuZSM5 cata-
`lysts. In addition, a fairly obvious weaker crystallinity for the
`overexchanged CuZSM5 compared with the underexchanged
`one is seen. It closely agrees with the decreasing trends of BET
`and micropore surface areas on the hydrothermal treatment of
`the catalysts as listed in Table 1. The irreversible decrease of the
`catalyst surface areas mainly causes a permanent loss of NO re-
`moval activity for overexchanged CuZSM5 catalysts aged at the
`high sintering temperature. It may be also attributed to the si-
`multaneous effects of the combination of high Cu loading and
`harsh aging conditions on the framework structure of ZSM5-
`type zeolite catalyst [34]. In addition, a broad XRD peak can be
`◦
`◦
`observed at 2θ of ca. 35.5
`and 39
`in the patterns of CuZSM5-
`150 catalysts, indicating formation of CuO on the catalyst sur-
`face, even for the fresh counterpart catalyst, but basically no
`CuO peak for the slightly underexchanged CuZSM5 catalyst,
`regardless of the aging temperatures. This may cause the de-
`crease of NO removal activity of overexchanged CuZSM5 cat-
`◦
`alyst aged at 800
`C, as mentioned earlier, but catalyst deactiva-
`tion for underexchanged CuZSM5 catalysts can barely be elu-
`cidated. Note that the additional XRD peaks for the formation
`of CuO and Cu2O on the catalyst surface of underexchanged
`CuZSM5 also cannot be identified, as shown in Fig. 5. Also
`note that formation of the Cu-aluminate related to the octahe-
`dral Al of the catalyst was hardly observed on the basis of the
`27Al MAS-NMR and XRD results, although we did not con-
`duct a TPR analysis to confirm its formation, as was done by
`Yan et al. [28].
`
`◦
`C
`Fig. 1. Activity of a series of CuZSM5 catalysts before and after aging at 700
`[(F) CuZSM5-61-fresh, (2) CuZSM5-92-fresh, (a) CuZSM5-124-fresh,
`(") CuZSM5-150-fresh,
`(E) CuZSM5-61-700,
`(1) CuZSM5-92-700,
`(e) CuZSM5-124-700, (!) CuZSM5-150-700] for the reduction of NO by
`NH3. Feed gas composition is 500 ppm NO, 500 ppm NH3, 5% O2, 10% H2O
`in N2 balance.
`
`(3)
`
`the oxidation of NH3 at the high reaction temperature with oxy-
`gen leading to a significant decrease in NH3 selectivity for NOx
`reduction as follows [32]:
`4NH3 + 3O2 → 2N2 + 6H2O.
`Among the catalysts before and after aging, CuZSM5-124 ex-
`hibited the best performance for NO removal activity main-
`◦
`C, which
`tenance, even at a reaction temperature below 200
`is critical for the application of the present urea SCR tech-
`nology to light-duty diesel engines. However, the activity for
`◦
`C decreases
`underexchanged CuZSM5-61 catalyst aged at 700
`significantly.
`Figs. 2 and 3 illustrate the aging effect of a series of
`CuZSM5 catalysts with respect to the copper content and
`the aging temperatures on NO removal activity in a reac-
`◦
`tion temperature range of 150–450
`C. For the catalysts con-
`taining <4% copper loading, NO removal activity increases
`as the Cu loading increases, regardless of the catalyst ag-
`ing. However, for a highly overexchanged catalyst such as
`CuZSM5-150, the activity decrease due to the catalyst sinter-
`ing is somewhat milder than that for slightly overexchanged
`catalyst, CuZSM5-124. For an underexchanged catalyst con-
`taining low copper content, such as CuZSM5-65, the worst
`catalyst deactivation can be observed, even for the catalyst
`◦
`aged at 600
`C. However, the cause of the sintering of Cu-
`based zeolite catalyst for SCR reaction has not yet been sys-
`tematically examined. It has been simply speculated, probably
`due to the structural destruction of thezeolite, and the alter-
`
`Exhibit 2024.004
`
`

`
`J.-H. Park et al. / Journal of Catalysis 240 (2006) 47–57
`
`51
`
`◦
`◦
`◦
`Fig. 2. Effect of Cu loading on the conversion of NO for CuZSM5 catalysts: (A) fresh, (B) aged at 600
`C, (C) aged at 700
`C, and (D) aged at 800
`C. Feed gas
`−1. [Reaction temperature: (") 150, (a) 175, (2) 200, (F) 300,
`composition is 500 ppm NO, 500 ppm NH3, 5% O2, 10% H2O in N2 balance, and SV is 100,000 h
`◦
`and (Q) 400
`C.]
`
`3.3. Local structure of copper ions by XAFS
`
`The shape and position of the Cu K-edge XANES provide
`information on the electronic structure and the local coordina-
`tion geometry of the absorbing Cu atom [22–24,35–42]. Fig. 6
`shows the XANES spectra for CuZSM5 catalysts, which are
`quite typical for the existence of the distorted octahedral Cu2+
`on the catalyst surface. The first derivatives of the XANES
`spectra for the catalyst samples on the hydrothermal treatment
`are given in Fig. 7; the general patterns are basically identical.
`In fact, the weak pre-edge peaks (I), corresponding to an elec-
`tric dipole-forbidden, but quadruply and vibronically allowed
`1s → 3d electronic transition and providing direct information
`on the oxidation state of copper, can be observed at ca. 8979 eV,
`
`species on the catalyst surface
`
`which is the fingerprint of Cu2+
`[23,36].
`The primary absorption peaks (II and III) for all of the cata-
`lysts appear at ca. 8986–8988 and 8995–8998 eV, respectively,
`which are attributed to the charge transfer from the metal ligand
`to the 3d orbital (ligand-to-metal charge transfer) correspond-
`ing to the degree of Cu(3d)–O(2p) bond covalence, and 1s →
`4p transition, respectively [23,36]. The shift of peak positions
`has been also observed at the lower absorption energy. The evo-
`lution of the absorption peaks (II and III) in Fig. 7 with respect
`to the Cu content and the aging condition may be due to the
`gradual polymerization of Cu2+
`ions with oxygen ion (O2−
`)
`on the catalyst surface. To identify and evaluate the peak posi-
`tion of the absorption edge (E0) defined as the maximum of the
`
`Exhibit 2024.005
`
`

`
`52
`
`J.-H. Park et al. / Journal of Catalysis 240 (2006) 47–57
`
`◦
`Fig. 3. Effect of aging temperatures on the conversion of NO for CuZSM5 catalysts at reaction temperatures of (A) 175, (B) 200, (C) 300, and (D) 400
`C. Feed
`−1 [(F) CuZSM5-61, (2) CuZSM5-92, (a) CuZSM5-124,
`gas composition is 500 ppm NO, 500 ppm NH3, 5% O2, 10% H2O in N2 balance, and SV is 100,000 h
`(") CuZSM5-150].
`
`first derivative of XANES spectra for CuZSM5 catalysts, the
`references including CuO, Cu2O, and Cu metal have been also
`analyzed and compared as listed in Table 2, where the apparent
`trend of the shift to the lower absorption energy can be seen.
`For the catalysts aged at the harshest sintering temperature
`◦
`of 800
`C except CuZSM5-61, an additional peak (IV) appears
`at ca. 8984 eV, which is due to the transition of copper ion into
`the bulk-type CuO [23]. It indicates that the copper on the sur-
`face of the catalyst can be transferred to a crystalline CuO by
`the hydrothermal treatment. These results were reconfirmed by
`EXAFS studies on the state of Cu on the catalyst surface.
`It has been generally recognized that EXAFS is an effec-
`tive tool for evaluating interatomic distances and coordination
`numbers to explore the geometric structure of reaction sites,
`due mainly to the interaction between an absorbing atom and
`a nearly surrounding atom [22,23,36,39,40]. Fig. 8 shows the
`radial distribution functions (RDFs) of Cu of CuZSM5 cata-
`lysts with CuO as a reference compound by Cu K-edge k2-
`
`Fig. 4. 27Al-MAS-NMR spectra of CuZSM5-150 before and after aging:
`◦
`C [∗ symbol indicating spinning
`(a) fresh, (b) aged at 600, (c) 700, and (d) 800
`sidebands].
`
`Exhibit 2024.006
`
`

`
`J.-H. Park et al. / Journal of Catalysis 240 (2006) 47–57
`
`53
`
`Fig. 5. XRD patterns of the under-exchanged (CuZSM5-92) and over-ex-
`changed (CuZSM5-150) before and after aging.
`(a) CuZSM5-92-fresh,
`(b) CuZSM5-92-700,
`(c) CuZSM5-92-800,
`(d) CuZSM5-150-fresh,
`(e) CuZSM5-150-700, and (f) CuZSM5-150-800. (Inset figure is a de-
`◦
`◦
`tailed XRD patterns for the identical catalysts ranged from 33
`to 43
`of the
`diffraction angle and the dashed lines are the characteristic peak positions of
`bulk CuO.)
`
`−1 (cid:2)
`weighted Fourier transforms (FTs) in the range of 2.0 Å
`−1 with a Kaiser–Bessel function. Dominant spectra
`k (cid:2) 11.5 Å
`with a Cu–O distance of ca 1.96 Å were commonly observed for
`the catalysts examined in this study [22,23]. As clearly shown in
`Fig. 9, the best fits of the EXAFS data for the CuZSM5 samples
`were attained; their parameters are summarized in Table 2. It is
`noteworthy that the coordination number of the Cu–O shell for
`CuZSM5 samples decreases with respect to the aging tempera-
`◦
`C, indicating that Cu2+
`ion may migrate
`tures from 600 to 800
`to the inaccessible sites or transform isolated Cu2+
`ions into
`small CuO-like phases.
`An additional weak Cu–(O)–Cu bond at ca. 2.95 Å [23]
`was seen for all but the underexchanged catalysts, including
`CuZSM5-61 and CuZSM5-92-fresh, due to the local structure
`of crystalline CuO formed on the catalyst surface. The am-
`plitude of the Cu–(O)–Cu peak for the CuZSM5 catalyst was
`much weaker than that for the bulk oxides, with coordination
`number <4 [23]. As the aging condition became more severe
`and the Cu loading content increased, the amplitude of the peak
`became evident. For highly overexchanged ZSM5 catalysts,
`a Cu–(O)–Cu peak was observed even for its fresh counterpart.
`It may be concluded that Cu zeolites prepared by the solution
`ion-exchange method can contain Cu–O–Cu dimeric oxoca-
`
`Fig. 6. Cu K-edge XANES spectra for CuZSM5 catalysts before and after
`aging: (A) CuZSM5-61, (B) CuZSM5-92, (C) CuZSM5-124, and (D) Cu-
`◦
`◦
`ZSM5-150 [(a) fresh, (b) aging at 600
`C, (c) aging at 700
`C, (d) aging at
`◦
`800
`C, and reference (e) CuO].
`
`tions that may be transferred to crystalline CuO as catalyst
`aging proceeds and/or the Cu loading content of the catalyst in-
`creases. Although the XAFS studies can examine the oxidation
`state of copper on the catalyst surface, the interatomic distance,
`and the coordination number of copper ion on the surface of
`CuZSM5 catalyst to characterize the geometric structure of the
`Cu atoms, it may not be sufficient to clearly assess the charac-
`teristics of Cu2+
`species, the amount of the isolated Cu2+
`, the
`degree of transformation of Cu2+
`into bulk-type CuO, and other
`features that may be critical to elucidate the aging mechanism
`of CuZSM5 catalyst to remove NO from automotive diesel en-
`gine by urea.
`
`3.4. Migration of copper among the reaction sites over the
`catalyst on aging
`To examine the alteration of the local characteristics of Cu2+
`ions by sintering, the CuZSM5 catalysts before and after hy-
`◦
`drothermal treatment were pretreated in situ at 500
`C for 2 h in
`pure O2 flow, evacuated, and sealed off for ESR study [22]. To
`
`Exhibit 2024.007
`
`

`
`54
`
`J.-H. Park et al. / Journal of Catalysis 240 (2006) 47–57
`
`Sample
`
`Sigmad R-factor
`
`CuZSM5-61-fresh
`CuZSM5-61-600
`CuZSM5-61-700
`CuZSM5-61-800
`
`CuZSM5-92-fresh
`CuZSM5-92-600
`CuZSM5-92-700
`CuZSM5-92-800
`
`CuZSM5-124-fresh
`CuZSM5-124-600
`CuZSM5-124-700
`CuZSM5-124-800
`
`5.3
`4.9
`5.1
`4.9
`
`5.6
`4.9
`5.1
`3.9
`
`5.4
`5.3
`5.0
`3.8
`
`0.0040
`0.0040
`0.0049
`0.0044
`
`0.0057
`0.0047
`0.0053
`0.0021
`
`0.0053
`0.0050
`0.0054
`0.0015
`
`0.0010
`0.0030
`0.0053
`0.0055
`
`0.0060
`0.0017
`0.0042
`0.0137
`
`0.0031
`0.0042
`0.0055
`0.0137
`
`Table 2
`Cu K-edge EXAFS least-square fitting results in CuZSM5 samples with respect
`to copper loadings and aging temperatures
`b Ec
`C.N.a Reff
`1.976 −6.6
`1.960 −7.7
`1.959 −8.3
`1.957 −7.8
`1.956 −7.5
`1.955 −6.4
`1.952 −7.2
`1.947 −7.4
`1.955 −7.1
`1.951 −7.3
`1.954 −6.8
`1.956 −5.2
`1.955 −7.1
`1.954 −7.6
`1.959 −6.6
`1.947 −6.4
`
`Edge
`energye
`8992.2
`8992.2
`8993.2
`8992.2
`
`8991.6
`8990.8
`8991.0
`8991.6
`
`8991.6
`8990.8
`8991.8
`8990.6
`
`8991.6
`8992.2
`8991.8
`8991.4
`
`0.0048
`0.0057
`0.0050
`0.0050
`
`0.0025
`0.0025
`0.0022
`0.0020
`
`5.1
`CuZSM5-150-fresh
`5.5
`CuZSM5-150-600
`5.3
`CuZSM5-150-700
`4.6
`CuZSM5-150-800
`a Coordination number.
`b Energy shift.
`c Interatomic distance.
`d Debye–Waller factor.
`e Determined from the position of the maximum of the derivative of XANES
`spectra for the samples.
`
`Fig. 7. First-derivative of XANES spectra for CuZSM5 catalysts before and
`after aging: (A) CuZSM5-61, (B) CuZSM5-92, (C) CuZSM5-124, and (D) Cu-
`◦
`◦
`ZSM5-150 [(a) fresh, (b) aging at 600
`C, (c) aging at 700
`C, (d) aging at
`◦
`800
`C, and reference (e) CuO].
`
`obtain better resolution of ESR spectra and investigate the ef-
`fect of measurement temperature on the state of Cu ions in the
`framework of the ZSM5 catalyst, the ESR studies were con-
`ducted at both room temperature and liquid nitrogen tempera-
`ture. No significant variation of the ESR spectra was observed.
`Fig. 10 shows the X-band ESR spectra for the CuZSM5-125
`catalysts, an axially symmetrical signal split into four hyper-
`fine lines that can be resolved as the parallel component of
`ESR spectra. This fourfold hyperfine splitting of the spectra
`was caused by the coupling between the 3d unpaired electron
`and the copper (I = 3/2) nuclear spin. For an orientationally
`disordered solid assuming axial symmetry, g anisotropy pro-
`duces a powder pattern in which the sharp features are referred
`to as parallel and perpendicular edges [20,22,24,41,42].
`The spectra of fresh samples exhibited two characteristic
`perpendicular g values: Cu2+
`in a square planar environment
`(by g(cid:6) = 2.27–2.30 and A(cid:6) = 150–170 G) and square pyra-
`midal (g(cid:6) = 2.33–2.35 and A(cid:6) = 135–145 G) [20,22,24,41,
`42]. For CuZSM-5 catalysts after aging, two additional Cu2+
`species, containing g(cid:6) = 2.30 and A(cid:6) = 165 G and g(cid:6) = 2.36
`
`and A(cid:6) = 130 G of ESR signals, have been found on the cata-
`lyst surface [20,24]. The exact geometry of the third and fourth
`Cu2+
`species is not completely understood, however.
`Based on the ESR parameters examined in the present study,
`the third species (g(cid:6) = 2.30 and A(cid:6) = 165 G) obtained over
`the aged catalyst can be coordinated to a distorted square sites
`anchored on the locations recessed from the main channels
`(i.e., in the side pockets of the zeolite structure), and the fourth
`species (g(cid:6) = 2.36 and A(cid:6) = 130 G) can be assigned to slightly
`distorted octahedral [20,43]. A recessed location would be ex-
`pected to alter the redox cycle of copper presumably during
`the SCR of NOx by impairing the accessibility of the reac-
`tants, hence decreasing the overall NO removal activity [18,42].
`Similar results were also observed for the CuZSM5 catalysts
`with differing copper contents. The g and hyperfine interaction
`constant values listed in Table 3 are nearly identical to those
`reported earlier for CuZSM5 [20,24].
`Double-integral (DI) values of ESR spectra for CuZSM5 cat-
`alysts were obtained for a comparative study to quantify the
`number of active Cu species on the catalyst surface, assum-
`ing the isolated Cu2+
`ions as the active reaction sites [22,42]
`for NO removal activity by NH3 SCR; these are given in Ta-
`ble 4. The altered number of reaction sites may be represented
`by the relative percentage of DI values of the ESR spectra for
`CuZSM5 catalysts, where the absorption area of ESR spectrum
`for CuZSM5-92-fresh is used as a basis for the calculation to
`confirm the migration of the copper ions among the reaction
`sites, an arbitrary criterion set for this comparative study. The
`relative percentages of DI values for fresh CuZSM5 catalyst
`
`Exhibit 2024.008
`
`

`
`J.-H. Park et al. / Journal of Catalysis 240 (2006) 47–57
`
`55
`
`Fig. 8. EXAFS RDF of Cu K-edge for CuZSM5 catalysts before and after
`aging: (A) CuZSM5-61, (B) CuZSM5-92, (C) CuZSM5-124, and (D) Cu-
`◦
`◦
`ZSM5-150 [(a) fresh, (b) aging at 600
`C, (c) aging at 700
`C, and (d) aging
`◦
`C, (e) CuO].
`at 800
`
`Fig. 9. Cu K-edge k2 weighted EXAFS data (!) and their best fits (–) for
`CuZSM5 cata

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket