Next Article in Journal
Solventless Mechanochemical Fabrication of ZnO–MnCO3/N-Doped Graphene Nanocomposite: Efficacious and Recoverable Catalyst for Selective Aerobic Dehydrogenation of Alcohols under Alkali-Free Conditions
Next Article in Special Issue
Present Challenges in Catalytic Emission Control for Internal Combustion Engines
Previous Article in Journal
Single-Crystal-to-Single-Crystal Transformation and Catalytic Properties of New Hybrid Perhalidometallates
Previous Article in Special Issue
The Impact of Lanthanum and Zeolite Structure on Hydrocarbon Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The H2O Effect on Cu Speciation in Cu-CHA-Catalysts for NH3-SCR Probed by NH3 Titration

1
Laboratory of Catalysis and Catalytic Processes, Dipartimento di Energia, Politecnico di Milano, via La Masa 34, 20156 Milano, Italy
2
Johnson Matthey Technology Centre, Blounts Court Road, Sonning Common, Reading RG4 9NH, UK
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(7), 759; https://doi.org/10.3390/catal11070759
Submission received: 28 May 2021 / Revised: 9 June 2021 / Accepted: 14 June 2021 / Published: 23 June 2021

Abstract

:
The present work is focused on the effect of water on NH3 adsorption over Cu-CHA SCR catalysts. For this purpose, samples characterized by different SAR (SiO2/Al2O3) ratios and Cu loadings were studied under both dry and wet conditions. H2O adversely affects NH3 adsorption on Lewis acid sites (Cu ions) over all the tested catalysts, as indicated by the decreased NH3 desorption at low temperature during TPD. Interestingly, the NH3/Cu ratio, herein regarded as an index for the speciation of Cu cations, fell in the range of 3–4 (in the presence of gaseous NH3) or 1–2 (no gaseous NH3) in dry conditions, in line with the formation of different NH3-solvated Cu species (e.g., [CuII(NH3)4]2+ and [CuII(OH)(NH3)3]+ with gaseous NH3, [Z2CuII(NH3)2]2+ and [ZCuII(OH)(NH3)]+ without gaseous NH3). When H2O was fed to the system, on the contrary, the NH3/Cu ratio was always close to 3 (or 1), while the Brønsted acidity was slightly increased. These results are consistent both with competition between H2O and NH3 for adsorption on Lewis sites and with the hydrolysis of a fraction of Z2CuII species into ZCuIIOH.

1. Introduction

NH3/urea SCR is recognized worldwide as an effective technology in controlling automotive emissions to meet increasingly stricter NOx emission regulations [1,2]. Amongst state-of-the-art metal-promoted zeolite catalysts used in this process, Cu-exchanged chabazite (Cu-CHA) shows excellent activity and extraordinary hydrothermal stability over a broad temperature range under SCR conditions, triggering intensive research on the relationship between its unique structure and its DeNOx performance [3,4,5,6,7,8,9,10,11]. In this regard, the SiO2/Al2O3 ratio (SAR) and the Cu loading content are important structure indexes, which could affect the nature of Cu cations existing in the catalyst framework and, based on their nature, location and redox properties, further impact the SCR activity [3,4,5,6,7,8,9]. Two Cu species are recognized to be present in the Cu-CHA catalyst: Z2CuII (doubly coordinated to the framework) and ZCuIIOH (balanced by one framework negative charge), whose relative ratio can be predicted based on Si/Al and Cu/Al ratios [5,6,11]. The proportion of such Cu species could also be affected by several factors, for example, hydrothermal aging [4,6,12] or hydrated/dehydrated conditions [5,6]. It was also proposed, based on DFT calculations, DRIFTS, XAS and NH3-TPD tests [4,5,8,13], that these two species are able to coordinate a different number of NH3 molecules, thus influencing, in turn, the NH3 storage capacity, a relevant parameter for the SCR process. In line with Paolucci et al. [5], our recent paper concluded that upon full catalyst saturation, the Z2CuII species could coordinate up to four NH3 molecules, while the hydroxylated CuII site could bind only three molecules [8]. The NH3 adsorption chemistry on the Lewis acid sites can be summarized by the following reactions [13]:
Z2CuII (s) + 4NH3 (g) → Z2CuII(NH3)4 (s)
ZCuIIOH (s) + 3NH3 (g) → ZCuIIOH(NH3)3 (s)
When Cu-CHA catalysts were not fully saturated and Cu ions were not able to detach from the zeolite framework, however, we observed a stoichiometry different from reaction (1) and (2) for the two copper species: Z2CuII ions adsorb two NH3 molecules while only one NH3 is bound to ZCuIIOH, in agreement with Luo et al. [4].
NH3 can also be adsorbed on the Brønsted acid sites typical of the zeolite framework, according to the adsorption chemistry reported below [12]:
NH3 (g) + ZH (s) → ZNH4 (s)
Herein, we carried out NH3 adsorption + TPD tests under both dry and wet conditions over several Cu-CHA samples with different SARs and Cu loadings, which therefore contain distinct fractions of the two Cu species populations and different Brønsted sites [5,8]. The aim of the present investigation is to verify if H2O affects the Lewis and Brønsted acidity in chabazite-based materials and the possibility to use NH3-TPD as a probe technique for Cu speciation. In the literature, this method was implemented to characterize Cu-CHA catalysts in dry environments only in [8] or in the presence of water only in [4,12]. On the contrary, to assess the role of H2O in NH3 adsorption, we compared the performance of different samples under both wet and dry conditions.

2. Results and Discussion

An example of an NH3 adsorption/desorption run is provided in Figure 1A, depicting the three stages of the test, namely isothermal adsorption, isothermal desorption, and TPD. The run was performed after having pre-oxidized the sample at 550 °C. The aim was not only to evaluate the NH3 storage capacity, but more specifically to identify the different sites present in the catalyst [8].
For this purpose, we analyze the TPD phase. As shown in Figure 1A, the temperature ramp featured a release of NH3 together with a small N2 peak (about 25 ppm). The integral of the N2 released (i.e., ≈0.68 µmol) is in line with the full reduction in CuII ions (i.e., 4.3 µmol of Cu loaded) by adsorbed NH3 according to the following stoichiometries [13,14]:
6CuII (s) + 2NH3 (g) → N2 (g) + 6H (s) + 6CuI (s)
6CuIIOH (s) + 2NH3 (g) → N2 (g) + 6H2O (g) + 6CuI (s)
Figure 1B displays the NH3 and N2 traces during the TPD, whereas the corresponding integral balances are reported in Table 1. The N balance error improves significantly if the production of nitrogen is taken into account. In line with this result, Borfecchia et al. showed by means of XAS spectroscopy and UV-vis-NIR experiments that at the end of the TPD, after NH3 adsorption, the Cu sites in their Cu-CHA catalysts were almost completely reduced [13].
In line with the literature, the TPD curve in Figure 1B showed two distinct NH3 release peaks, ascribed to energetically different populations of acidic sites [4,8,15] (Figure 2). The physical nature of the different sites was estimated via Gaussian deconvolution of the NH3-TPD trace, to which the contribution of N2 according to the reactions discussed above (4) and (5) was added (green line in Figure 1B and Figure 2). The N2 trace was not measured in our previous investigation [8]; however, it is important to include this contribution to quantify more accurately the various Cu sites taking part in the SCR chemistry.
As established in several works within the literature [4,8,12,16], the first peak in Figure 1B is related to the NH3 adsorbed onto both Cu species and a small amount of Al extra-framework (quantified in [8] and reported in Table S1), while the second one is attributed to NH3 desorption from Brønsted sites associated with the bare zeolite framework. It is notable that the nitrogen production in Figure 1A was well aligned with the low temperature peak of the NH3-TPD curve, implying that mostly Lewis NH3 is responsible for the reduction in CuII cations into CuI according to (4) and (5).

2.1. H2O Effect with Changes in Cu Loading

Figure 3 shows the Cu loading effect on the NH3-TPD data collected over each catalyst under both dry and wet conditions.
Figure 3A refers to the simplest case, i.e., the TPD profiles over the unpromoted chabazite catalyst with SAR 25. The curves obtained in the presence and absence of H2O are almost overlapped, showing a single desorption peak centered at about 440 °C with a similar intensity in both tests. This peak corresponds to the desorption of NH3 stored onto the Brønsted acid sites of the zeolite [11,17,18,19]. The small shoulder visible at 250 °C under dry conditions (black line) is linked to the extra-framework aluminum species, which confers a slight Lewis acidity to the catalyst [6,17,18,19,20,21]. This peak disappeared under wet conditions, in accordance with findings in the literature [6,20,21].
Interestingly, H2O significantly affects the weakly bound NH3, since H2O molecules are stored preferentially on Lewis acid sites. A clear confirmation of this behavior was obtained comparing the NH3 desorption profiles of the copper-exchanged samples under dry and wet conditions. Figure 3B–D show the TPD results for the zeolite catalysts with the same SAR (25) and loaded with 0.7, 1.7 and 2.1% w/w copper, respectively. The presence of H2O considerably lowered the NH3 storage capacity of all catalysts, affecting mostly the low-T peak, associated predominantly with Cu ions. The high-temperature Brønsted NH3 peak was less affected by H2O. The negative effect of H2O on the NH3 storage capacity of Lewis sites was confirmed by the faster dynamics during the catalyst saturation in the isothermal adsorption phase, when the samples were previously treated with H2O (not shown): indeed, the dead time recorded during this phase was reduced from 530–860 s under dry conditions to 380–595 s under wet conditions. In addition, the integral of the TPD curves (NH3 + 2N2) was lower in the presence of H2O even considering the negligible contribution of the extra-framework Al sites (Tables S1 and S2).

2.2. Water Effect with Changes in SAR

Results were also collected over Cu-CHA catalysts with fixed Cu loading (1.7% w/w) and different SARs (10, 13, 22, 25): the corresponding TPD curves are shown in Figure 4. Under dry conditions, the differences in the relative intensities of the low and high temperature peaks indicate that on increasing the SAR, the overall stability of the adsorbed NH3 molecules decreased [8]. In fact, in line with our previous studies, intensities of the low-T peaks were lower than those of the high-T peaks for high SARs (22, 25), and the opposite trend was seen at low SARs (10, 13), where the intensities of the low-T peaks were higher [8]. In all cases, the overall amount of chemisorbed NH3 calculated from the integration of the TPD curves was found fairly constant upon changing the SAR, in the range of 950–1120 μmol/g.
A similar effect was observed in presence of H2O (light-blue lines). Comparing the TPD under wet and dry conditions, it is apparent that H2O reduced ammonia storage (Figure 4). Similarly to the data presented in the previous section, the low-temperature desorption feature diminished as a result of H2O addition. Furthermore, the overall chemisorbed NH3 decreased to 830–880 μmol/g for all the catalysts. Accordingly, the deadtime recorded during the adsorption phase was affected by the presence of water (not shown), decreasing from about 800 to 650 s.

2.3. H2O Effect on Cu Speciation

At this point, we used Gaussian deconvolution of the bimodal TPD profiles to compare the amount of NH3 to Cu atoms and to Brønsted acid sites in wet and dry atmosphere. Notice that the NH3 legated to the Al extra-framework was completely lost in the presence of water, as apparent in Figure 3A for H-CHA (SAR 25) and in Figure S1 for H-CHA (SAR 13).
The results in Figure 5 and in Table S1 indicate that by cofeeding water: (i) the quantity of NH3 stored on Brønsted sites slightly increased, while (ii) the Lewis NH3 decreased. This trend could be related to the hydrolysis of Z2CuII species, reaction (6) [6,22]:
Z2CuII (s) + H2O (g) → ZCuIIOH (s) + ZH (s)
In fact, according to this reaction, the transformation of Z2CuII into ZCuIIOH implies a loss in Cu-NH3 ligands. Indeed, the Z2CuII species can bind up to four NH3 molecules, while the hydroxylated Cu ions can coordinate only three molecules [5,8,23].
According to our deconvolution results, however, the increment in Brønsted NH3 in wet conditions was less than the decrease in Lewis NH3 for some samples. This behavior could be explained by a further loss of Lewis NH3 due to the competitive adsorption of water on the Cu sites [20], as indicated by the decrease in the overall NH3 storage capacity (i.e., the sum of Brønsted and Lewis NH3 in Figure 5).
The fraction of Z2CuII transformed into ZCuIIOH was estimated by the following equation:
Amount   of   Z 2 Cu II   hydrolised   initial   amount   of   Z 2 Cu II ( Table   2 )   =   N H 3   Brønsted   wet     N H 3   Brønsted   dry initial   amount   of   Z 2 Cu II ( Table   2 ) %
From the results in Table 2 it can be noticed that only a fraction of Z2CuII (40–70%) was hydrolyzed after the wet NH3 adsorption. Furthermore, even if 1.7Cu-SAR25 and 1.7Cu-SAR22 are characterized by similar speciation and Cu/Al ratio, they display different behaviors in a wet atmosphere, more specifically the 1.7Cu-SAR22 shows a more enhanced hydrolysis of the Z2CuII sites. One possible parameter affecting this reaction could be the distribution of the Al atoms in the zeolite framework. In addition, two different types of the Z2CuII sites are viable, with para and metal configuration, respectively, which show distinct features in EPR spectra and distinct reactivity under Standard SCR conditions [24]. In principle, these species could be associated with a different response to the hydrolysis reaction. A dedicated investigation of the hydrolysis reaction is ongoing.
At this point we calculated the NH3/Cu ratio, taking into account the area under the low-T TPD peak, proportional to the NH3 stored only on Cu cations, and the number of copper atoms known from the weight fraction of Cu in each sample (Table 3). Figure 6 illustrates how this ratio varies with the Cu/Al ratio.
Under dry conditions, the experimental data were in accordance with the literature [8,23]: the ratio was in the range 1–2 when only the Cu-NH3 desorbed during the TPD was considered, and in the 3–4 range when the physisorbed NH3 was also considered (area highlighted in orange, Figure 1A). In particular, the increase in Cu/Al lead to a decrease in the NH3/Cu ratio, from a value close to 4 (or 2 considering only the TPD phase) for the catalysts with the lowest SAR and Cu loading, to a value closer to 3 (or 1 considering only the TPD phase) for the other tested samples (Figure 6). When water was added to the system, we observed a shift of all the NH3/Cu ratios evaluated from the TPD to values closer to 3 (or 1 considering only the TPD phase). These results are consistent both with competition between H2O and NH3 adsorption onto Lewis sites and with the hydrolysis of a fraction of Z2CuII species into ZCuIIOH according to reaction (6), as discussed above.
At this stage of the discussion, our opinion is that analysis of the NH3-TPD data is not the best method to probe the Cu speciation. First, quantification of the Lewis and Brønsted acid sites requires deconvolution of the NH3-TPD curves. Second, during the temperature ramp the Cu atoms are reduced, producing N2; hence, it would be important to include its trace in the TPD plot to accurately analyze the experimental data. Finally, under wet conditions the Cu species are modified by the hydrolysis reaction. The occurrence of these phenomena complicates the interpretation of experimental TPD traces, thus it is suggested not to solely use this method to quantify the fraction of Z2CuII and ZCuIIOH species in the catalysts. The Cu speciation could be quantified by means of simpler techniques such as, for example, NO2 adsorption + TPD and H2 TPR, as discussed in [8,11,25]. However, NH3-TPD runs under wet conditions provide useful experimental evidence of the hydrolysis reaction over Cu-CHA materials.

3. Material and Methods

For this study, we investigated two sets of Cu-exchanged chabazite (Cu-CHA), research catalysts supplied in the form of powders by Johnson Matthey. One set was purposely characterized by different SiO2/Al2O3 (SAR) ratios (10, 13, 22, 25), with fixed Cu loading = 1.7% w/w. The other had different Cu loadings (0, 0.7, 1.7, 2.1% w/w, with fixed SAR = 25). The textural properties of some samples, comprising the BET surface area and the micropore volume, are reported in Table S3 and are typical of commercial CHA based materials [26,27]. The samples were conditioned by heating them up to 600 °C at 5 °C/min and holding the maximum temperature for 5 h, while 10% O2 and 10% H2O were fed in to remove all the impurities and the species that may be formed on the sample due to contact with atmosphere.
Sixteen milligrams of catalyst powder, sieved to obtain an average particle size of 90 micron and diluted up to 130 mg with cordierite, was loaded into a quartz flow reactor (6 mm ID) inserted in a vertical electric furnace. The temperature was controlled by a K-type thermocouple directly immersed in the catalyst bed.
Helium was used as balance gas in all the micro-reactor runs. Gases were controlled using mass flow controllers (Brooks Instruments, Hatfield, PA, USA). Water vapor was fed in by means of a saturator and the water concentration was controlled by adjusting the saturator’s temperature according to Antoine’s Law. NH3 was fed in using a 6-way valve, enabling step changes in the reactant concentrations. Temporal evolution of the species at the reactor outlet was followed by a mass spectrometer (QGA Hiden Analytical, Warrington, UK) and a UV analyzer (ABB LIMAS 11 HW, Zurich, Switzerland) arranged in a parallel configuration, which allowed the simultaneous measurement of all the species involved.
The interaction between NH3 molecules and the catalyst was studied by performing NH3 adsorption + temperature programmed desorption (TPD) runs. The catalysts were saturated using 500 ppm of NH3 and balance He (GHSV = 266,250 cm3/(gcat*h) STP), both under dry and wet conditions (5% w/w H2O), at 150 °C. After that, the catalysts were purged with He (or He + H2O) for 1 h to desorb weakly bound ammonia, and then heated in He only (or He + H2O) from 150 °C up to 550 °C at a rate of 15 °C/min to release the strongly adsorbed molecules. Prior to each experiment, in order to control the oxidation state of the catalysts, the samples were pre-treated at 550 °C, feeding 8% O2 in He for 1 h, then the adsorption temperature was reached by cooling down under the same gas feed, with a final purge in He for 15 min.
The catalysts were characterized in our previous work [8], the amount of Cu loaded and the Cu speciation is reported in Table 3. In particular, NO + NH3 TPR and ICP-MS analysis were used to quantify the number of reducible CuII sites. Meanwhile, the NO2 adsorption -TPD protocol was implemented to assess the amount of ZCuIIOH species [8]. The samples show the typical deNOx activity of CHA based materials as apparent from the NO conversions in Standard SCR conditions reported in Table 3 (i.e., GHSV = 937,500 cm3/h/gcat (STP), gcat = 0.016 g, NH3 = 500 ppm, NO = 500 ppm, H2O = 5%, O2 = 8%, T = 225 °C).

4. Conclusions

Herein, we have studied the effect of H2O on the NH3 storage onto different Cu-CHA catalysts, using standard NH3 adsorption/TPD experiments. According to our preliminary results, during the temperature ramp of the TPD run the adsorbed NH3 fully reduces the initially oxidized Cu sites, releasing a stoichiometric amount of N2. This effect is often overlooked in the literature but impacts the quantification of Cu-related and Brønsted NH3 adsorption sites.
On increasing both the SAR and Cu loading, we found the adsorbed NH3/Cu ratio decreased from 4 to 3 when H2O was absent from the gas mixture (and from 2 to 1 in the absence of gaseous NH3), which correlates nicely with the corresponding expected increase in the proportion of ZCuIIOH over Z2CuII species. Upon addition of water to the feed stream, all the tested catalysts showed NH3/Cu ratios closer to 3 (or to 1 in the absence of gaseous NH3). These results are consistent both with the hydrolysis of a fraction of Z2CuII species to form ZCuIIOH, even though competition between H2O and NH3 for adsorption onto Lewis sites cannot be ruled out.
This work is novel because it includes both the Cu reduction and the hydrolysis reaction in the analysis of the NH3-TPD data. However, the occurrence of these phenomena complicates the interpretation of the experimental TPD runs, thus it is suggested not to solely use this method to probe the Cu speciation in Cu-CHA catalysts [8].

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11070759/s1, Figure S1: Comparison between dry and wet NH3-TPD after adsorption at 150 °C on H-CHA catalysts with SAR = 13, Table S1: NH3 adsorbed on different sites as estimated from NH3-TPD deconvolution under dry conditions, Table S2: NH3 adsorbed on different sites as estimated from NH3-TPD deconvolution under dry conditions.

Author Contributions

Conceptualization, J.C., A.P.E.Y. and D.T.; Data curation, R.V., F.G. and U.I.; Investigation, R.V.; Supervision, I.N. and E.T.; Visualization, S.L., J.C., A.P.E.Y. and D.T.; Writing—original draft, R.V.; Writing—review & editing, F.G., U.I. and M.P.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Güthenke, A.; Chatterjee, D.; Weibel, M.; Krutzsch, B.; Kočí, P.; Marek, M.; Nova, I.; Tronconi, E. Current status of modeling lean exhaust gas aftertreatment catalysts. Adv. Chem. Eng. 2007, 33, 103–283. [Google Scholar] [CrossRef]
  2. Nova, I.; Tronconi, E. Urea-SCR Technology for deNOx After Treatment of Diesel Exhausts; Springer: New York, NY, USA, 2014. [Google Scholar]
  3. Wang, D.; Zhang, L.; Li, J.; Kamasamudram, K.; Epling, W.S. NH3-SCR over Cu/SAPO-34-Zeolite acidity and Cu structure changes as a function of Cu loading. Catal. Today 2014, 231, 64–74. [Google Scholar] [CrossRef]
  4. Luo, J.; Gao, F.; Kamasamudram, K.; Currier, N.; Peden, C.H.F.; Yezerets, A. New insights into Cu/SSZ-13 SCR catalyst acidity. Part I: Nature of acidic sites probed by NH3 titration. J. Catal. 2017, 348, 291–299. [Google Scholar] [CrossRef] [Green Version]
  5. Paolucci, C.; Parekh, A.A.; Khurana, I.; Di Iorio, J.R.; Li, H.; Albarracin Caballero, J.D.; Shih, A.J.; Anggara, T.; Delgass, W.N.; Miller, J.T.; et al. Catalysis in a cage: Condition-dependent speciation and dynamics of exchanged cu cations in ssz-13 zeolites. J. Am. Chem. Soc. 2016, 138, 6028–6048. [Google Scholar] [CrossRef] [PubMed]
  6. Gao, F.; Peden, C.H.F. Recent progress in atomic-level understanding of Cu/SSZ-13 selective catalytic reduction catalysts. Catalysts 2018, 8, 140. [Google Scholar] [CrossRef] [Green Version]
  7. Rizzotto, V.; Chen, D.; Tabak, B.M.; Yang, J.Y.; Ye, D.; Simon, U.; Chen, P. Spectroscopic identification and catalytic relevance of NH4+ intermediates in selective NOx reduction over Cu-SSZ-13 zeolites. Chemosphere 2020, 250, 126272. [Google Scholar] [CrossRef] [PubMed]
  8. Villamaina, R.; Liu, S.; Nova, I.; Tronconi, E.; Ruggeri, M.P.; Collier, J.; York, A.; Thompsett, D. Speciation of Cu Cations in Cu-CHA Catalysts for NH3-SCR: Effects of SiO2/AlO3 Ratio and Cu-Loading Investigated by Transient Response Methods. ACS Catal. 2019, 9, 8916–8927. [Google Scholar] [CrossRef]
  9. Song, J.; Wang, Y.; Walter, E.D.; Washton, N.M.; Mei, D.; Kovarik, L.; Engelhard, M.H.; Prodinger, S.; Wang, Y.; Peden, C.H.F.; et al. Toward Rational Design of Cu/SSZ-13 Selective Catalytic Reduction Catalysts: Implications from Atomic-Level Understanding of Hydrothermal Stability. ACS Catal. 2017, 7, 8214–8227. [Google Scholar] [CrossRef]
  10. Fan, C.; Chen, Z.; Pang, L.; Ming, S.; Zhang, X.; Albert, K.B.; Liu, P.; Chen, H.; Li, T. The influence of Si/Al ratio on the catalytic property and hydrothermal stability of Cu-SSZ-13 catalysts for NH3-SCR. Appl. Catal. A Gen. 2018, 550, 256–265. [Google Scholar] [CrossRef]
  11. Gao, F.; Washton, N.M.; Wang, Y.; Kollár, M.; Szanyi, J.; Peden, C.H.F. Effects of Si/Al ratio on Cu/SSZ-13 NH3-SCR catalysts: Implications for the active Cu species and the roles of Brønsted acidity. J. Catal. 2015, 331, 25–38. [Google Scholar] [CrossRef] [Green Version]
  12. Daya, R.; Joshi, S.Y.; Luo, J.; Dadi, R.K.; Currier, N.W.; Yezerets, A. On kinetic modeling of change in active sites upon hydrothermal aging of Cu-SSZ-13. Appl. Catal. B Environ. 2020, 263, 118368. [Google Scholar] [CrossRef]
  13. Borfecchia, E.; Negri, C.; Lomachenko, K.A.; Lamberti, C.; Janssens, T.V.W.; Berlier, G. Temperature-dependent dynamics of NH3-derived Cu species in the Cu-CHA SCR catalyst. React. Chem. Eng. 2019, 4, 1067–1080. [Google Scholar] [CrossRef]
  14. Giordanino, F.; Borfecchia, E.; Lomachenko, K.A.; Lazzarini, A.; Agostini, G.; Gallo, E.; Soldatov, A.V.; Beato, P.; Bordiga, S.; Lamberti, C. Interaction of NH3 with Cu-SSZ-13 catalyst: A complementary FTIR, XANES, and XES study. J. Phys. Chem. Lett. 2014, 5, 1552–1559. [Google Scholar] [CrossRef]
  15. Ruggeri, M.P.; Nova, I.; Tronconi, E.; Collier, J.E.; York, A.P.E. Structure–Activity Relationship of Different Cu–Zeolite Catalysts for NH3–SCR. Top. Catal. 2016, 59, 875–881. [Google Scholar] [CrossRef]
  16. Hu, W.; Selleri, T.; Gramigni, F.; Fenes, E.; Rout, K.R.; Liu, S.; Nova, I.; Chen, D.; Gao, X.; Tronconi, E. On the Redox Mechanism of Low-Temperature NH3-SCR over Cu-CHA: A Combined Experimental and Theoretical Study of the Reduction Half Cycle. Angew. Chem. 2021, 133, 7273–7280. [Google Scholar] [CrossRef]
  17. Gao, F.; Wang, Y.; Washton, N.M.; Kollár, M.; Szanyi, J.; Peden, C.H.F. Effects of Alkali and Alkaline Earth Cocations on the Activity and Hydrothermal Stability of Cu/SSZ-13 NH3-SCR Catalysts. ACS Catal. 2015, 5, 6780–6791. [Google Scholar] [CrossRef]
  18. Han, S.; Ye, Q.; Cheng, S.; Kang, T.; Dai, H. Effect of the hydrothermal aging temperature and Cu/Al ratio on the hydrothermal stability of CuSSZ-13 catalysts for NH3-SCR. Catal. Sci. Technol. 2017, 7, 703–717. [Google Scholar] [CrossRef]
  19. Bagnasco, G. Improving the selectivity of NH3 TPD measurements. J. Catal. 1996, 159, 249–252. [Google Scholar] [CrossRef]
  20. Sjovall, H.; Blint, R.J.; Olsson, L. Detailed Kinetic Modeling of NH3 and H2O Adsorption, and NH3 Oxidation over Cu-ZSM-5. J. Phys. Chem. C 2009, 113, 1393–1405. [Google Scholar] [CrossRef]
  21. Bolis, V.; Busco, C.; Ugliengo, P. Thermodynamic study of water adsorption in high-silica zeolites. J. Phys. Chem. B 2006, 110, 14849–14859. [Google Scholar] [CrossRef] [PubMed]
  22. Gao, F.; Mei, D.; Wang, Y.; Szanyi, J.; Peden, C.H.F. Selective Catalytic Reduction over Cu/SSZ-13: Linking Homo- and Heterogeneous Catalysis. J. Am. Chem. Soc. 2017, 139, 4935–4942. [Google Scholar] [CrossRef] [PubMed]
  23. Paolucci, C.; Khurana, I.; Parekh, A.A.; Li, S.; Shih, A.J.; Li, H.; Di Iorio, J.R.; Albarracin-Caballero, J.D.; Yezerets, A.; Miller, J.T.; et al. Dynamic multinuclear sites formed by mobilized copper ions in NOx selective catalytic reduction. Science 2017, 357, 898–903. [Google Scholar] [CrossRef] [Green Version]
  24. Godiksen, A.; Isaksen, O.L.; Rasmussen, S.B.; Vennestrøm, P.N.R.; Mossin, S. Site-Specific Reactivity of Copper Chabazite Zeolites with Nitric Oxide, Ammonia, and Oxygen. ChemCatChem 2018, 10, 366–370. [Google Scholar] [CrossRef]
  25. Hun Kwak, J.; Zhu, H.; Lee, J.H.; Peden, C.H.F.; Szanyi, J. Two different cationic positions in Cu-SSZ-13? Chem. Commun. 2012, 48, 4758–4760. [Google Scholar] [CrossRef] [PubMed]
  26. Rutkowska, M.; Duda, M.; Kowalczyk, A.; Chmielarz, L. Modification de la zéolithe CHA commerciale pour les besoins de la réduction catalytique des oxydes d’azote. C. R. Chim. 2017, 20, 850–859. [Google Scholar] [CrossRef]
  27. McEwen, J.S.; Anggara, T.; Schneider, W.F.; Kispersky, V.F.; Miller, J.T.; Delgass, W.N.; Ribeiro, F.H. Integrated operando X-ray absorption and DFT characterization of Cu-SSZ-13 exchange sites during the selective catalytic reduction of NOx with NH3. Catal. Today 2012, 184, 129–144. [Google Scholar] [CrossRef]
Figure 1. (A) NH3 adsorption + TPD test over 1.7Cu-CHA 25 catalyst. GHSV = 266,250 cm3/(gcat*h) (STP). Pre-oxidized catalysts. NH3 = 500 ppm, TPD heating rate = 15 °C/min, He. (B) TPD plot.
Figure 1. (A) NH3 adsorption + TPD test over 1.7Cu-CHA 25 catalyst. GHSV = 266,250 cm3/(gcat*h) (STP). Pre-oxidized catalysts. NH3 = 500 ppm, TPD heating rate = 15 °C/min, He. (B) TPD plot.
Catalysts 11 00759 g001
Figure 2. Deconvolution of the NH3-TPD profile over 1.7Cu-CHA 25 catalyst in Figure 1B.
Figure 2. Deconvolution of the NH3-TPD profile over 1.7Cu-CHA 25 catalyst in Figure 1B.
Catalysts 11 00759 g002
Figure 3. Comparison between dry and wet NH3-TPD (NH3 + 2N2) after adsorption at 150 °C on Cu-CHA catalysts with SAR = 25 and different Cu loadings: (A) 0% w/w; (B) 0.7% w/w; (C) 1.7% w/w; (D) 2.1% w/w. NH3 = 500 ppm, H2O = 0–5% v/v; heating rate = 15 °C/min, He. GHSV = 266,250 cm3/(gcat*h) (STP). Pre-oxidized catalysts.
Figure 3. Comparison between dry and wet NH3-TPD (NH3 + 2N2) after adsorption at 150 °C on Cu-CHA catalysts with SAR = 25 and different Cu loadings: (A) 0% w/w; (B) 0.7% w/w; (C) 1.7% w/w; (D) 2.1% w/w. NH3 = 500 ppm, H2O = 0–5% v/v; heating rate = 15 °C/min, He. GHSV = 266,250 cm3/(gcat*h) (STP). Pre-oxidized catalysts.
Catalysts 11 00759 g003
Figure 4. Comparison between dry and wet NH3-TPD (NH3 + 2N2) after adsorption at 150 °C on Cu-CHA catalysts with Cu loading = 1.7% w/w and different SAR: (A) 10; (B) 13; (C) 22; (D) 25. NH3 = 500 ppm, H2O = 0–5% v/v; heating rate = 15 °C/min, He. GHSV = 266,250 cm3/(gcat*h) (STP). Pre-oxidized catalysts.
Figure 4. Comparison between dry and wet NH3-TPD (NH3 + 2N2) after adsorption at 150 °C on Cu-CHA catalysts with Cu loading = 1.7% w/w and different SAR: (A) 10; (B) 13; (C) 22; (D) 25. NH3 = 500 ppm, H2O = 0–5% v/v; heating rate = 15 °C/min, He. GHSV = 266,250 cm3/(gcat*h) (STP). Pre-oxidized catalysts.
Catalysts 11 00759 g004
Figure 5. Deconvolution of the TPD profiles: evaluation of NH3 stored on Lewis and Brønsted sites. Comparison between dry and wet conditions over different Cu-CHA samples: (A) 0.7% w/w Cu, SAR = 25; (B) 1.7% w/w Cu, SAR = 25; (C) 2.1% w/w Cu, SAR = 25; (D) 1.7% w/w Cu, SAR = 10; (E) 1.7% w/w Cu, SAR = 13; (F) 1.7% w/w Cu, SAR = 22. NH3 = 500 ppm, H2O = 0–5% v/v, He. GHSV = 266,250 cm3/(gcat*h) (STP). Pre-oxidized catalysts. Red symbols = Error bar.
Figure 5. Deconvolution of the TPD profiles: evaluation of NH3 stored on Lewis and Brønsted sites. Comparison between dry and wet conditions over different Cu-CHA samples: (A) 0.7% w/w Cu, SAR = 25; (B) 1.7% w/w Cu, SAR = 25; (C) 2.1% w/w Cu, SAR = 25; (D) 1.7% w/w Cu, SAR = 10; (E) 1.7% w/w Cu, SAR = 13; (F) 1.7% w/w Cu, SAR = 22. NH3 = 500 ppm, H2O = 0–5% v/v, He. GHSV = 266,250 cm3/(gcat*h) (STP). Pre-oxidized catalysts. Red symbols = Error bar.
Catalysts 11 00759 g005
Figure 6. NH3/Cu ratio evaluated considering only NH3 stored on Cu ions (low-T peak of TPD) and NH3 stored on Cu ions plus NH3 physisorbed: comparison between dry and wet conditions.
Figure 6. NH3/Cu ratio evaluated considering only NH3 stored on Cu ions (low-T peak of TPD) and NH3 stored on Cu ions plus NH3 physisorbed: comparison between dry and wet conditions.
Catalysts 11 00759 g006
Table 1. Integral balances for the NH3 adsorption + TPD test over 1.7Cu-CHA25 shown in Figure 1.
Table 1. Integral balances for the NH3 adsorption + TPD test over 1.7Cu-CHA25 shown in Figure 1.
1.7Cu-CHA25
Cutot (ICP) (µmol) 4.3
NH3 adsorbed (µmol) 23.4
NH3 physisorbed (µmol) 8.0
NH3 chemissorbed (µmol) 14.1
N2 produced (µmol) 0.7
N balance error w/o N2 (%) 5.6
N balance error with N2 (%) 0.1
Table 2. Estimates of the fraction of Z2CuII hydrolyzed according to reaction (3).
Table 2. Estimates of the fraction of Z2CuII hydrolyzed according to reaction (3).
SampleCu Loaded, µmolZ2CuII from NO2 Adsorption, µmol [8]NH3 Brønsted dry, µmolNH3 Brønsted wet, µmolZ2CuII Hydrolised/Initial Z2CuII, %Z2CuII Hydrolised/Cutot, %
0.7% w/w Cu, SAR = 251.80.88.028.384520
1.7% w/w Cu, SAR = 254.31.87.98.674319
2.1% w/w Cu, SAR = 255.21.77.078.075919
1.7% w/w Cu, SAR = 104.32.46.207.545631
1.7% w/w Cu, SAR = 134.33.18.309.503928
1.7% w/w Cu, SAR = 224.31.88.009.287130
Table 3. Characterization of the tested catalyst samples [8].
Table 3. Characterization of the tested catalyst samples [8].
SampleCu Loaded, µmolCu/Al Ratio, -NO2 Adsorption Experiment, µmolZCuIIOH, %Z2CuII, %NOx Conversion at 225 °C, %
0.7% w/w Cu, SAR = 251.80.111.0554516
1.7% w/w Cu, SAR = 254.30.242.5584291
2.1% w/w Cu, SAR = 255.20.293.5673396
1.7% w/w Cu, SAR = 104.30.121.9445658
1.7% w/w Cu, SAR = 134.30.171.2287267
1.7% w/w Cu, SAR = 224.30.222.5584183
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Villamaina, R.; Gramigni, F.; Iacobone, U.; Liu, S.; Nova, I.; Tronconi, E.; Ruggeri, M.P.; Collier, J.; York, A.P.E.; Thompsett, D. The H2O Effect on Cu Speciation in Cu-CHA-Catalysts for NH3-SCR Probed by NH3 Titration. Catalysts 2021, 11, 759. https://doi.org/10.3390/catal11070759

AMA Style

Villamaina R, Gramigni F, Iacobone U, Liu S, Nova I, Tronconi E, Ruggeri MP, Collier J, York APE, Thompsett D. The H2O Effect on Cu Speciation in Cu-CHA-Catalysts for NH3-SCR Probed by NH3 Titration. Catalysts. 2021; 11(7):759. https://doi.org/10.3390/catal11070759

Chicago/Turabian Style

Villamaina, Roberta, Federica Gramigni, Umberto Iacobone, Shaojun Liu, Isabella Nova, Enrico Tronconi, Maria Pia Ruggeri, Jillian Collier, Andrew P. E. York, and David Thompsett. 2021. "The H2O Effect on Cu Speciation in Cu-CHA-Catalysts for NH3-SCR Probed by NH3 Titration" Catalysts 11, no. 7: 759. https://doi.org/10.3390/catal11070759

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop