Lewis Acidity Inherent to the Framework of Zeolite Mordenite

Switzerland. 2) Laboratory for Catalysis and Sustainable Chemistry, Paul Scherrer Institute, Villigen 5232, Switzerland. Supporting Information Placeh...
2 downloads 0 Views 436KB Size
Subscriber access provided by Bethel University

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Lewis Acidity Inherent to the Framework of Zeolite Mordenite Manoj Ravi, Vitaly L. Sushkevich, and Jeroen A. van Bokhoven J. Phys. Chem. C, Just Accepted Manuscript • Publication Date (Web): 28 May 2019 Downloaded from http://pubs.acs.org on May 28, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Lewis Acidity Inherent to the Framework of Zeolite Mordenite. Manoj Ravi1, Vitaly L. Sushkevich2, Jeroen A. van Bokhoven1,2* 1) Institute for Chemical and Bioengineering, Department of Chemistry and Applied Biosciences, ETH Zurich, 8093 Zurich, Switzerland 2) Laboratory for Catalysis and Sustainable Chemistry, Paul Scherrer Institute, Villigen 5232, Switzerland Supporting Information Placeholder ABSTRACT: Despite being used extensively as heterogeneous catalysts and supports at the academic and industrial levels alike, the nature of aluminum-based Lewis acidity in zeolites is not completely understood. In this study, we discovered a correlation between Lewis acidity and aluminum structure in the framework of zeolite mordenite. The amount of octahedrally coordinated aluminum in these samples present under wet conditions correlates to the number of Lewis acid sites as detected by FTIR spectroscopy of adsorbed probe molecules in fully dehydrated state. We illustrate that this aluminum species, which are better considered as framework-associated and not extra-framework aluminum, have Lewis acidic properties when treated under vacuum at high temperature. This observation constitutes important progress in understanding the structure of Lewis acid sites. The Lewis acid sites are framework aluminum species which undergo a change in coordination, from tetrahedral to octahedral, depending on the charge-balancing cation and degree of hydration. These Lewis acid sites exist in the case of chargebalancing by protons.

Introduction Zeolites continue to be an indispensable class of materials in catalysis research, as well as in chemical industries, to catalyze a wide array of reactions that includes cracking,1 alkylation,2 acylation,3 and other reactions of interest for the production of fine chemicals.4, 5 Over the years, a number of synthesis and characterization techniques have been developed for these microporous, crystalline aluminosilicates. The structure of aluminum species in zeolites and that of catalytically active sites, particularly those of Lewis acidic aluminum sites remain a focal point in zeolite research. Lewis acid sites in zeolites, often obtained by introducing heteroatoms such as titanium, tin and zirconium, have been exploited to catalyze reactions such as epoxidation,6 aldol condensation,7 and Baeyer-Villiger oxidation.8 In general, Lewis acidic zeolites also find extensive applications in the conversion of biomass.9 This includes the conversion of sugars to lactates and dehydration of sugars to furan compounds.10 Particularly for zeolites, 27Al MAS NMR is the most common technique used to characterize different aluminum species in terms of their coordination,11 while the acidic properties of zeolites can be examined by a range of methods,12 including 1H MAS NMR,13, 14 Fourier-transform Infrared (FTIR) spectroscopy of adsorbed probe molecules,15 such as pyridine and carbon monoxide, and NH3-Temperature Programmed Desorption.16 In a 27Al MAS NMR spectrum of a hydrated zeolite, tetrahedrally coordinated framework aluminum (Td) is characterized by a narrow resonance between about 50 and 60 ppm and octahedrally

coordinated aluminum (Oh) by a resonance around 0 ppm. In some zeolites, such as zeolite BEA, the resonance at 0 ppm is generally a superposition of two contributions: well-ordered octahedral aluminum and distorted octahedral aluminum.17 Numerous studies, with many of them based on zeolite BEA, have tried to discern the role of these octahedrally coordinated aluminum species, distinguishing them as extra-framework or framework-associated, depending on aluminum’s ability to reversibly change its coordination.18-22 In the case of the latter, octahedrally coordinated aluminum in protonic zeolites quantitatively reverts to tetrahedral coordination on heating the zeolite to higher temperatures or exchanging the zeolite to its sodium or ammonium form.18-25 By primarily studying the effects of steaming and activity in the the Meerwein-Ponndorf-Verley (MPV) reaction, a wealth of novel, if at times conflicting, findings have emerged. Octahedral aluminum species connected to the framework structure of zeolite BEA was found to exhibit characteristics of Lewis acid sites.26 However, the absence of a correlation between catalytic activity in the Lewis acid-catalyzed MPV reduction of ketones and the relative amounts of octahedrally coordinated aluminum in different H-BEA samples led to the conclusion that such aluminum species are not mandatory for catalytic activity.27 Likewise, it was reported that framework aluminum atoms in a non-tetrahedral coordination may be associated with Lewis acidic properties28, 29 and consequently, partially hydrolyzed framework aluminum was proposed as the precursor for the MPV active site.23 Furthermore, quantitative studies hint at cationic extra-framework aluminum species being located at ion-exchange positons in zeolite BEA.17 Emanating from these studies is a lack of consensus on the structure of Lewis acid sites in zeolites, a vague understanding of the role of framework aluminum in catalytic activity, and an improbability to relate aluminum of a particular structure or coordination definitively to Lewis acidity. Experimental Section Materials Preparation Mordenite zeolites of different Si/Al ratios were obtained from commercial suppliers as follows: CBV10A, Si/Al = 6.5, in sodium form; CBV21A, Si/Al = 10, in ammonium form; CBV90A, Si/Al = 46, in proton form were purchased from Zeolyst International. Zeoflair 800, Si/Al = 8.5, in sodium form; Zeoflair 810, Si/Al = 19, in proton form were obtained from Zeochem. Ammonium exchange procedure As received MOR zeolites were converted to the ammonium form by the following procedure: 1 g of zeolite was stirred in 80 mL of 0.1 - 0.5 M solution of ammonium nitrate (> 99%, SigmaAldrich) at room temperature or 353 K overnight with pH

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

monitoring. The suspension was then filtered at room temperature and rinsed with 300 mL of deionized water. This procedure was repeated twice to obtain the mordenite zeolite in ammonium form. The ammonium form was calcined in an oven in static air at 823 K for 6 h with a heating ramp rate of 1 K min-1. The calcined Hforms are labelled as H-MOR(8), H-MOR(9), H-MOR(11), HMOR(19), and H-MOR(48) with the number in the brackets indicating the Si/Al ratio of the zeolite. Acid treatment procedure Acid treatment of MOR samples resulting in proton-exchanged MOR were obtained as follows: Na-MOR was stirred in suitable concentrations of nitric acid (0.1-1 M) for the desired time (1-24 h) at the desired temperature (298-323 K). The resulting solids were washed extensively with milli-Q water and dried at 373 K overnight. Subsequently, the sample was calcined in static air at 823 K for 6 h with a heating ramp rate of 1 K min-1. Sodium exchange procedure The following procedure was used to exchange zeolites to the sodium form: 1 g of zeolite was stirred in 60 mL of 0.1 M solution of sodium nitrate (> 99%, Sigma-Aldrich) at room temperature overnight with pH monitoring. The suspension was then filtered and rinsed with 300 mL of deionized water. The procedure was repeated twice. The zeolite was dried overnight at 373 K in a drying oven and subsequently calcined in static air at 823 K for 6 h with a heating ramp rate of 1 K min-1. Material Characterization Nitrogen Physisorption N2 sorption–desorption isotherms were measured at 77 K using a Micromeritics Tristar automatic surface area and pore size analyzer. Prior to the measurements, samples were degassed for 4h under vacuum for accurate mass measurements. Solid state 27Al-MAS-NMR 27Al-MAS-NMR spectra of the zeolites were recorded on a Bruker 400 MHz Ultra-Shield spectrometer. The spectra were accumulated from 3000 scans using a 4 mm probe at a sample spinning rate of 10 kHz. The 27Al chemical shift was referenced to AlNH4(SO4)2·12H2O at −0.54 ppm and the spectra were acquired using a pulse sequence at a pulse angle of 90 deg and a recovery delay of 1s. The mass of the sample loaded into the ZrO2 rotor was recorded to normalize the corresponding spectra per unit mass. The relative amount of aluminum in octahedral coordination was found based on the relative area of the peak centered at 0 ppm. FTIR spectroscopy of CO adsorbed over zeolites IR spectra were recorded on a Thermo Nicolet iS50 FTIR spectrometer equipped with DTGS detector at a 4 cm-1 optical resolution and 128 scans. Prior to the measurements, the sample (about 20mg) was pressed in self-supporting discs and activated in the IR transmission cell attached to a vacuum line at 723 K for 4 h. A low temperature vacuum cell cooled with liquid nitrogen was used for carbon monoxide (CO) adsorption measurements. Calibrated aliquotes of the gas was introduced into the cell and the spectra were collected immediately. The pressure was measured by a Pfeiffer gauge. Difference spectra were obtained by the subtraction of the spectra of the activated samples from the spectra of samples with the adsorbate. The presented spectra were normalized to the weight of the samples and compared at similar CO coverage. The subtraction was performed using the OMNIC 9.1 software package. FTIR spectroscopy of pyridine adsorbed over zeolites The activation procedure was the same as described above using CO as the probe molecule. For pyridine (Py) adsorption, the samples were exposed to 3 torr of Py at 423 K for 30 min for the

Page 2 of 7

complete diffusion of probe molecule and then evacuated at different temperatures for 30 min. For quantifications, molar integral extinction coefficients of 2.22 cm/µmol and 1.67 cm/µmol were used for Lewis and Brønsted acid sites, respectively.1 The peaks at 1455 cm-1 and 1555 cm-1 were used for determination of Lewis and Brønsted acid sites, respectively. Atomic absorption spectroscopy The silicon and aluminum content of various zeolites were determined by AAS. The zeolites (20-50 mg) were dissolved in a mixture of hydrofluoric acid (2-5 mL) and 60% nitric acid (2-5 mL) and diluted with DI-water (25-40 mL). The mixture was stirred at room temperature and diluted to the range of calibration. The resulting solutions were measured with an Agilent SpectrAA 220FS instrument. Results and Discussion While often in the past, extra-framework aluminum has conveniently been cited as the reason for Lewis acidity, curated data suggests that such a correlation does not necessarily hold true at all times (Figure S1). This does not mean that extra-framework debris does not have Lewis acid sites, but that such acidity can also stem from other sites that are inherent to the zeolite framework structure and the endeavor of the present work is to identify such sites. To address this pressing question of Lewis acidity-aluminum structure relationship, we turned to MOR zeolites. Mordenite is widely used as a heterogeneous catalyst and support in extensively researched reactions, such as hydroisomerization of hydrocarbons,30 methane-to-methanol31 and selective catalytic reduction.32 Five MOR zeolite samples, with different Si/Al ratios given in brackets, were used in this study. Physical characterization by nitrogen adsorption-desorption indicate the high micropore volume of the MOR samples irrespective of the aluminum content (Table 1). The sample with the highest Si/Al ratio, H-MOR(48), has the highest mesoporous volume (approx. 0.11 cm3/g) and this might be due to the post-synthetic treatment of the zeolite by the supplier to remove aluminum.

Table 1. Characterization of MOR zeolites used in this study. Sample Micropore Total pore volume (Bulk Si/Al) volume [cm3/g]a [cm3/g]b

Al content Octahedral [µmol/g]c Al relative to total Al [%]d

H-MOR(8)

0.18

0.21

1815

12

H-MOR(9)

0.20

0.23

1621

13

H-MOR(11) 0.18

0.22

1402

11

H-MOR(19) 0.19

0.22

818

10

H-MOR(48) 0.17

0.28

302

1

aFrom t-plot of Nitrogen-physisorption experiments, bBased on single point adsorption at p/po = 0.97, cCalculated based on Si/Al ratio obtained from AAS, dDetermined from the area under the 0 ppm peak of 27Al MAS NMR spectra.

The 27Al MAS NMR spectra of the H-MOR zeolites after being calcined in static air and stored under ambient conditions show two resonances: a peak at 58 ppm, associated to tetrahedral framework aluminum (Td) and a sharp peak centered at 0 ppm, indicative of well-ordered octahedral aluminum species (Oh) with the absence of distorted octahedral aluminum (Figure 1). The tetrahedral peak has a broad shoulder attached to it, potentially suggesting the presence of other aluminum sites in a distorted

ACS Paragon Plus Environment

Page 3 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry tetrahedral coordination. In addition, this broad extension is more apparent in samples with higher aluminum content.

Figure 1. 27Al MAS NMR spectra of the H-MOR zeolites of different Si/Al ratios.

The population of aluminum in octahedral coordination is not present in the sodium or ammonium forms, which both show only one resonance characteristic of tetrahedrally coordinated framework aluminum (Figure S3). Moreover, this octahedrally coordinated aluminum species is best considered as frameworkassociated and not as extra-framework aluminum. Since this aluminum species adopts a typical framework tetrahedral coordination on back-exchanging to its sodium or ammonium form (Figure S3), the octahedral aluminum can be visualized as at least being partially attached to the zeolite framework or linked to its original location. Aluminum species that are characterized by such a sharp peak at around 0 ppm in other zeolite frameworks are flexible to be reinserted into the framework.33, 34 Extra-framework aluminum with octahedral coordination generally results in a broader feature at around 0 ppm and reinsertion of this species back into the framework of the zeolite is generally not possible. However with the MOR zeolites, we observe an enrichment of the framework with tetrahedrally coordinated aluminum upon exchange of the protonic zeolite to its ammonium form (Table S1), thereby demonstrating the importance to view the octahedral species as being linked to the framework. The flexibility of aluminum coordination in zeolite MOR has been observed by XANES at the Al K-edge21 and such framework-associated aluminum has previously been identified in zeolites such as ZSM-5,24 zeolite Y25 and BEA.18 Likewise, it has been shown that aluminum eliminated from the framework of zeolites BEA and MOR can easily be reinserted back into the framework.35 Interestingly, the possibility to drive octahedral aluminum in H-MOR back into tetrahedral coordination has also been hypothesized during copper exchange.36 The acid sites in zeolite mordenite were probed by FTIR spectroscopy using pyridine and carbon monoxide as probe molecules. The two molecules differ in terms of strength and size and may be used in parallel to obtain deeper insight on the nature of acid sites being probed. Figure 2 shows the FTIR spectra of adsorbed pyridine over the calcined H-MOR samples. The peaks at 1455 cm-1 and 1621 cm-1 are associated to the interaction of pyridine with Lewis acid sites. The bands at 1545 and 1635 cm-1 are assigned to Brønsted bound pyridine and that at 1490 cm-1 is structure insensitive.37, 38 Likewise, these acid sites were examined with carbon monoxide as the probe molecule (Figure 3). The absorption feature at 2223 cm-1 is characteristic of the interaction of carbon monoxide with three-coordinated Lewis acidic aluminum while that at 2198 cm-1 is attributed to five-coordinated Lewis acidic aluminum probed by carbon monoxide. The band at 2173 cm-1 arises from the interaction of carbon monoxide with Brønsted acid sites in the zeolite and the feature at 2138 cm-1 is due to physisorbed liquid-like carbon monoxide. All samples have a greater proportion of three-coordinated Lewis acidic aluminum than five-coordinated Lewis acidic aluminum sites. Furthermore, appreciable amounts of the five-coordinated Lewis acidic species can only be found in samples with high aluminum content (Figure 3). These three and five-coordinated Lewis acidic aluminum sites as observed here by FTIR spectroscopy is due to dehydration during sample pre-treatment. The removal of water from the coordination sphere facilitates the interaction of carbon monoxide with such sites.

Figure 2. FTIR spectra of adsorbed pyridine over H-MOR zeolites of different Si/Al ratios. (L) indicates Lewis bound pyridine, (B) indicates Brønsted bound pyridine and (L+B) indicates pyridine associated with both Brønsted and Lewis sites.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

spectra (Figure 1) suggests the presence of such species in HMOR.

Figure 4. Correlating Lewis acidity determined by FTIR spectroscopy of adsorbed pyridine (squares, left ordinate axis) and adsorbed carbon monoxide (diamonds, right ordinate axis) in HMOR to the amount of octahedral aluminum determined by 27Al MAS NMR. Quantitative data on Lewis acidity in the experiments with pyridine were extracted based on the peak at 1455 cm-1. Relative amounts of Lewis acidity based on experiments with carbon monoxide were calculated by the sum of the areas of the deconvoluted peaks centered at 2223 and 2198 cm-1.

Figure 3. FTIR spectra of adsorbed carbon monoxide over HMOR zeolites of different Si/Al ratios. (B) indicates Brønsted acid sites probed by carbon monoxide. Lewis acidic aluminum in coordination with carbon monoxide are denoted as (Al3c-CO)δ+ and (Al5c-CO)δ+ Figure 4 illustrates a linear relation between the amount of octahedrally coordinated aluminum in the wet conditions of a solid-state NMR experiment and the Lewis acidity as detected by FTIR spectroscopy in the H-MOR samples. This correlation, which is observed irrespective of the choice of the probe molecule, suggests that aluminum species that are in octahedral coordination at room temperature in a hydrated environment are associated with Lewis acidity when treated at high temperatures under vacuum. Thus, these results, which are obtained by using two vastly varying techniques in terms of sample environment, aid in unequivocally assigning six-coordinated aluminum in HMOR to Lewis acidic properties. The generation of octahedral aluminum in zeolites from tetrahedral framework aluminum is driven by the adsorption of water and hydrolysis reactions.25, 39, 40 Bearing this in mind, the relationship between octahedral aluminum and Lewis acidity (Figure 4) is intuitive to envisage because an aluminum species that is capable of changing its coordination from tetrahedral to octahedral by interacting with lone pairs of water is concurrent with the definition of Lewis acidity. The y-intercept of this correlation, which refers to the Lewis acidity in the absence of any octahedral aluminum, is likely to originate from other Lewis-acidic aluminum species, such as distorted four-coordinated and/or penta-coordinated aluminum. The broad shoulder of the peak at 58 ppm in the 27Al MAS NMR

It is vital to understand that the structure probed by FTIR spectroscopy is not the same as that in 27Al MAS NMR as the two techniques differ in their pretreatments of the zeolite sample and these varying treatments bring about a change in aluminum coordination.18, 19, 23, 25, 41, 42 While 27Al MAS NMR is measured on samples stored under ambient conditions, FTIR spectroscopy entails the degassing of the zeolite under vacuum. We demonstrate this difference in sample pre-treatment by mimicking the wet environs of an NMR experiment in an FTIR spectroscopic measurement using carbon monoxide as the probe molecule. Upon pre-adsorbing very small quantities of water (20 µmolg-1, approximately 10% of total Lewis acid content) before exposure to carbon monoxide, we observe a decrease in the absorbance band at 2223 cm-1 and 2187 cm-1, associated with threecoordinated Lewis acidic aluminum and Brønsted acid sites respectively. (Figure S4). This shows the tendency of water molecules to adsorb over Lewis and Brønsted acid sites and induce a change in coordination of aluminum that can be regarded as associated to the framework. As has been reported in the past, moisture drives the formation of octahedrally coordinated aluminum in H-mordenite and this aluminum species is unstable at temperatures higher than 395 K, quantitatively reverting back to tetrahedral coordination.21 The other takeaway from Figure 4 is the influence of Si/Al ratio on Lewis acidity. The higher Lewis acidity in samples with greater aluminum content (Table 1) underscores the importance of having framework aluminum close to each other to generate sites with Lewis acidity. Furthermore, our experiments highlight the necessity to have protons as charge-compensating cations in order to generate Lewis-acidic aluminum sites. When examined by FTIR spectroscopy using carbon monoxide as the probe molecule, the MOR zeolite in its sodium form did not show any signature for Lewis acidic aluminum sites (Figure S5). As expected, this sample also had very little to no octahedral aluminum under the conditions of an NMR experiment (Figure S3). The simultaneous absence of Lewis acidity and octahedral aluminum in hydrated state reaffirms the relationship of six-coordinated aluminum to Lewis acidity.

ACS Paragon Plus Environment

Page 4 of 7

Page 5 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Subsequently, treating Na-MOR(9) in nitric acid solution results in the generation of Lewis acidity (absorbance at 2223 cm1, Figure S4) and a population of aluminum species with octahedral coordination (Acid treated MOR(9), Figure S6) Thus, the octahedrally coordinated aluminum species is only present in the case of proton-exchanged zeolites, suggesting that two close Brønsted acid sites are needed to generate the defective site able to change the aluminum coordination into octahedral. Besides, this observation is in good agreement with in situ XAS experiments that ascribe the formation of ocatahedral aluminum, without loss of aluminum from the zeolite lattice, to the interaction of water with Brønsted acid sites.40 Conclusions To summarize, the experiments in this study constitute a prominent step forward in understanding the structure of aluminum-based Lewis acidity in zeolites. Aluminum species in octahedral coordination under hydrated conditions are associated with Lewis acidity in the dehydrated environment of FTIR spectroscopy. These Lewis acid sites only form in protonic zeolites that have close aluminum sites and such a Lewis acidity is intrinsic to the zeolite framework. The structure of Lewis acid sites in zeolites is a function of conditions and the octahedral aluminum describing this Lewis acidity is associated to the framework. In addition, as reflected by the offset in the FTIRNMR correlation, there do exist other aluminum species in the zeolite with Lewis acidic properties. In the light of our findings, octahedral aluminum that is associated to the framework can be associated to Lewis acidity without ambivalence and the existence of such sites is strictly influenced by the Si/Al ratio of the zeolite. We propose that this correlation of octahedral frameworkassociated aluminum to Lewis acidity holds true for zeolite topolgies beyond just mordenite. Gil et al. studied the acidic properties of SSZ-35 and SSZ-33 using infrared spectroscopy and MAS NMR.34 NH4-SSZ-35 was found to possess two tetrahedral aluminum sites and no octahedral aluminum. Meanwhile, the proton form of SSZ-33 exhibited a sharp resonance at around 0 ppm in the 27Al MAS NMR and this octahedral aluminum species was shown to adopt a tetrahedral coordination on exchange with ammonium nitrate. The Lewis acidity per unit cell of this sample was far greater than that of SSZ-35 and we postulate that this difference in Lewis acidity could be explained by the vastly different amounts of octahedral aluminum in the two samples. AUTHOR INFORMATION

Corresponding Author Prof. Dr. Jeroen A. van Bokhoven E-mail: [email protected]

Author Contributions MR synthesized the samples, performed and analyzed the NMR experiments and prepared the manuscript. VS performed and analyzed the FTIR experiments and provided feedback during the preparation of the manuscript. JAvB devised the overall idea, supervised the experimental progress and provided feedback during the preparation of the manuscript.

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT The authors gratefully acknowledge the ESI Platform, Paul Scherrer Institute and ETH Zurich for financial support and the assistance of Dr. René Verel for NMR measurements. Manoj

thanks Teng Li and Amy Knorpp for help with AAS measurements.

ASSOCIATED CONTENT Supporting Information Available Figures showing curated data for correlation of Lewis acidity with extra-framework aluminum, nitrogen adsorption-desorption isotherms, the reversible octahedral-tetrahedral aluminum transformation in zeolite MOR, effect of pre-adsorbed water on Lewis and Brønsted acidity, FTIR spectra of adsorbed carbon monoxide and 27Al MAS NMR spectra of Na-MOR(9) and acid treated MOR(9) are included in the supporting information. The Supporting Information is available free of charge via the Internet at http://pubs.acs.org

REFERENCES 1. Marcilly, C. Present Status and Future Trends in Catalysis for Refining and Petrochemicals. J. Catal. 2003, 216, 47-62. 2. Degnan Jr, T. F.; Smith, C. M.; Venkat, C. R. Alkylation of Aromatics with Ethylene and Propylene: Recent Developments in Commercial Processes. Appl. Catal. A: Gen. 2001, 221, 283-294. 3. Sartori, G.; Maggi, R. Use of Solid Catalysts in Friedel− Crafts Acylation Reactions. Chem. Rev. 2006, 106, 1077-1104. 4. Van Bekkum, H.; Kouwenhoven, H. W. Zeolites and Fine Chemicals. Stud. Surf. Sci. Catal. 1988, 41,45-59. 5. Clerici, M. G. Zeolites for Fine Chemicals Production. Top. Catal. 2000, 13, 373-386. 6. Boronat, M.; Corma, A.; Renz, M.; Viruela, P. M. Predicting the Activity of Single Isolated Lewis Acid Sites in Solid Catalysts. Chem. Eur. J. 2006, 12, 7067-7077. 7. Lewis, J. D.; Van de Vyver, S.; Román‐Leshkov, Y. Acid–base Pairs in Lewis Acidic Zeolites Promote Direct Aldol Reactions by Soft Enolization. Angew. Chem. Int. Ed. 2015, 54, 9835-9838. 8. Corma, A.; Nemeth, L. T.; Renz, M.; Valencia, S. Sn-zeolite Beta as a Heterogeneous Chemoselective Catalyst for Baeyer–Villiger Oxidations. Nature 2001, 412, 423-425. 9. Luo, H. Y.; Lewis, J. D.; Román-Leshkov, Y. Lewis Acid Zeolites for Biomass Conversion: Perspectives and Challenges on Reactivity, Synthesis, and Stability. Annu. Rev. Chem. Biomol. Eng. 2016, 7, 663-692. 10. Taarning, E.; Osmundsen, C. M.; Yang, X.; Voss, B.; Andersen, S. I.; Christensen, C. H. Zeolite-catalyzed Biomass Conversion to Fuels and Chemicals. Energy Environ. Sci. 2011, 4, 793-804. 11. Lippmaa, E.; Samoson, A.; Magi, M. High-Resolution Aluminum-27 NMR of Aluminosilicates. J. Am. Chem. Soc. 1986, 108, 1730-1735. 12. Derouane, E. G.; Vedrine, J. C.; Pinto, R. R.; Borges, P. M.; Costa, L.; Lemos, M.; Lemos, F.; Ribeiro, F. R. The Acidity of Zeolites: Concepts, Measurements and Relation to Catalysis: a Review on Experimental and Theoretical Methods for the Study of Zeolite Acidity. Catal. Rev. 2013, 55, 454-515. 13. Freude, D.; Hunger, M.; Pfeifer, H.; Schwieger, W. 1H MAS NMR Studies on the Acidity of Zeolites. Chem. Phys. Lett. 1986, 128, 6266. 14. Hunger, M.; Freude, D.; Pfeifer, H. Magic-Angle Spinning Nuclear Magnetic Resonance Studies of Water Molecules Adsorbed on Brønsted-and Lewis-Acid Sites in Zeolites and Amorphous Silica– Aluminas. J. Chem. Soc., Faraday Trans. 1991, 87, 657-662. 15. Hadjiivanov, K. I.; Vayssilov, G. N. Characterization of Oxide Surfaces and Zeolites by Carbon Monoxide as an IR Probe Molecule. Adv. Catal. 2002, 47, 307-511. 16. Hidalgo, C. V.; Itoh, H.; Hattori, T.; Niwa, M.; Murakami, Y. Measurement of the Acidity of Various Zeolites by TemperatureProgrammed Desorption of Ammonia. J. Catal. 1984, 85, 362-369. 17. Maier, S. M.; Jentys, A.; Lercher, J. A. Steaming of Zeolite BEA and its Effect on Acidity: a Comparative NMR and IR Spectroscopic Study. J. Phys. Chem. C 2011, 115, 8005-8013. 18. Bourgeat-Lami, E.; Massiani, P.; Di Renzo, F.; Espiau, P.;

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fajula, F.; Des Courières, T. Study of the State of Aluminium in Zeolite-β. Appl. Catal. 1991, 72, 139-152. 19. van Bokhoven, J. A.; Sambe, H.; Ramaker, D. E.; Koningsberger, D. C. Al K-Edge Near-Edge X-Ray Absorption Fine Structure (NEXAFS) Study on the Coordination Structure of Aluminum in Minerals and Y Zeolites. J. Phys. Chem. B 1999, 103, 7557-7564. 20. Omegna, A.; van Bokhoven, J. A.; Prins, R. Flexible Aluminum Coordination in Alumino− Silicates. Structure of Zeolite H− USY and Amorphous Silica− Alumina. J. Phys. Chem. B 2003, 107, 88548860. 21. van Bokhoven, J. A.; Van der Eerden, A. M. J.; Koningsberger, D. C. Three-Coordinate Aluminum in Zeolites Observed with in situ XRay Absorption Near-Edge Spectroscopy at the Al K-Edge: Flexibility of Aluminum Coordinations in Zeolites. J. Am. Chem. Soc. 2003, 125, 74357442. 22. Abraham, A.; Lee, S.-H.; Shin, C.-H.; Hong, S. B.; Prins, R.; van Bokhoven, J. A. Influence of Framework Silicon to Aluminium Ratio on Aluminium Coordination and Distribution in Zeolite Beta Investigated by 27 Al MAS and 27 Al MQ MAS NMR. Phys. Chem. Chem. Phys. 2004, 6, 3031-3036. 23. Kunkeler, P. J.; Zuurdeeg, B. J.; Van der Waal, J. C.; van Bokhoven, J. A.; Koningsberger, D. C.; Van Bekkum, H. Zeolite Beta: the Relationship Between Calcination Procedure, Aluminum Configuration, and Lewis Acidity. J. Catal. 1998, 180, 234-244. 24. Woolery, G. L.; Kuehl, G. H.; Timken, H. C.; Chester, A. W.; Vartuli, J. C. On the Nature of Framework Brønsted and Lewis Acid Sites in ZSM-5. Zeolites 1997, 19, 288-296. 25. Wouters, B. H.; Chen, T. H.; Grobet, P. J. Reversible Tetrahedral− Octahedral Framework Aluminum Transformation in Zeolite Y. J. Am. Chem. Soc. 1998, 120, 11419-11425. 26. Kuehl, G. H.; Timken, H. K. C. Acid sites in Zeolite Beta: Effects of Ammonium Exchange and Steaming. Microporous Mesoporous Mater. 2000, 35, 521-532. 27. Creyghton, E. J.; Ganeshie, S. D.; Downing, R. S.; Van Bekkum, H. Stereoselective Meerwein–Ponndorf–Verley and Oppenauer Reactions Catalysed by Zeolite BEA1. J. Mol. Catal. A: Chem. 1997, 115, 457-472. 28. Jia, C.; Massiani, P.; Barthomeuf, D. Characterization by Infrared and Nuclear Magnetic Resonance Spectroscopies of Calcined Beta Zeolite. J. Chem. Soc., Faraday Trans. 1993, 89, 3659-3665. 29. Beck, L. W.; Haw, J. F. Multinuclear NMR Studies Reveal a Complex Acid Function for Zeolite Beta. J. Phys. Chem. 1995, 99, 10761079. 30. Minachev, K. M.; Garanin, V. I.; Kharlamov, V. V.; Isakova, T. A.; Senderov, E. E. Catalytic Properties of Synthetic Mordenite in the Isomerization, Hydrogenation, and Hydroisomerization of Certain Hydrocarbons. Russ. Chem. Bull. 1969, 18, 1611-1615. 31. Groothaert, M. H.; Smeets, P. J.; Sels, B. F.; Jacobs, P. A.; Schoonheydt, R. A. Selective Oxidation of Methane by the Bis (μ-oxo) Dicopper Core Stabilized on ZSM-5 and Mordenite Zeolites. J. Am. Chem. Soc. 2005, 127, 1394-1395. 32. Rahkamaa-Tolonen, K.; Maunula, T.; Lomma, M.; Huuhtanen, M.; Keiski, R. L. The Effect of NO2 on the Activity of Fresh and Aged Zeolite Catalysts in the NH3-SCR Reaction. Catal. Today 2005, 100, 217222. 33. Haouas, M.; Kogelbauer, A.; Prins, R. The Effect of Flexible Lattice Aluminium in Zeolite Beta During the Nitration of Toluene with Nitric Acid and Acetic Anhydride. Catal. Lett. 2000, 70, 61-65. 34. Gil, B.; Zones, S. I.; Hwang, S.-J.; Bejblová, M.; Čejka, J. Acidic Properties of SSZ-33 and SSZ-35 Novel Zeolites: a Complex Infrared and MAS NMR Study. J. Phys. Chem. C 2008, 112, 2997-3007. 35. Oumi, Y.; Nemoto, S.; Nawata, S.; Fukushima, T.; Teranishi, T.; Sano, T. Effect of the Framework Structure on the Dealumination– Realumination Behavior of Zeolite. Mater. Chem, Phys. 2003, 78, 551557. 36. Pappas, D. K.; Martini, A.; Dyballa, M.; Kvande, K.; Teketel, S.; Lomachenko, K. A.; Baran, R.; Glatzel, P.; Arstad, B.; Berlier, G. et al. The Nuclearity of the Active Site for Methane to Methanol Conversion in Cu-Mordenite: a quantitative assessment. J. Am. Chem. Soc. 2018, 140, 15270-15278. 37. Emeis, C. A. Determination of Integrated Molar Extinction Coefficients for Infrared Absorption Bands of Pyridine Adsorbed on Solid Acid Catalysts. J. Catal. 1993, 141, 347-354. 38. Maache, M.; Janin, A.; Lavalley, J. C.; Joly, J. F.; Benazzi, E. Acidity of Zeolites Beta Dealuminated by Acid Leaching: An FTi. r.

Study Using Different Probe Molecules (Pyridine, Carbon Monoxide). Zeolites 1993, 13, 419-426. 39. De Ménorval, L. C.; Buckermann, W.; Figueras, F.; Fajula, F. Influence of Adsorbed Molecules on the Configuration Of Framework Aluminum Atoms in Acidic Zeolite-β. A 27Al MAS NMR Study. J. Phys. Chem. 1996, 100, 465-467. 40. Drake, I. J.; Zhang, Y.; Gilles, M. K.; Teris Liu, C. N.; Nachimuthu, P.; Perera, R. C. C.; Wakita, H.; Bell, A. T. An in situ Al KEdge XAS Investigation of the Local Environment of H+-and Cu+Exchanged USY and ZSM-5 Zeolites. J. Phys. Chem. B 2006, 110, 1166511676. 41. van Bokhoven, J. A.; Koningsberger, D. C.; Kunkeler, P.; Van Bekkum, H. Influence of Steam Activation on Pore Structure and Acidity of Zeolite Beta: An Al K Edge XANES Study of Aluminum Coordination. J. Catal. 2002, 211, 540-547. 42. Jiao, J.; Altwasser, S.; Wang, W.; Weitkamp, J.; Hunger, M. State of Aluminum in Dealuminated, Nonhydrated Zeolites Y Investigated by Multinuclear Solid-State NMR Spectroscopy. J. Phys. Chem. B 2004, 108, 14305-14310.

ACS Paragon Plus Environment

Page 6 of 7

Page 7 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

TOC Image

ACS Paragon Plus Environment

7