Delicate Distinction between OH Groups on Proton-Exchanged H

Sep 27, 2015 - Research and Development Center, PQ Corporation, 280 Cedar Grove Road, Conshohocken, Pennsylvania 19428, United States. ‡ Research an...
1 downloads 6 Views 2MB Size
Subscriber access provided by UNIV OF LETHBRIDGE

Article

Delicate Distinction Between OH Groups on Proton Exchanged H-Chabazite and H-SAPO-34 Molecular Sieves Istvan Halasz, Bjorn Moden, Anton Petushkov, Jian-Jie Liang, and Mukesh Agarwal J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.5b09247 • Publication Date (Web): 27 Sep 2015 Downloaded from http://pubs.acs.org on October 7, 2015

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 free 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 accessible to all readers and 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.

The Journal of Physical Chemistry C 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 32

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

1

Delicate Distinction Between OH Groups on Proton Exchanged H-Chabazite and H-SAPO-34 Molecular Sieves

Istvan Halasz1*, Bjorn Moden2, Anton Petushkov2 Jian-Jie Liang3, Mukesh Agarwal1 1

Research and Development Center, PQ Corporation, 280 Cedar Grove Road, Conshohocken,

PA 19428 (USA); [email protected] 2

Research and Development Center, Zeolyst International, 280 Cedar Grove Road,

Conshohocken, PA 19428 (USA) 3

BIOVIA, Dassault Systèmes, 5005 Wateridge Vista Dr., San Diego, CA 92121(USA)

*Corresponding author

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

2

ABSTRACT We observed surprising difference in the FTIR (Fourier transform infrared) hydroxyl spectra of the structurally isomorphous, proton exchanged H-CHA and H-SAPO-34 molecular sieves when measured by transmission (TR) or diffuse reflectance (DRIFT) techniques. Experimental and density functional theory (DFT) based model evidence is presented in this paper to prove that the essential reason for this spectral difference is that DRIFT emphasizes the vibrations of surface hydroxyl sites. Vibrations of the bulk Brønsted acidic hydroxyls shift to higher frequencies when they become surface species, the IR beam is reflected from approximately the top ~15 to 20 Å thick layer of the particles, hence the proportion of surface related IR bands becomes significant compared to the bulk related ones in the DRIFT spectra while the opposite is valid for the TR spectra. We demonstrate that the surface hydroxyls are Brønsted acidic both on the HCHA and the H-SAPO-34 particles and the upshifted vibrations noticed primarily in the DRIFT spectra are Al-OH vibrations on the surface even of H-SAPO-34, not P-OH groups as most researchers believe. We also show that the bulk Brønsted sites might involve HO1, HO2 and HO4 type hydroxyls associated with the known geometrically different oxygen positions on both molecular sieves, but only HO1 surface hydroxyls are associated with the red-shifted vibration intensified in the DRIFT spectra. Moreover, a single surface model cannot account for every vibration observed in DRIFT spectra. From the combination of IR vibrations of three adequate surface models one can as properly match the experimental DRIFT spectra as the TR spectra from the combination of the calculated bulk HO1…HO4 vibrations of these molsieve crystals.

Keywords: FTIR, DRIFT, molecular sieve, zeolite, hydroxyl, Brønsted acidity, surface sites, DFT, computer model

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

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

3

1. INTRODUCTION Chabazite (CHA), an aluminosilicate zeolite, also frequently referred as SSZ-131, and its crystallographically isomorphous silicoaluminophosphate, SAPO-34, are well studied molecular sieves with approximately d ~ 3.8 Å and ~ 4.3 Å diameter microchannels2-4, respectively. These materials are important ingredients of catalysts for the recently developed advanced methanol-tohydrocarbon (MTH) technology5,6 and the urea-SCR (Selective Catalytic Reduction) based NOx reduction of diesel engine exhausts7, 8. The Brønsted acidic hydroxyl groups (BA-OH) of both zeolites have pivotal catalytic role in these processes; hence their accurate characterization is a regular task. Along with all other hydroxyls, this characterization is most frequently made by Fourier Transform Infrared Spectroscopy (FTIR). Depending on the analysts’ preference, two major FTIR sampling techniques are in use: i) transmission (TR), mostly with thin, selfsupported pellets; ii) diffuse reflectance (DR), carried out with powdered samples9. Results are generally considered to be identical with each other, especially when the DRIFT (diffuse reflectance infrared Fourier transform) results are modified by Kubelka-Munk or other functions to compensate for non-linearity9-17. Therefore, it came as a total surprise for us to see that the hydroxyl FTIR spectra of proton exchanged H-CHA and H-SAPO-34 samples gave quite different spectra when we incidentally measured their OH contents by both DRIFT and TR technique18. As an example, illustrated in Fig. 1a, a band 3677 cm-1 is much more intense in the DRIFT spectrum of HSAPO-34 than in its TR spectrum. In a very similar manner, Fig 1b shows that the 3662 cm-1 band is much more intense in the DRIFT spectrum of H-CHA than in its TR spectrum. Since such comparative FTIR measurements with these two techniques have not been reported on these samples, we surveyed the literature for a potential explanation of this unusual phenomenon.

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

Page 4 of 32

4

Fig. 1 The FTIR spectra of a) a H-SAPO-34 sample with nominal composition of Al5.9P5.0 Si1.1H1.1O24 and b) a H-CHA sample with nominal composition of Si5.6Al0.4H0.4O12

show

substantial difference when measured by DRIFT vs. transmission (TR) technique. The 3600/3627 cm-1 bands of H-SAPO-34 and the 3594/3612 cm-1 bands of H-CHA have been univocally associated in the literature with their Brønsted acidic Si-OH and Al-OH sites, respectively. It is clear from the crystal geometry that these low and high frequency (LF and HF) vibrational band pairs might actually arise from four different Brønsted protons connected to four, geometrically differently positioned oxygen atoms when they surround an Al3+ ion, isomorphously substituting a tetrahedral Si4+ ion in the aluminosilicate CHA. These oxygens are indicated in Fig. 2, which shows the unit cell of a SSZ-13 crystal and allows visualizing the CHA structure as interconnected double-six member siloxane rings (D6R). The geometry of H-SAPO34 is identical but its tetrahedral building blocks are alternating [PO4]+ and [AlO4]- units, from which some of the P5+ ions are isomorphously substituted by Si4+ ions which generates the Brønsted acidity of this crystal. There is no total agreement between theorists, which of the four possible hydroxyls is responsible for the experimentally observed FTIR vibrations in these two

ACS Paragon Plus Environment

Page 5 of 32

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

5

Fig. 2 The unit cell of purely siliceous H-SSZ-13 structure shows two interconnected D6R units; its four geometrically distinct oxygen positions are marked (every [SiO4] tetrahedral unit is equivalent in this crystal lattice). Color codes: Si = Yellow; O = Red. molecular sieves. For example Shah et al.19 have found that the HO1 and HO3 groups are energetically most stable on both SSZ-13 and H-SAPO-34, but their computed anharmonic νOH vibrations, which are actually seen in the experimental FTIR spectra, do not fit accurately the experimental values. Similar results were reported by Mihaleva et al.20 for H-CHA, based on 8T ring and 7T cluster calculations. Smith et al.21, 22 on the other hand computed with high accuracy that HO2 and HO4 Brønsted acidic sites fit the experimental spectra in both zeolites, but also observed shifting the HO4 proton to HO1 position. Lo and Trout23 essentially also supports this result.

Using QM-Pot method Sierka and Sauer24 even found that the protons of the [AlO4]

units in various zeolites jump from position to position but on the H-CHA they found that the HO1 and HO2 positions (see Fig. 2) are preferred. Jeanvoine et al.25, 26 also suggest that the H-

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

Page 6 of 32

6

CHA acidic bands are associated with HO1 and HO2 proton positions and in H-SAPO-34 that HO4 protons might contribute to them, as well. Hemelsoet et al.27 also identified the HO4 proton being responsible for the HF vibration of H-SAPO-34. Miki Niwa’s group28 computed that the HO3 site of H-CHA cannot be part of the experimental FTIR spectrum of Brønsted acidic sites since vibrates at lower energies where the experimental spectra are largely empty. They have found that the computed vibrations of the other three hydroxyls fit the experimental spectra. This group28,

29

also proved that in general the phosphorous containing zeolites are

weaker Brønsted acids than their aluminosilicate structure-isomers.

The plane wave based

calculations of Sauer et al.30 also resulted in this conclusion for the H-SAPO-34 and H-CHA structures. The interpretation of the respective 3677 cm-1 and 3662 cm-1 bands of H-SAPO-34 and H-CHA is more controversial than that of the named Brønsted acidic sites. They do not show up in the computer models of these molecular sieves, i.e., they seem not to be associated with the structural Brønsted acidic sites. In the course of our literature survey we have found that those researchers who use DRIFT technique31-40 generally find these bands more intense than those who use TR technique41-54. In case of the H-SAPO-34 most researchers assign the band near ~3680 cm-1 to surface P-OH vibration, following Peri’s55 early empirical assignment from the TR spectra of an amorphous AlPO4. Yet, his spectra also showed the parallel appearance surface Al-OH vibrations near 3800 cm-1, which are absent from our, and many other authors’ H-SAPO34 spectra (Fig. 1a). The P-OH assignment certainly cannot be valid for the ~3662 cm-1 band of the H-CHA. To our best knowledge only Bordiga et al.56 assigned this band to “extra lattice or partially extra lattice Al-OH vibrations”. Many researchers associate the appearance of both the ~3680 cm-1 H-SAPO-34 and ~3660 cm-1 H-CHA FTIR bands with sample hydrolysis, showing

ACS Paragon Plus Environment

Page 7 of 32

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

7

that the intensity of these bands substantially increases in the presence of H2O41-45,56,57. It has been also reported that the acidic OH groups of both molecular sieves tend to connect at least one, but rather 3 or 4 H2O molecules via hydrogen bonds34,58,59. However, our measurements were carried out on samples calcined at 500 oC or higher temperatures in air and in addition they were also in situ evacuated at ~10-5 torr at 500 oC before both the TR and the DRIFT measurements, which were carried out at room temperature afterward.

This pretreatment

removes all molecular water from the zeolites as one could confirm from the lack of any measurable 1640 cm-1 FTIR band, associated with the well-known bending vibration of molecular water. It follows that hydrolysis could not take place during our FTIR measurements. What is more, none of these band assignments could answer the question why can be the hydroxyl spectra substantially differ when measured by TR or DRIFT technique? To clarify this issue, we carried out a number of experiments and DFT (Density Functional Theory) based model calculations. We show in this paper that the distinct sensitivity of the two FTIR sampling techniques toward the surface and bulk hydroxyl groups is the key factor, which is not uniquely restricted to the H-CHA and H-SAPO-34 structures. In the course of this work we also identified and corrected some misconceptions about published vibrational peak assignments and associated every hydroxyl band visible in the TR and the DRIFT spectra with surface and bulk OH species. 2. EXPERIMENTAL AND COMPUTATIONAL METHODS 2.1. Sample sources and preparation Every molecular sieve tested was either a commercial product of Zeolyst International (www.zeolyst.com) or its slightly modified version to compare the effect of crystal size and/or

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

8

elemental composition. Spectroscopic quality pyridine (PY) was purchased from Aldrich. FTIR quality KBr was obtained from the International Crystal Labs. Accurate refractive index liquids were purchased from the Cargille Labs. 2. 2. Analytical techniques 2.2.1. FTIR (Fourier Transform Infra-Red) measurements Transmisson (TR) FTIR analyses were carried out on a Nicolet 6700 spectrometer from Thermo Scientific, by placing self-supported material disks into a sample holder from CIC Photonics, Inc., evacuating it at 500 oC and 5 x 10-6 mbar for at least 2 hours and obtaining the hydroxyl spectra after cooling them under vacuum to room temperature (estimated 23±2 °C). DRIFT measurements were made on a Bruker IFS/66 spectrometer. 10% sample powder was thoroughly mixed and grinded with KBr and the mixture was placed either into the sample holder of a Praying Mantis unit from Harrick Scientific, Inc., or into the sample holder of a Diffuse IRTM unit from PIKE Technologies. The sample pretreatment was identical with that of the TR technique.

All DRIFT spectra were also measured at room temperature and were

converted into Kubelka-Munk (KM) units. For pyridine (PY) adsorption measurements, samples were pretreated the same way as for the OH-measurements, but cooled only to 150 oC. Liquid PY was placed into a quartz container and kept under vacuum to avoid moisture inlet. Before adsorption measurements, the liquid was frozen with liquid N2 while the space above the frozen liquid and the overall inlet line was thoroughly evacuated. After this the liquid N2 was removed and the PY was allowed to warm up to room temperature. Next, a Swagelok valve, which separated the PY line from the evacuated sample, was opened and the system was let to equilibrate. After equilibration the Swagelok valve was closed, the sample chamber was evacuated and the sample was cooled from 150 oC to

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32

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

9

room temperature under vacuum for FTIR measurement. For desorption at higher temperatures, the sample was heated up still under vacuum and kept at the elevated temperature for 2 hours. For measurement, it was cooled down again to room temperature.

2.2.2. Refractory index measurements The refractory index of molecular sieve powders was measured by the Becke line method60. We checked the halo around the particles in a drop of calibrated refractive index liquid by using a Leitz Wetzlar Transmission Phase Contrast microscope. The correct index was gained at the borderline when, upon raising the focus, the higher and lower refractive index liquids caused to move the halo inward or outward, respectively. We used liquid sets differing by n = 0.004 refractivity index. Hence, this determined the accuracy of measurements.

2.3. Molecular Simulation Density functional theory61 implemented in the program CASTEP62 was used to obtain equilibrium crystal structures.

Each structure was geometry optimized using the Broyden,

Fletcher, Goldfarb, and Shannon (BFGS) minimizer. Norm-conserving potentials were used. For H, O, Si, Al, and P the valence electrons included were 1s1; 2s2 2p4; 3s2 3p2; 3s2 3p1; and 3s2 3p3, respectively. The plane wave basis set cutoff was 750 eV. The k-point grid was kept to maintain a spacing of ca. 0.08 Å-1. For generalized gradient approximation (GGA) the functional of Perdew, Burke and Ernzerhof (PBE)63 was employed. The convergence criteria for total energy, max force, max stress, max displacement and SCF (Self Consistent Field) iterations were 2 x e−5 eV/atom, 0.05 eV/Å, 0.1GPa, 2 x e−3 Å and 2 x e-6 eV/atom, respectively. To simulate the FTIR spectra of periodic crystals, we started with these geometry optimized structures and computed the phonon density of states/vibrational spectra with

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

Page 10 of 32

10

CASTEP’s implementation of density functional perturbation theory (DFPT)64. A k-spacing of 0.0286 Å-1 was employed in sampling the vibrational density of states with sufficient resolution. The convergence criteria for SCF iterations of the ground state electronic structure was set to 5 x e-11 eV/atom or lower, to ensure convergence of the DFPT wavefunctions. To see the deviation of computed bulk spectra from the spectral vibrations of surface species, we calculated the IR spectra of periodic vacuum slabs created from the minimized molsieve structures. For charge equilibration of the dangling bonds at the discontinued crystal surface we created various bond “healings” (interconnecting [TO4] tetrahedral units via oxygen bridges; T could be Al, Si, or P atom depending on the material at hand), probed the effect of increased coordination from tetrahedral to higher values, and capped dangling oxygen atoms with protons. These surface modifications were carried out so that the overall vacuum slab became electrically neutral.

Due to these changes, each vacuum slab was individually

minimized again at the same way as the bulk crystal was using the CASTEP program and thereafter the DFPT feature of this program was deployed to compute the vibrational spectra from the phonon density of states. The computed FTIR vibrations are illustrated by using Lorentzian line shapes.

To

approximate the line width of the experimental spectra we adjusted them to their width at half height, usually 20 cm-1 to 50 cm-1 depending on the spectrum at hand. We also used GRAMS32 program from Thermo Scientific to combine the adequately increased or decreased computed surface spectra to match the DRIFT spectra.

ACS Paragon Plus Environment

Page 11 of 32

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

11

RESULTS AND DISCUSSION After seeing repeated results of different DRIFT vs. TR spectra with variously composed H-CHA and H-SAPO-34 samples similar to shown in Fig. 1, we investigated the hydroxyl FTIR spectra on a number of other zeolites and amorphous silicas using both DRIFT and TR techniques. In Fig. 3a the spectra of a commercial, ultra-stabilized (US-Y) zeolite, CBV 500, exemplifies that the differently measured spectra are indeed roughly identical with each other on numerous samples as most researchers expect. Slight differences in the intensity ratios of the various bands are not unusual, but the peak positions are the same. On the other hand, the example of Ferrierite (FER) in Fig. 3b illustrates that the H-CHA and H-SAPO-34 are not alone with the position differences in peak maxima, which makes the differently measured spectra quite different. It is noteworthy to mention that always some higher wavenumber bands get more intense in the DRIFT measurement. Since DRIFT is a reflection technique, with which the vibrational spectra are known to be affected by the particle size, we tested zeolites of various

Fig. 3 Comparison of the FTIR spectra of a) the commercial, proton-exchanged, H-Y zeolite, CBV 500 and b) the smaller and larger crystallite versions of the commercial, proton-exchanged Ferrierite (H-FER), CP 914C, zeolites measured by DRIFT and transmission (TR) technique.

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

Page 12 of 32

12

crystallite sizes. Fig. 3b illustrates that this parameter has not been found to have as significant effect on the overall appearance of spectra as the switch between the DRIFT and the TR techniques. Another factor which can affect the reflection of material particles is their refractory index (n). For various zeolites and amorphous silica samples we found only small variation in this parameter, within the range from n = 1.440±0.004 to n = 1.452±0.004. The differences did not correlate with the appearance or lack of peak shifts between the TR and DRIFT spectra. To see if the most intense bands in the DRIFT spectra of H-SAPO-34 and H-CHA are Brønsted acidic or not, we carried out ion exchange with alkaline and other cations. As an example with sodium exchange illustrates in Fig. 4, every hydroxyl band got substantially reduced in both materials, no matter which sampling technique was used. Consequently every hydroxyl has exchangeable, Brønsted acidic character, although those OH-groups which are represented by the 3677 cm-1 and 3670 cm-1 bands in the DRIFT spectra appear to be less prone

Fig. 4 Effect of Na-exchange on the hydroxyl FTIR bands of a) a H-SAPO-34 sample with nominal composition of Al3P1.8Si1.2H1.2O12 and b) a H-CHA sample with nominal composition of Si2.2Al0.8H0.8O12..

ACS Paragon Plus Environment

Page 13 of 32

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

13

for the exchange, i.e., probably less acidic than the 3600/3627 cm-1 and 3617/3643 cm-1 band pair represented Brønsted acidic sites in these materials, as discussed in the Introduction. Note that the composition of H-CHA in Fig. 4 is somewhat different from that of the high Si/Al ratio H-CHA in Fig. 1, which caused some peak shifts, but the difference of its DRIFT and TR spectra remained valid. The bands near 3740 cm-1 are well-known as usually non-acidic or very weakly acidic, isolated terminal silanol groups, hence the ion exchange might or might not affect them. Thus, we had to conclude that neither the chemical composition, nor the particle size or the refractory index could cause the appearance of intense bands of acidic hydroxyls which appear in the DRIFT spectra at higher frequencies than the structural Brønsted acidic hydroxyl bands. Therefore, we speculated that these bands might arise from hydroxyls on the particle surfaces magnified by the reflection technique. To check the viability of this hypothesis we carried out adsorption measurements with pyridine (PY) on these molecular sieves.

This

molecule is frequently used to measure selectively the Brønsted acidic sites by FTIR spectroscopy on various oxides since it gives a characteristic, distinctly measurable peak at 1540 cm-1. However, its kinetic diameter is about 5.85 Å, which does not allow it to enter the narrow channels of H-CHA and H-SAPO-34. Thus, any measurable PY adsorption must occur on the surface of these molecular sieves. As the example of H-CHA illustrates in Fig. 5, pyridine virtually did not affect the Brønsted acidic hydroxyl bands in the transmission cell, which shows mainly the bulk acidic sites. However, a dramatic drop in the hydroxyl intensity occurred when the same adsorption was carried out in the DRIFT cell. Consequently the DRIFT spectrum mainly shows surface Brønsted acidic sites. Note that the strongest drop occurred with the 3668 cm-1 band while the lower wavelength bands remained partly unaffected, presumably because they represent “bulk” acidic sites so deep from the surface, to where the PY molecules could not

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

Page 14 of 32

14

Fig. 5 The FTIR spectrum of hydroxyl groups on a H-CHA a) remains largely unchanged when pyridine (PY) is adsorbed onto the sample in a transmission (TR) sample holder at 150 oC, but b) substantially drops when this adsorption is done in a DRIFT cell and recovers upon desorbing the PY at higher temperatures. penetrate. Notwithstanding, the intensity drop in this hydroxyl vibration range indicates that these bands mainly represent OH groups so close to the surface that the PY could interact with them, i.e., presumably not further than one or two atomic layers (vide infra at surface models). At elevated temperatures up to about 400 oC, the adsorbed PY molecules can be desorbed and the original spectrum recovered. We got exactly the same results on an H-SAPO-34 sample as well. This strong experimental proof for the preferred surface characterization of DRIFT spectra tempted us to learn more about the specificities of these surface hydroxyls, since the facilitated accessibility of external Brønsted acidic sites compared to those, which are in the micropores of these molecular sieves, might bear even of catalytic importance65. For this end we carried out DFT based model calculations for the IR spectra of bulk and surface hydroxyls both on H-CHA and H-SAPO-34. First, we tested the accuracy of our modeling parameters by

ACS Paragon Plus Environment

Page 15 of 32

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

15

computing the bulk Brønsted acidic OH-vibrations associated with the isomorphously substituted Al atoms in H-CHA and Si atoms in H-SAPO-34. We considered all four geometrically possible Brønsted sites, indicated in Fig. 2, for both structures. To accelerate the calculations, they were made on the primitive cells of these crystals instead of their rhombohedral unit cell structures22, 66 shown in Fig. 2. With several parallel calculations we confirmed that the computed vibrational results with the primitive cells are totally identical with those obtained by computing the larger unit cell. Actually, these primitive cells correspond to the trigonal unit cells identified for mineral chabazites long time ago2, 3. One Al and one Si atom was inserted into the model primitive cells of H-CHA and H-SAPO-34, respectively. Fig. 6 illustrates these structures showing HO1 and HO2 positioned Brønsted sites. Their nominal compositions are close to the composition of the experimental samples in Fig. 1, thus our computed models were compared to these spectra.

Fig. 6 The primitive cell of a) a H-SAPO-34 structure with nominal composition of Al6P5SiHO24; its extralattice proton is connected to an O1 atom; and b) a H-CHA sample with nominal composition of Si11AlHO24; its extra lattice proton is connected to an O2 atom (see the numbered O-atom positions in Fig 2). Color codes: white = H; red = O; pink = Al; yellow = Si; green = 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

Page 16 of 32

16

As Fig. 7 indicates, the computed spectra gave an excellent fit to the experimental TR FTIR bands:

in both structures the O1 positioned hydroxyl is responsible for the higher

frequency (wavenumber) bands while the O2 and O4 positioned hydroxyls overlap each other at the lower frequency band position. The ∆E (eV) energy differences for the HO1, HO2, HO3, and HO4 bonds, which characterize their stabilities, were found to be 0, 0.03, 0.37, 0.08 for HCHA and 0, 0.05, 0.46, 0.08 for H-SAPO-34, respectively. These values are close to published data except for HO3 which we found much less stable than previously reported19,

26

. The

structures we used for calculation can be seen in the Supplementary Information as .cif files which allow to see every detail of the lattice parameters and atom positions. As mentioned in the caption of Fig. 7, the computed vibrations of H-CHA had to be shifted up by a factor of 1.0033. This shift corresponds to some calculations which also used plane wave approach19, 26. Note however that we are not using the customary notation of scaling factor for computed frequencies, as a scaling factor is typically based on a full range (or a very limited number of sub-ranges) of vibrational frequencies of a target system67. In the present work, we focus on just the O-H stretching frequencies. The somewhat differing amounts of shifts, zero for H-SAPO-34 and 0.33% for H-CHA, can originate from difference in electron correlation and/or anharmonicity of the vibrations in the target systems67, given identical basis set treatments in both calculations. Moreover, the difference is small enough to be comparable to an rms of difference between computed and experimental frequencies in the same target system67, not to mention across different systems. None of the computed vibrations were related to the 3677 cm-1 and 3662 cm-1 vibrations observed in the DRIFT spectra of H-SAPO-34 and H-CHA, respectively (Fig. 1). The most confined O3 related hydroxyls vibrate at much lower frequencies, near 3200 cm-1, where

ACS Paragon Plus Environment

Page 17 of 32

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

17

Fig. 7 Experimental TR (copied from Fig. 1) and computed OH spectra of H-SAPO-34 and HCHA with different HOn positions where n = the oxygen position numbers in Fig 2. The model spectra for H-SAPO-34 are as computed and for H-CHA they are shifted up by 0.33%. experimental FTIR bands were not observed Consequently HO3 hydroxyls are only theoretical possibilities, but are not present in the real H-SAPO-34 and H-CHA structures. This result fits for example the result reported by Suzuki et al.28, although their computed HO3 vibration was significantly higher, 3538 cm-1, than ours. For better visibility the computed HO3 bands were omitted from Fig. 7, but we show them in Fig. S1 in the Supporting Information section, along with the detail model structures. As the selected computational parameters gave very reliable vibrational spectra for the bulk hydroxyls, we also used them to compute the surface specific hydroxyls. For this we created two to five primitive cell deep surface slabs. This allowed us to see how far from the surface vibrate the bulk Brønsted acidic sites unbiased and also to control if the calculation went in order, i.e., the bulk Brønsted sites gave the same spectra as seen from the periodic bulk calculation in Fig. 7. We created slabs with various hkl index faces; at the top, at 2/3rd and at one half of the top primitive cell; and with putting D6R to the slab surface.

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

Page 18 of 32

18

Moreover, we tested all four possible hydroxyl positions for each slab variation and closed the inevitable dangling bonds at the surface discontinuation of the periodic lattice in ways which are widely considered to be realistic in the chemistry of metal oxides. These included bridging of oxygens between different tetrahedral units, capping the singular oxygens with hydrogens, making surface oxygens double bounded to the surface cation, or capping the incompletely coordinated top Al, Si, or P atoms with hydroxyl groups. We accepted the surface rearrangement only when the whole periodic slab became electronically neutral. Because of these inevitable distortions on the surface, every surface model had to be geometry optimized and energy minimized before IR model calculations could be carried out. In the course of these virtual experiments we learned several things: i) The Brønsted acidic sites vibrated totally undisturbed and gave the same spectra as indicated in Fig. 7, when they were only in the second or third primitive cell, i.e. quite close to the surface. Considering that the thickness of one cell is about 9.4 Å, their estimated distance from the surface was around 15 to 25 Å depending on how deep the first primitive cell was cut. In accordance with our deductions from the PY adsorption experiments, this implies that the surface sites become dominating in the DRIFT spectra only if the IR beam in the DRIFT cell does not penetrate deeper than one to three primitive cells.

ii) Very few surface models resulted in vibrations near

the 3677 cm-1 and 3662 cm-1 positions, for which we conducted these calculations. Exclusively HO1 vibrations on surface cuts from 100 and -100 directions gave intense bands near these positions without also generating other bands at positions where the experimental spectra showed nothing. iii) In case of H-CHA, the HO1 vibration was close to the 3662 cm-1 target whether it appeared on the surface as Si-connected (3663 cm-1) or Al-connected (3651 cm-1) group. In case of H-SAPO-34 the computed 3678 cm-1 band only appeared when the surface HO1 was

ACS Paragon Plus Environment

Page 19 of 32

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

19

connected to Al atoms. iv) Since the experimental DRIFT spectra did not indicate any bands near or above 3800 cm-1 it is unlikely that any significant AlxOy(OH)z type surface accumulation could occur68-70.

In line with this, the cross-section test of our H-CHA and H-SAPO-34

crystallites by EDAX showed totally homogeneous Al-dispersion without any surface accumulation. v) In contrast to the common assignments, these higher frequency bands could never be associated with surface P-OH vibrations, which always gave vibrations below 3570 cm1

.

vi) We tested the option of “healing” the surfaces by using higher than tetrahedral

coordinations for Al and P which is known stable state for them in many non-zeolytic oxides and hydroxides. These models resulted in IR spectra quite different from the experimental ones. vii) None of the computed surfaces could account alone for every IR vibration observed in the experimental DRIFT spectra. The main reason for this is that on a surface, constructed from periodic repetition of a truncated unit cell one can only test one type of Brønsted acidic hydroxyls at a time; let’s say HO2 connected to either its Al-O or Si-O neighbor in a CHA unit cell. It is also limited how many types of other surface hydroxyls can be modeled within a cell, let’s say by capping dangling P-O or Al-O oxygens by H away from the Brønsted acidic sites in a SAPO-34 structure. Same is valid for the interconnecting oxygens between different [TO4] tetrahedra on the surface, where T represents Si, Al, or P atoms. We have found however three surface slab results for both the H-CHA and the H-SAPO-34, which could be combined to mimic quite nicely the experimental DRIFT spectra when the peak intensities were adequately adjusted, as Fig. 8 illustrates. Fig. 9 shows two-two surfaces for both molecular sieves which accounted for most of their model vibrations. The top figure in Fig. 9/a shows a H-SAPO-34 surface with HO1 sites connected to Al and Si atoms resulting in model IR vibrations at 3679 cm-1 and 3721 cm-1, respectively. On the bottom H-SAPO-34 surface in Fig 9/a the Al-HO1 bond resulted in

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

Page 20 of 32

20

3674 cm-1 computed vibration and the calculated P-OH related vibration was 3559 cm-1. Both calculations gave back the expected vibration near 3627 cm-1 associated with the bulk HO1 Brønsted acidic site. A third surface slab with HO2 Brønsted sites was needed to generate the band at 3588 cm-1, which combined with the other two spectra gave exactly the well-known 3600 cm-1 peak illustrated in Fig. 8. This third surface also resulted in a 3745 cm-1 vibration of Al connected HO2 on the surface, also indicated in Fig. 8. In a very similar manner the reader can make the comparison between the H-CHA related surface and bulk species shown in Fig. 9/b and the computed model spectrum in Fig. 8. Details of the surface slab models and their computed IR spectra are shown in the Supplemental Information section.

Fig. 8 Experimental DRIFT (copied from Fig. 1) and computed OH spectra of surface slabs of H-SAPO-34 and H-CHA. The HO1 and HO2 positions are illustrated in Fig 2. Below the peak position and O assignment it is indicated if the vibration belongs to surface or bulk OH groups.

ACS Paragon Plus Environment

Page 21 of 32

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

21

Fig. 9 Surface models for a) H-SAPO-34 and b) H-CHA which gave most band for their combined spectra in Fig. 8. Color codes are the same as in Fig. 6. Finally, we wish to address a paradox phenomenon: namely that the intensity of external isolated silanol groups vibrating near 3740 cm-1 is usually much lower in the DRIFT spectra than the intensity of the 3662 cm-1 and 3678 cm-1 terminal Brønsted acidic sites. Note that compared to the internal Brønsted site intensities (near 3615 cm-1 and 3630 cm-1) the intensity of the 3740 cm-1 band is about the same both in the DRIFT and the TR spectra. This suggests that the much higher intensity of the surface Brønsted sites in the DRIFT might be much more abundant on the crystallite surfaces than the non- acidic, isolated, terminal Si-OH. Notwithstanding, in lack of any relevant study, one cannot exclude either that DRIFT has an intrinsic quantitation problem.

For example, using reflection Raman measurements for

measuring the intensity of Si-O vibrations, it was found that even a 100-fold intensity difference can occur when the vibration belongs to a [SiO4] unit connected to three other neighbors versus the Si-O vibration of a terminal [SiO4] connected to only one other neighbor71.

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

Page 22 of 32

22

CONCLUSIONS In this study we showed that the TR and DRIFT sampling techniques can result in quite different FTIR spectra. We show experimental and DFT based model evidence for H-CHA and H-SAPO-34 molecular sieves that the reason for the differences is that DRIFT emphasizes the vibrations of surface OH groups when such groups are present on the surface of crystallites. We demonstrated that the surface hydroxyls of these molecular sieves are Brønsted acidic and in contrast to the common belief they represent surface Al-OH and not P-OH groups.

The

computer models also showed that almost exclusively HO1 hydroxyls appear on the surface and a single surface model cannot account for every vibration observed in DRIFT spectra. We managed to get an excellent fit, however, when combined tree adequate surface slab results for both H-SAPO-34 and H-CHA. The presented experimental and computer modeling studies allow a deeper understanding of fine details in the internal and external molecular structure of these practically important molecular sieves. We believe that the presented data adequately prove that the DRIFT technique represents the constitution of a thin surface layer of solid particles in general, which might or might not differ from the bulk composition.

ACKNOWLEDGEMENT The authors appreciate support and permission for publication from Zeolyst International and PQ Corporation. We thank Ms. Larissa Ding the electron microscopic and EDAX analysis of our samples.

ACS Paragon Plus Environment

Page 23 of 32

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

23

REFERENCES (1) Zones, S. I. Zeolite SSZ-13 And Its Method of Preparation. US Patent 4,544,538, 1985. (2) Dent, L. S., Smith, J. V. Crystal Structure of Chabazite, a Molecular Sieve. Nature, 1958 181, 1794-1796. (3) Smith, J. V.; Rinaldi, F.; Dent Glasser, L. S. Crystal Structures With a Chabazite Framework. II. Hydrated Ca-Chabazite at Room Temperature. Acta Crystallogr. 1963, 16, 45-53. (4) Lok, B. M.; Messina, C. A.; Patton, R. L.; Gajek, R. T.; Cannan, T. R.; Flanigen, E. M. Silicoaluminophosphate Molecular Sieves: Another New Class of Microporous Crystalline Inorganic Solids. J. Am. Chem. Soc. 1984, 106, 6092-6093. (5) Vora, B.; Chen, J. Q.; Bozzano, A.; Glover, B.; Barger, P. Various Routes to Methane Utilization – SAPO-34 Catalysis Offers The Best Option. Catalysis Today 2009, 141, 77-83. (6) Galadima, A.; Muraza, O. Recent Developments on Silicoaluminates and Silicoaluminophosphates in the Methanol-to-Propylene Reaction: A Mini Review. Industrial & Engineering Chemistry Research 2015, 54, 4891-4905. (7) Schmieg, S. J.; Oh, S. H.; Kim, C. H.; Brown, D. B.; Lee, J. H.; Peden, C. H. F.; Kim, D. H. Thermal Durability of Cu-CHA NH3-SCR Catalysts for Diesel NOx Reduction. Catalysis Today 2012, 184, 252-261. (8) Ma, L.; Cheng, Y.; Cavataio, G.; McCabe, R. W.; Fu, L.; Li, J. Characterization of Commercial Cu-SSZ-13 and Cu-SAPO-34 Catalysts with Hydrothermal Treatment for NH3SCR of NOx in Diesel Exhaust. Chem. Engineering Journal 2013, 225, 323-330.

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

Page 24 of 32

24

(9) Karge, H. K.; Geidel, E. Vibrational Spectroscopy. In Molecular Sieves; Science and Technology; Karge, H. G., Weitkamp, J., Eds.; Springer, Berlin; 2004, 4, 1-200. (10) Kortüm, G. Reflectance Spectroscopy; Springer, New York; 1969. (11)

Fuller, M. P.; Griffiths, P. R. Diffuse Reflectance measurement by Infrared Fourier

Transform Spectrometry. Analytical Chemistry 1978, 50, 1906-1910. (12) Kubelkova, L.; Hoser, H.; Riva, A.; Trifiro, F. Infrared Diffuse Reflectance measurement of Zeolites and Adsorbed Species. Zeolites 1983, 3, 244-248. (13) Leyden, D. E.; Murthy, R. S. S. Diffuse Reflectance Fourier Transform IR Spectroscopy. Trends in Analytical Chemistry 1988, 7, 164-169. (14) Loeffler, E.; Peuker, Ch.; Zibrowius, B.; Zscherpel, U.; Schnabel, K. H. Application of Diffuse Reflectance FTIR Spectroscopy in Molecular Sieve Research. Proc. 9th International Zeolite Conference, Montreal 1992.; von Ballmoos, R., Higgins, J. B., Treacy, M. M. J., Eds.; Butterworth-Heinemann, Boston; 1993, 1, 521-528. (15) Milosevic, M.; Berets, S. L. A Review of FT-IR Diffuse Reflection Sampling Considerations. Applied Spectroscopy Reviews 2002, 37, 347-364. (16) Sirita, J.; Phanichpant, S.; Meunier, F. C., Quantitative Analysis of Adsorbate Concentrations by Diffuse Reflectance FT-IR. Anal. Chem. 2007, 79, 3912-3918. (17) Reeves, III, J. B. Does the Spectral Format Matter in Diffuse Reflection Spectroscopy? Applied Spectroscopy 2009, 63, 669-677.

ACS Paragon Plus Environment

Page 25 of 32

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

25

(18) Halasz, I.; Moden, B.; Petushkov, A.; Agarwal, M.; Liang, J. J. FTIR Distinction of Internal and External Brønsted Acidic Sites on Zeolite Particles. Proc. 23rd North American Catalysis Society Meeting, Louisville, KY; 2013, O-TH-JON-1. (19) Shah, R.; Gale, J. D.; Payne, M. C. The Active Sites of Microporous Solid Acid Catalysts. Phase Transitions 1997, 61, 67-81. (20) Mihaleva, V. V.; van Santen, R. A.; Jansen, A. P. The Heterogeneity of the Hydroxyl Groups in Chabasite. J. Chem. Phys. 2003, 119, 13053-13060. (21) Smith, L.; Cheetham, A. K.; Marchese, L.; Thomas, J. M.; Wright, P. A.; Chen, J. A Quantitative Description of the Active Sites in the Dehydrated Acid Catalyst HSAPO-34 for the Conversion of Methanol to Olefins. Catal. Letters 1996, 41, 13-16. (22) Smith, L. J.; Davidson, A.; Cheetham, A. K. A Neutron Difraction and Infrared Spectroscopy Study of the Acid Form of the Aluminosilicate Zeolite, Chabazite (H-SSZ-13). Catal. Lett. 1997, 49, 143-146. (23) Lo, C.; Trout, B. L. Density-Functional Theory Caracterization of Solid Acid Sites in Chabasite. J. Catal. 2004, 227, 77-89. (24) Sierka, M.; Sauer, J. Proton Mobility in Chabazite, Faujasite, and ZSM-5 Zeolite Catalysts. Comparison Based on ab Initio Calculations. J. Phys. Chem. B 2001, 105, 1603-1613. (25) Jeanvoine, Y.; Angyan, J. G.; Kresse, G.; Hafner, J. On the Nature of Water Interacting with Brønsted Acidic Sites. Ab Initio Molecular Dynamics Study of Hydrated H-SAPO-34. Phys. Chem. B 1998, 102, 7307-7310. (26) Jeanvoine, Y.; Angyan, J. G.; Kresse, G.; Hafner, J. Brønsted Acid Sites in H-SAPO-34 and Chabazite: An An Initio Structural Study. J. Phys. Chem. B 1998, 102, 5573-5580.

ACS Paragon Plus Environment

J.

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

26

(27) Hemelsoet, K.; Ghysels, A.; Mores, D.; De Wispelaere, K.; Van Speybroeck, V.; Weckhuysen, B. M.; Waroquier, M. Experimental and Theoretical IR Study of Methanol and Ethanol Conversion Over H-SAPO-34. Catalysis Today 2011, 177, 12-24. (28) Suzuki, K; Sastre, G.; Katada, N.; Niwa, M. Ammonia IRMS-TPD Measurements and DFT Calculation on Acidic Hydroxyl Groups in CHA-Type Zeolites. Phys. Chem. Chem. Phys. 2007, 9, 5980–5987 (29) Katada, N.; Kazuma Nouno, K.; Lee, J. K.; Jiho Shin, J.; Suk Bong Hong, S. B.; Niwa, M. Acidic Properties of Cage-Based, Small-Pore Zeolites with Different Framework Topologies and Their Silicoaluminophosphate Analogues. J. Phys. Chem. C 2011, 115, 22505–22513(30) Sauer, J.;Schroder, K-P.; Termath, V. Comparing the Acidities of Microporous Aluminosilicate and Silico-Aluminophosphate Catalysts: a Combined Quantum Mechanics-Interatomic Potential Funktion Study. Collect. Czech. Chem. Commun. 1998, 63 1394-1408. (31) Löffler, E.; Peuker, C.; Finger, G.; Girnus, I.; Jahn, E.; Zubowa, H-L. IR-spectroskopishe Untersuchung der Hydroxylgruppen in Molekularsieben des SAPO-Type. Z. Chem. 1990, 30, 255-256.

(32) Zubkov, S. A.; Kustov, L. M.; Kazansky, V. B. Investigation od Hydroxyl Groups in Crystalline Silicoaluminophosphate SAPO-34 by Diffuse Reflectance Infrared Spectroscopy. J. Chem. Soc. Faraday Trans. 1991, 87, 897-900. (33) Zibrowius, B.; Löffler, E.; Hunger, M. Multinuclear MAS NMR and IR Spectroscopic Study of Silicon Incorporation into SAPO-5, SAPO-31, and SAPO-34 Molecular Sieves. Zeolites 1992, 12, 167-174. (34) Marchese, L.; Chen, J.; Wright, P. A.; Thomas, J. M. Formation of H3O+ at the Brönsted Site in SAPO-34 Catalysts. J. Phys. Chem. 1993, 97, 8109-8112.

ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32

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

27

(35) Campelo, J. M.; Lafont, F.; Marinas, J. M.; Ojeda, M. Studies of Catalyst Deactivation in Methanol Conversion with High, Medium and Small Pore Silicoaluminophosphates. Appl. Catalysis A: General 2000, 192, 85-96. (36) Itakura, M.; Goto, I.; Takahashi, A.; Fujitani, T.; Ide, Y.; Sadakane, M. Synthesis of HighSilica CHA Type Zeolite by Interzeolite Conversion of FAU Type Zeolite in the Presence of Seed Crystals. Microporous Mesoporous Materials 2011, 144, 91-96. (37) Wang, L.; Li, W.; Qi, G.; Weng, D. Location and Nature of Cu Species in Cu/SAPO-34 for Selective Catalytic Reduction of NO with NH3. J. Catal. 2012, 289, 21-29. (38) Wang, D.; Zhang, L.; Kamasamudram, K.; Epling, W. S. In Siru-DRIFTS Study of Selective Catalytic Reduction of NOx by NH3 over Cu-Exchanged SAPO-34. ACS Catal. 2013, 3, 871-881. (39) Sedighi, M.; Bahrami, H.; Darian, J. T. Thorough Investigation of Varying template Combinations on SAPO-34 Synthesis, Catalytic Activity and Stability in the Methanol Conversion to Light Olefin. RSC Advances 2014, 4, 49762-49769. (40) Duan, Y.; Wang, J.; Yu, T.; Shen, M.; Wang, J. The Role and Activity of Various Adsorbed Ammonia Species on Cu/SAPO-34 Catalyst During Passive-SCR Process. RSC Advances 2015, 5, 14103-14113. (41) Mortier, W. J.; King, G. S. D.; Sengler, L. Crystal Structures of Dehydrated H Chabazite Pretreated at 320 oC, and at 600 oC After Steaming. J. Phys. Chem. 1979, 83, 2263-2266. (42) Beyer, H. K.; Belenkaja, I. M.; Borbely, G.; Tielen, M.; Jacobs, P. A. Deammoniation and Dehydroxylation of Calcium Ammonium Chabazites. J. Chem. Soc. Faraday Trans. I 1985, 81, 3049-3058.

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

28

(43) Smith, L.; Cheetam, A. K.; Morris, R. E.; Marchese, L.; Thomas, J. M.; Wright, P. A.; Chen, J. On the Nature of Water Bound to a Solid Acid Catalysts. Science 1996, 271, 799-802. (44) Zhu, Q.; Kondo, J. N.; Ohnuma, R.; Kubota, Y.; Yamaguchi, M.; Tatsumi, T. The Study of Methanol-To-Olefin Over Proton Type Aluminosilicate CHA Zeolites. Microporous Mesoporous Materials 2008, 112, 153-161. (45) Zhu, Q.; Hinode, M.; Yokoi, T.; Yoshioka, M.; Kondo, J. N.; Tatsumi, T. Synthesis, Characterization and Catalytic Studies of CHA Zeotype Materials Containing Boron and Iron. Catalysis Communications 2009, 10, 447-450. (46) Makarova, M. A.; Ojo, A. F.; Karim, K.; Hunger, M.; Dwyer, J. FTIR Study of Weak Hydrogen Bonding of Brønsted Hydroxyls in Zeolites and Aluminophosphates. J. Phys. Chem. 1994, 98, 3619-3623. (47) Chen, J.; Wright, P. A.; Thomas, J. M.; Natarajan, S.; Marchese, L.; Bradley, S. M.; Sankar, G.; Catlow, C. R. SAPO-18 Catalysts and Their Brønsted Acid Sites. J. Phys. Chem. 1994, 98, 10216-10224. (48) Chen, J.; Thomas, J. M.; Wright, P. A.; Townsend, R. P. Silicoaluminophosphate Number Eighteen (SAPO-18): a New Microporous Solid Acid Catalyst. Catal. Lett. 1994, 28, 241-248. (49) Marchese, L.; Frache, A.; Gatti, G.; Coluccia, S.; Lisi, L.; Guoppolo, G.; Russo, G.; Pastore, H. O. Acid SAPO-34 Catalyst for Oxidative Dehydrogenation of Ethane. J. Catal. 2002, 208, 479-484. (50) Tan, J.; Liu, Z.; Bao, X.; Liu, X; Han, X; He, C.; Zhai, R. Crystallization and Si Incorporation Mechanisms of SAPO-34. Microporous Mesoporous Materials 2002, 53, 97-108.

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32

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

29

(51) Frache, A., Gianotti, E., Marchese, L., Spectroscopic CXharacterization of Microporous Aluminophosphate Materials with Potential Application in Environmental Catalysis. Catalysis Today 2003, 77, 371-384. (52) Martins, G. A. V.; Berlier, G.; Coluccia, S.; Pastore, H. O.; Superti, G. B.; Gatti, G.; Marchese, L. Revisiting the Nature of the Acidity in Chabasite-Related Silicoaluminophosphates: Combined FTIR and 29Si MAS NMR Study. J. Phys. Chem. C 2007, 111, 330-339. (53) Izadbakhsh, A.; Farhadi, F.; Khorasheh, F.; Sahebdelfar, S.; Asadi, M.; Feng, Y. Z. Effect of SAPO-34’s Composition on its Phisico-Chemical Properties and Deactivation in MTO Process. Appl. Catal. A: General 2009, 364, 48-56. (54) Jang, H-G.; Min, H-K.; Lee, J. K.; Hong, S. B.; Seo, G. SAPO-34 and ZSM-5 Nanocrystals‘ Size Effect on Their Catalysis of Methanol-To-Olefin Reactions. Appl. Catal. A: General 2012, 437-438, 120-130. (55) Peri, J. B. Surface Chemistry of AlPO4 – A Mixed Oxide of Al and P. Disc. Faraday Trans. Soc. 1971, 55-65. (56) Bordiga, S.; Regli, L.; Cocina, D.; Lamberti, C.; Bjørgen, M.; Lillerud, K. P. Assessing the Acidity of High Silica Chabazite H-SSZ-13 by FTIR Using CO as Molecular Probe: Comparison with H-SAPO-34. J. Phys. Chem. B 2005, 109, 2779-2784. (57) Bordiga, S.; Regli, L.; Lamberti, C.; Zecchina, A. FTIR Adsorption Studies of H2O and CH3OH in the Isostructural H-SSZ-13 and H-SAPO-34: Formation of H-Bonded Adducts and Protonated Clusters. J. Phys. Chem. B 2005, 109, 7724-7732. (58) Termath, V.; Haase, F.; Sauer, J.; Hutter, J.; Parrinello, M. Undertsanding the nature of Water Bound to Solid Acid Surfaces. Ab Initio Simulation on H-SAPO-34. J. Am. Chem. Soc. 1998, 120, 8512-8516.

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

30

(59) Vener, M. V.; Rozanska, X.; Sauer, J. Protonation of Water Clusters in the Cavities of Acidic Zeolites: (H2O)n. H-Chabazite, n = 1-4. Phys. Chem. Chem. Phys. 2009, 11, 1702-1712. (60) Allen, R. M. Practical Refractometry by Means of the Microscope. Publ. by R. P. Cargille Laboratories 1985, pg. 6-7. (61) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. B 1964, 136, 864−871. (62) Clark, S. J.; Segall, M. D.; Pickard, C. J.; Hasnip, P. J.; Probert, M. J.; Refson, K.; Payne, M. C. First Principles Methods Using CASTEP. Z. Kristallogr. 2005, 220 (5−6), 567−570. (63) Perdew, J. B.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (64) Baroni, S.; de Gironcoli, S.; dal Corso, A.; Giannozzi, P. Phonons and Related Crystal Properties from Density-Functional Perturbation Theory. Rev. Mod. Phys. 2001, 73, 515−562. (65) Liu, D.; Zhang, X.; Bhan, A.; Tsapatsis, M. Activity and Selectivity Differences of External Brønsted Acid Sites of Single-Unit-Cell Thick and Conventional MFI and MWW Zeolites. Microporous Mesoporous Materials 2014, 200, 287-290. (66) Dias-Cabanas, M-J; Barrett, P. A.; Camblor, M. A. Synthesis and Structure of Pure SiO2 Chabazite: the SiO2 Polymorph with the Lowest Framework Density. Chem. Commun. 1998, 1881-1882. (67) Scott, A. P.; Radom, L. Harmonic Vibrational Frequencies: An Evaluation of Hartree-Fock, Møller-Plesset, Quadratic Configuration Interaction, Density Functional Theory, and Semiempirical Scale Factors. J. Phys. Chem. 1996, 100, 16502-16513. (68) Peri, J. B. A Model for the Surface of γ-Alumina. J. Phys. Chem. 1965, 69, 220-230. (69) Peri, J. B. Infrared Study of Adsorption of Ammonia on Dry γ-Alumina. J. Phys. Chem. 1965, 69, 231-239.

ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32

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

31

(70) Knozinger, H.; Ratnasamy, P. Catalytic Aluminas: Surface Models and Characterization of Surface Sites. Catal. Rev.-Sci. Eng. 1978, 17, 31-70. (71) Halasz, I.; Kierys, A.; Goworek, J.; Liu, H.; Patterson, R. E. 29Si NMR and Raman Glimpses into the Molecular Structures of Acid and Base Set Silica Gels Obtained from TEOS and Na-Silicate. J. Phys. Chem. C 2011, 115, 24788–24799.

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

32

TOC picture

ACS Paragon Plus Environment

Page 32 of 32