Synthesis and Growth Kinetics of Zeolite SSZ-39 - Industrial


Synthesis and Growth Kinetics of Zeolite SSZ-39 - Industrial...

0 downloads 70 Views 5MB Size

Article pubs.acs.org/IECR

Synthesis and Growth Kinetics of Zeolite SSZ-39 Ross Ransom,‡ Jonathan Coote,‡ Roger Moulton,† Feng Gao,† and Daniel F. Shantz*,‡ †

SACHEM, Inc., 821 East Woodward St., Austin, Texas 78704, United States Department of Chemical and Biomolecular Engineering, Tulane University, 6823 St. Charles Avenue, New Orleans, Louisiana 70118, United States



ABSTRACT: A thorough study of SSZ-39 formation, a next generation deNOx catalyst, is presented. The presence of the trans isomer is beneficial to the growth kinetics leading to enhancements in the growth rate, in some cases of over 40%. The formation of SSZ-39 is also sensitive to the composition of the faujasite used as an aluminum source as gels with identical compositions but different faujasites do not lead to SSZ-39 formation. Once SSZ-39 begins to form, its growth rate is linear and appears to equal the rate of faujasite dissolution. Finally, the Si/Al ratio of the material is influenced by the cis/trans ratio of the SDA. These results provide new insights into the formation of this industrially relevant catalyst.



analogue of the AEI (AlPO4-18) topology.13,17−25 This zeolite has been synthesized using both quaternary ammonium compounds21−23 and phosphonium17,18,20 compounds. Moreover, this material shows promise for deNOx catalysis,18,21 as well as methane to methanol26 and methanol to olefins.22 Thus, a simple and scalable synthesis is something that would be highly desirable, as would detailed information about growth kinetics and synthesis yields and how they depend on synthesis conditions. A variety of organic SDAs have been shown to afford SSZ-39 at various levels of efficacy, including N,Ndimethyl-3,5-dimethylpiperidinium, the organic SDA employed here. Prior work from the Davis lab in collaboration with SACHEM scientists has suggested that the isomer identity has a significant effect on growth kinetics.23 With the above in mind, the current work set out to quantify how the isomer identity impacts crystallization kinetics and to understand how the composition of the parent faujasite used as the aluminum source impacted the synthesis kinetics and yield and how wide of a gel composition space could be used to make phase pure SSZ-39. Experimental Section. Materials. Sodium silicate (PQ Brand N, 28.9 wt % SiO2, 8.9 wt % Na2O) was kindly provided by PQ Corporation. Faujasite (FAU) zeolites of various compositions were obtained from Alfa Aesar (manufactured by Zeolyst). N,N-Dimethyl-3,5-dimethylpiperidinium hydroxide and chloride (various cis/trans ratios) were provided by SACHEM, Inc. Sodium hydroxide (ACS reagent grade) was purchased from VWR. All reagents were used as received. Zeolite Synthesis. The general procedure used for the synthesis of SSZ-39 using N,N-dimethyl-3,5-dimethylpiperidi-

INTRODUCTION Zeolites are a technologically relevant class of microporous crystalline aluminosilicates.1−3 They are used in industrial processes ranging from fluidized catalytic cracking to aromatics upgrading, i.e., BTX interconversion, to light gas separations.4 This is a consequence of their crystalline structures that have uniform size pores in the 0.5−1 nm range. This coupled with the ability to introduce acidity via substitution of aluminum for silicon has led to widespread application of zeolites in the refining and chemical industries. Clear structure−property relationships can be deduced that relate the zeolite topology to the property of interest, e.g., reactivity. Given this, there is and will continue to be strong interest in the synthesis of new zeolites.5−7 Zeolites are typically synthesized hydrothermally in alkaline media.8 Zeolites are synthesized in either the presence of large quantities of alkali cations (sodium, potassium) in the case of low-silica zeolites or in the presence of organic quaternary ammonium compounds in the case of high-silica zeolites. These organic molecules are often referred to as structure-directing agents (SDA) as there is often a clear correlation between the identity of the organic molecule and the zeolite structure obtained.9,10 These synthesis mixtures typically contain a silica source, an aluminum source, alkali, SDA, and water. Often the aluminum sources used are molecular salts such as aluminum sulfate or nitrate. Another route to introducing aluminum into a synthesis gel is to use a zeolite as the aluminum source. While this has long been known to be a viable route to making highsilica zeolites,11−13 the use of zeolites as reagents, i.e., zeolite interconversion, has seen a significant recent increase in use in the open literature.14−16 In some cases, it clearly enables the synthesis of a zeolite in a gel composition space that would be unsuccessful using molecular aluminum salts. One zeolite that has been shown to be facile to make using faujasite as an aluminum source is SSZ-39, the aluminosilicate © XXXX American Chemical Society

Received: Revised: Accepted: Published: A

February 13, 2017 March 15, 2017 March 24, 2017 March 24, 2017 DOI: 10.1021/acs.iecr.7b00629 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

weighed after drying divided by the mass of faujasite added plus the mass of silicon from the sodium silicate in the form of SiO2. Finally, note that in the figures that each point shown represents an individual synthesis; i.e., the syntheses were performed in duplicate/quadruplicate to estimate variability. Analytical. Powder X-ray diffraction (PXRD) measurements were performed with a Rigaku MiniFlex 600 diffractometer with Cu Kα radiation over a range of 4−50° 2θ. Thermal gravimetric analyses (TGA) were performed using a Q500 instrument from TA Instruments over a temperature range of 25−600 °C using pure air as a carrier gas and a temperature ramping rate of 1 °C/min. Field emission scanning electron microscopy (FE-SEM) measurements were performed with a Hitachi 4800 high-resolution scanning electron microscope operating at 5.0 kV. Zeolite compositions were determined by Galbraith.

nium hydroxide as the SDA was based on a prior report by Dusselier and SACHEM scientists (Scheme 1).23 As a Scheme 1. Isomers of SDA Used in This Work

representative example is the following made with a gel composition of 1 SiO2:0.033 Al:0.57 NaOH:0.14 R+OH−:28 H2O. First, 4.33 g of sodium silicate were measured out gravimetrically and then placed in a Teflon jar on a stir plate with a magnetic stir bar. To this was added 0.5 mL of 1 M NaOH solution. To this solution was added 2.5 mL of 20 wt % N,N-dimethyl-3,5-dimethylpiperidinium hydroxide, followed by the addition of 6.2 mL of DI water. Here, 0.17 g faujasite (SiO2:Al2O3 = 5.4, Zeolyst CBV500) was then weighed out and added. This mixture was stirred at room temperature for 2 h, then transferred to two 23 mL Teflon-lined autoclaves and placed in the oven for the indicated amount of time at 140 °C in a rack rotating at approximately 60 rpm. The autoclaves were then removed from the oven and allowed to cool to room temperature. The solids were then collected via centrifugation, reserving the mother liquor, and washing and resuspending in DI water three times to bring the pH to approximately 7. The centrifuge tubes containing the wet solids were then placed in the oven at 100 °C overnight until dry. SDA Ion Exchange. Approximately 10 mmol of the SDA chloride (77.5% trans isomer) compound was dissolved in 25 mL of DI water. This was added to an exchange column filled with approximately 200 mL of Dowex Marathon A (hydroxide form) exchange resin and allowed to sit for 1 h. The column was then washed with three additions of 100 mL DI water. The resulting approximately 325 mL of solution was then concentrated to between 0.5 to 1.0 M using a rotary evaporator. The final concentration was confirmed by titration with 0.1 M HCl into a mixture of 0.5 mL of the SDA hydroxide solution and 50 mL of DI water, using phenol red as an indicator. Assessment of Phase Purity, Yield Definition. Here, we use the following criterion to describe the extent of SSZ-39 crystallization. The ratio shown in many figures is the ratio of the SSZ-39 peak intensity at 9.5° 2θ divided by the FAU peak intensity at 6.1° 2θ plus the SSZ-39 peak intensity at 9.5° 2θ or IAEI,9.5◦2θ



RESULTS AND DISCUSSION Prior work by Dusselier and co-workers23 was used as a starting point in this investigation to develop a more quantitative understanding of how the synthesis gel composition and SDA isomer identity impacted growth kinetics. Figure 1 shows a

Figure 1. Representative powder X-ray diffraction pattern for 1 SiO2:0.067 Al:0.14 R+OH−:0.51 NaOH:28 H2O after 3 days at 413 K.

representative PXRD pattern for a SSZ-39 synthesis of composition 1 SiO2:0.067 Al:0.14 R+OH−:0.51 NaOH:28 H2O heated for 3 days at 413 K. Per prior work, a small amount of analcime (ANA) is formed in these preparations. Figure 2 shows representative FE-SEM images for the parent faujasite zeolite used as well as the final SSZ-39 product. As shown, the faujasite (FAU) and SSZ-39 both have different particle sizes as well as morphologies. FE-SEM images of samples determined to be pure SSZ-39 by PXRD exhibit only one particle morphology which is that shown in Figure 2. Growth Kinetics. Figure 3 shows the rate of SSZ-39 formation estimated using PXRD and the solids yield versus time for 1 SiO2:0.067 Al:0.14 R+OH−:0.51 NaOH:28 H2O made with the pure cis isomer. The rate of SSZ-39 appearance is approximately linear, and there is minimal scatter in the data (four syntheses for each time point are shown) in both the

(IAEI,9.5◦2θ + IFAU,6.1◦2θ)

This ratio is bound between 0 (no AEI) and 1 (all AEI) and thus is a convenient, albeit qualitative, indication of the SSZ-39 crystallization extent. To assess the accuracy of this metric, physical mixtures of phase pure FAU and SSZ-39 were mixed in known mass fractions and analyzed using powder X-ray diffraction (PXRD). Those experiments (not shown) indicate this metric accurately captures the relative amount of the two phases to ±5%. No appreciable amorphous material (broad feature between 20−30° 2θ) was observed for any of the samples made here. Also, in many figures, a percent yields solid is shown. This is the amount of solid recovered from synthesis B

DOI: 10.1021/acs.iecr.7b00629 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

24 h until the completion of the synthesis at 3 days the solids yields vary only slightly. Only one of the 24 points is outside of the range of 34−39% solids yield, and that point is at 1.5 days corresponding to the low outlying XRD point. These two points (linear growth and effectively constant solids yields) can most easily be explained by the conclusion that the rate of SSZ39 formation is equal to the rate of FAU dissolution. The total absence of amorphous materials by PXRD is also consistent with this picture. In contrast, many syntheses with aluminum salts exhibit sigmoidal shaped growth curves. Along this theme of FAU dissolution controlling zeolite formation, if one assumes that the aluminum from the parent FAU is totally consumed and thus determines the SSZ-39 yield at a 40% solids yield, the Si/Al ratio should be approximately 6.4. This is in fact close to the value reported by Dusselier and co-workers (6.2).23 Thus, in all of these syntheses, the yield is below 50%; that is a result of the fact that the final SSZ-39 formed has a lower Si/Al ratio than the gel. This is observed in all cases (vide inf ra). On the basis of the results above, consistent with prior work, the SSZ-39 crystallization using a gel Si/Al = 15 is complete in 3 days and contains a small amount of analcime, and interestingly, the solids yield varies only slightly after 24 h into the synthesis. This issue of solids yields is described in more detail below. Effect of Cis/Trans Isomer Ratio. Prior work23 showed evidence that the cis and trans isomers led to different growth rates for SSZ-39. In Figure 4, this is demonstrated for the Si/Al

Figure 2. FE-SEM images of (top) faujasite and (bottom) SSZ-39. Scale bars are shown in the figure.

Figure 4. Relative amounts of SSZ-39 and FAU and solids yields as a function of trans isomer content at a synthesis time of 2 days for 1 SiO2:0.067 Al:0.14 R+OH−:0.51 NaOH:28 H2O.

= 15 synthesis (1 SiO2:0.067 Al: 0.14 R+OH−:0.51 NaOH:28 H2O). This data was measured after 2 days of crystallization to observe any growth enhancement as the pure cis fully crystallized in 3 days. The results in Figure 4 demonstrate a clear enhancement of the growth kinetics due to the presence of the trans isomer at contents above 40%. The small (but consistent) decrease at a trans isomer content of 14% was unanticipated, but clearly no enhancement is observed at 14% and 28% trans relative to syntheses performed with the pure cis isomer. Also noteworthy is that the yields for all syntheses are effectively constant (39− 42%); i.e., the yield appears insensitive to the cis/trans ratio in this synthesis.

Figure 3. Relative amounts of SSZ-39 and FAU and solids yields as a function of synthesis time for 1 SiO2:0.067 Al:0.14 R+OH−:0.51 NaOH:28 H2O using cis-N,N-dimethyl-3,5-dimethylpiperidinium hydroxide as SDA.

PXRD intensities and the solids yield. All four syntheses at 3 days appeared to be fully crystallized SSZ-39 with no FAU present, albeit there was a small amount of ANA per above. Perhaps the more interesting observation in this figure is from C

DOI: 10.1021/acs.iecr.7b00629 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research Effect of Aluminum Content. Syntheses were also performed with half the aluminum as those described above; i.e., Si/Al = 30. One difference from the Si/Al = 15 synthesis described above is that no analcime is observed. Figure 5 shows

Figure 5. Relative amounts of SSZ-39 and FAU and solids yields as a function of synthesis time for 1 SiO2:0.033 Al:0.14 R+OH−:0.57 NaOH:28 H2O using cis-N,N-dimethyl-3,5-dimethylpiperidinium hydroxide as SDA.

the SSZ-39 formation and solids yield using the pure cis isomer in a synthesis gel of composition 1 SiO2:0.033 Al:0.14 R+OH−:0.57 NaOH:28 H2O. Comparing the results in Figure 5 to those in Figure 3, two things stand out. First, the synthesis is complete in approximately 50% of the time; i.e., the Si/Al = 30 synthesis is done at approximately 1.5 days, whereas the Si/ Al = 15 synthesis is done at approximately 3 days. Second, while the yields are 18−19% at 12 h, beyond that time they are consistently in the 25−26% range with only a few syntheses giving yields in the 22−23% range (four syntheses were performed at each time point). Thus, the yield decreases by 40% when the aluminum content is decreased by 50%. If one again assumes that the aluminum from the FAU is the limiting reagent and that all the aluminum from the gel is in the SSZ-39 formed, then the theoretical Si/Al ratio should be approximately 7.5. ICP analysis on samples from this synthesis gives an Si/Al = 8.02, indicating a 93% yield based on aluminum. The effect of the trans isomer content was also investigated for the Si/Al = 30 synthesis. Figure 6 shows the SSZ-39 formation and solids yield for a range of isomer contents after 6, 12, and 24 h of heating. In contrast to the Si/Al = 15 syntheses, even at low trans contents a small, albeit reproducible, enhancement is observed. Again, at high trans (>40%) contents, a more noticeable growth enhancement is observed, as crystallization is complete at 24 h versus 36 h for the pure cis case shown in Figure 5. Also, interesting to note is that the pure cis synthesis seeded with 2 wt % SSZ-39 crystals after 24 h gives an acceleration of SSZ-39 formation (PXRD intensity ratio of 0.92 versus 0.8); however, the yield is essentially identical, i.e., 22% with seeds versus 24−25% without. The results above are consistent with the hypothesis that the aluminum is the limiting reagent and that rate of growth at a given isomer content is comparable for the two syntheses as halving the aluminum resulted in an approximately 50%

Figure 6. Relative amounts of SSZ-39 and FAU (top) and solids yields (bottom) as a function of time for various trans isomer contents for 1 SiO2:0.033 Al:0.14 R+OH−:0.57 NaOH:28 H2O.

reduction in crystallization time and approximately 40% decrease in yield. This led to the question of would this trend continue as the aluminum content of the gel decreased. Prior work by Wagner and co-workers indicated that increasing the silica to aluminum ratio of the gel led to MFI and MEL; however, those syntheses were performed at higher temperatures (433 and 443 K).13 Also of interest is if the phase selectivity will begin to shift as one continues to increase the Si/ Al of the gel. Finally, given for deNOx catalysis it is desirable to have higher silica contents (Si/Al > 10) versus what is obtained from the standard preps, we decided to pursue syntheses with even less aluminum. Silicon-Rich (Si/Al > 30) Syntheses. Figure 7 shows the SSZ-39 formation as a function of isomer content for a synthesis mixture with a Si/Al = 45 at 12 and 24 h of heating. Twenty-four hours of heating is all that is needed to form phase pure SSZ-39, and the synthesis is largely complete after 12 h. One can observe from Figure 7 that this synthesis is approximately 80% complete at 12 h for low trans contents but again complete at 12 h for 45% trans, showing the effect of the trans isomer. Attempting to quantify the exact synthesis D

DOI: 10.1021/acs.iecr.7b00629 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

Figure 7. Relative amounts of SSZ-39 and FAU as a function of trans isomer content for 1 SiO2:0.022 Al:0.14 R+OH−:0.57 NaOH:28 H2O. The synthesis time is denoted in the legend.

duration needed for these short (93%). The variation of the Si/Al of the as-made material with trans isomer is unexpected and indicates that the cis/trans ratio can be potentially used to control the Si/Al of the material obtained from synthesis. Finally, we have investigated syntheses where the gel Si/Al is 60 and 90. Twenty hours is sufficient to make pure SSZ-39 from a synthesis gel of Si/Al = 60 with a trans isomer content of 28% with a solids yield of approximately 15%. Syntheses from gels with a Si/Al = 90 were also successful in making phase pure SSZ-39 in 24 h with solids yield of 11%. The synthesis with a gel of Si/Al = 90 (pure cis isomer) yields a final material with a Si/Al = 10.1, indicating the aluminum uptake is 98%. These results are truly surprising, as one would anticipate that as the aluminum content is reduced that eventually one would make high-silica zeolites such as MFI or MEL per prior work by Wagner and the fact the SDA used here is very similar in structure to an organocation used to selectively form ZSM11 (MEL).27 That is not the case however based on our findings. While there are reports of high-silica SSZ-39 in the

Figure 8. (Top) Si/Al of the zeolite obtained and solids yields for the samples in Figure 7 after 24 h of heating (gel Si/Al = 45). (Bottom) Aluminum uptake and unit cell volumes determined by XRD for the as-synthesized samples; error bars represent one standard deviation.

literature, those reports involve fluoride-mediated syntheses. In our hands, an aluminum-rich faujasite is essential for the phase selectivity we observe; this is expounded upon below. Figure 9 summarizes the yield of solids for phase-pure SSZ-39 syntheses as a function of the gel composition. Effect of Other Synthesis Parameters. On the basis of the findings above, it is possible to make phase-pure SSZ-39 from aluminum deficient gels albeit at the cost of yield. Other synthesis parameters were investigated to increase the synthesis yield. Several syntheses were performed where the gel was the same composition of those above, but a different faujasite was used. In those syntheses, both a faujasite with a SiO2:Al2O3 = 12.3 and 30 were used. However, none of the syntheses resulted in the formation of SSZ-39. In all cases, the materials obtained were a mixture of analcime and MFI zeolites. This is at odds with prior literature20 for reasons that are not totally clear to us. Even the addition of 2 wt % SSZ-39 seeds to those synthesis gels did not result in the formation of SSZ-39. The main conclusion from these experiments is that in our hands the composition of the faujasite zeolite was crucial in leading to the rapid formation of phase pure SSZ-39. We were not successful in forming SSZ-39 with higher silica faujasite E

DOI: 10.1021/acs.iecr.7b00629 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research

observed in the synthesis outcomes. Ongoing work is exploring this issue further and will be reported elsewhere.



CONCLUSIONS A thorough study of SSZ-39 formation is presented here. There are four major conclusions from the results above. The first is that using a faujasite with a Si/Al = 2.7 it is possible to form phase pure SSZ-39 across a wide gel Si/Al ratio (15−90), something not previously reported. Second, the yield from these syntheses changes consistently and clearly shows that the aluminum is the limiting reagent, and in all syntheses, the yield based on aluminum is over 90%. Third, the trans isomer effect reported previously in the literature is confirmed and quantified across a wide range of cis/trans ratios in this work. Fourth, at an Si/Al = 45, one observes a clear increase in the Si/Al of the material obtained as one increases the trans isomer content of SDA. These results provide new insights into the formation of this industrially relevant catalyst.



Figure 9. Solids yield as a function of the Si/Al of the synthesis gel.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

materials. A possible explanation for this is given below (vide inf ra). Possible Formation Mechanism. A mechanism consistent with our findings is as follows. First, upon heating, the parent faujasite begins to dissolve. This is consistent with low solid yields observed early in the synthesis which become constant after appreciable SSZ-39 formation. At some point, one then begins to observe the formation of SSZ-39. Our results show that this induction period for a given gel composition is sensitive to the presence of the trans isomer and that this induction period is shortest at high trans (>50%) contents. This points to the trans isomer being more efficient than the cis isomer in nucleating SSZ-39. Prior literature suggests a better fit between the SDA and SSZ-39 structure in the case of the trans isomer.28 Perhaps more significant is that the solid yields are essentially constant once 20−30% (approximately) SSZ-39 is detected by XRD. This indicates that the rate of SSZ-39 growth is comparable to (or equal to) the rate of faujasite dissolution. That the synthesis duration systematically decreases as less aluminum is added to the gel is consistent with this picture, and that the aluminum is the limiting reagent under the synthesis conditions investigated here. That we cannot readily form SSZ-39 from faujasites of different composition, i.e., less aluminum, can be explained by a few possibilities. One is that the high-silica faujasites have different dissolution kinetics thus enabling other zeolites to form. This seems likely as the high-silica faujasites are steam treated and have extra-framework aluminum and silanol nests. Another possibility is that the solution species formed when the high-silica faujasites dissolve are different and inhibit SSZ-39 formation. This is also likely. While it is tempting to invoke the presence of certain aluminosilicate species in solution such as double-membered six rings as necessary to form SSZ-39, there is no experimental evidence in the literature that these species form and are stable at the synthesis conditions employed here. Thus, while it is certainly likely that the population of labile small (alumino)silicate species in solution are different for highsilica FAU materials, we discount the necessity of double sixmembered rings in solution and believe it is the dissolution kinetics which change and likely lead to the differences

Daniel F. Shantz: 0000-0002-3237-6120 Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge financial support from SACHEM, Inc.



REFERENCES

(1) Cejka, J. E. Introduction to Zeolite Molecular Sieves, 3rd ed.; Elsevier: Amsterdam, 2007. (2) Barrer, R. M. Hydrothermal Chemistry of Zeolites; Academic Press: London, 1982. (3) Breck, D. W. Zeolite Molecular Sieves: Structure, Chemistry and Use; Wiley: New York, 1974. (4) Handbook of Zeolite Science and Technology; Marcel Dekker: New York, 2005. (5) Li, J.; Corma, A.; Yu, J. Synthesis of New Zeolite Structures. Chem. Soc. Rev. 2015, 44, 7112−7127. (6) Jiang, J.; Yu, J.; Corma, A. Extra-Large-Pore Zeolites: Bridging the Gap between Micro and Mesoporous Structures. Angew. Chem., Int. Ed. 2010, 49, 3120−3145. (7) Corma, A.; Davis, M. E. Issues in the Synthesis of Crystalline Molecular Sieves: Towards the Crystallization of Low FrameworkDensity Structures. ChemPhysChem 2004, 5, 304−313. (8) Davis, M. E. Zeolite from a Materials Chemistry Perspective. Chem. Mater. 2014, 26, 239−245. (9) Moliner, M.; Rey, F.; Corma, A. Towards the Rational Design of Efficient Organic Structure-Directing Agents for Zeolite Synthesis. Angew. Chem., Int. Ed. 2013, 52, 13880−13889. (10) Li, Y.; Yu, J. New Stories of Zeolite Structures: Their Descriptions, Determinations, Predictions, and Evaluations. Chem. Rev. 2014, 114, 7268−7316. (11) Zones, S. I. Conversion of Faujasites to High-Silica Chabazite SSZ-13 in the Presence of N,N,N-trimethyladamantammonium Iodide. J. Chem. Soc., Faraday Trans. 1991, 87, 3709−3716. F

DOI: 10.1021/acs.iecr.7b00629 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Article

Industrial & Engineering Chemistry Research (12) Shantz, D. F.; Fild, C.; Koller, H.; Lobo, R. F. Guest-host interactions in As-made Al-ZSM-12: Implications for the synthesis of zeolite catalysts. J. Phys. Chem. B 1999, 103, 10858−10865. (13) Wagner, P.; Nakagawa, Y.; Lee, G. S.; Davis, M. E.; Elomari, S.; Medrud, R. C.; Zones, S. I. Guest/host relationships in the synthesis of the novel cage-based zeolites SSZ-35, SSZ-36, and SSZ-39. J. Am. Chem. Soc. 2000, 122, 263−273. (14) Goel, S.; Zones, S. I.; Iglesia, E. Synthesis of Zeolites via Interzeolite Transformations without Organic Structure-Directing Agents. Chem. Mater. 2015, 27, 2056−2066. (15) Goel, S.; Zones, S. I.; Iglesia, E. Encapsulation of Metal Clusters within MFI via Interzeolite Transformations and Direct Hydrothermal Syntheses and Catalytic Consequences of Their Confinement. J. Am. Chem. Soc. 2014, 136, 15280−15290. (16) Honda, K.; Itakura, M.; Matsuura, Y.; Onda, A.; Ide, Y.; Sadakane, M.; Sano, T. Role of Structural Similarity Between Starting Zeolite and Product Zeolite in the Interzeolite Conversion Process. J. Nanosci. Nanotechnol. 2013, 13, 3020−3026. (17) Maruo, T.; Yamanaka, N.; Tsunoji, N.; Sadakane, M.; Sano, T. Facile Synthesis of AEI Zeolites by Hydrothemal Conversoin of FAU Zeolites in the Presence of Tetraethylphosphonium Cations. Chem. Lett. 2014, 43, 302−304. (18) Kakiuchi, Y.; Yamasaki, Y.; Tsunoji, N.; Takamitsu, Y.; Sadakane, M.; Sano, T. One-Pot Synthesis of Phosphorus-Modified AEI Zeolites Derived by the Dual-Template Method as a Durable Catalyst with Enhanced Thermal/Hydrothermal Stability for the Selective Catalytic Reduction of NOx by NH3. Chem. Lett. 2016, 45, 122−124. (19) Moliner, M.; Franch, C.; Palomares, E.; Grill, M.; Corma, A. CuSSZ-39, an active and hydrothermally stable catalyst for the selective catalytic reduction of NOx. Chem. Commun. 2012, 48, 8264−8266. (20) Martin, N.; Li, Z. B.; Martinez-Triguero, J.; Yu, J. H.; Moliner, M.; Corma, A. Nanocrystalline SSZ-39 zeolite as an efficient catalyst for the methanol-to-olefin (MTO) process. Chem. Commun. 2016, 52, 6072−6075. (21) Martin, N.; Boruntea, C. R.; Moliner, M.; Corma, A. Efficient synthesis of the Cu-SSZ-39 catalyst for DeNOx applications. Chem. Commun. 2015, 51, 11030−11033. (22) Dusselier, M.; Deimund, M. A.; Schmidt, J. E.; Davis, M. E. Methanol-to-Olefins Catalysis with Hydrothermally Treated Zeolite SSZ-39. ACS Catal. 2015, 5, 6078−6085. (23) Dusselier, M.; Schmidt, J. E.; Moulton, R.; Haymore, B.; Hellums, M.; Davis, M. E. Influence of Organic Structure Directing Agent Isomer Distribution on the Synthesis of SSZ-39. Chem. Mater. 2015, 27, 2695−2702. (24) Bhadra, B. N.; Seo, P. W.; Jun, J. W.; Jeong, J. H.; Kim, T. W.; Kim, C. U.; Jhung, S. H. Syntheses of SSZ-39 and mordenite zeolites with N,N-dialkyl-2,6-dimethyl-piperidinium hydroxide/iodides: Phaseselective syntheses with anions. Microporous Mesoporous Mater. 2016, 235, 135−142. (25) Nakazawa, N.; Inagaki, S.; Kubota, Y. Direct Hydrothermal Synthesis of High-silica SSZ-39 Zeolite with Small Particle Size. Chem. Lett. 2016, 45, 919−921. (26) Wulfers, M. J.; Teketel, S.; Ipek, B.; Lobo, R. F. Conversion of methane to methanol on copper-containing small-pore zeolites and zeotypes. Chem. Commun. 2015, 51, 4447−4450. (27) Terasaki, O.; Ohsuna, T.; Sakuma, H.; Watanabe, D.; Nakagawa, Y.; Medrud, R. C. Direct observation of ’’pure MEL type’’ zeolite. Chem. Mater. 1996, 8, 463−468. (28) Schmidt, J. E.; Deem, M. W.; Lew, C.; Davis, T. M. Computationally-Guided Synthesis of the 8-Ring Zeolite AEI. Top. Catal. 2015, 58, 410−415.

G

DOI: 10.1021/acs.iecr.7b00629 Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX