You are on page 1of 203
Pe tebe ey 1) CHEMISTRY OF FCC NAPHTHA FORMATION i Us Seem ON Smcen Nae THE HYDROCARBON CHEMISTRY OF FCC NAPHTHA FORMATION Proceedings of the Symposium of the Division of Petroleum Chemistry, Inc. American Chemical Society Miami, Florida, September 10-15, 1989 Edited by Herman J. Loyink Lloyd A. Pine ‘Akzo Chemicals, Amersfoort Excon Research and Development ‘The Netherlands Laboratories, USA 1980 EDITIONS TECHNIP 27 RUE GINOUX 75737 PARIS ceoex 15 techni] (© 1990 Editions Techni AL rights reserva. No part of the publication may be reproduced or tansitied in any form or by any ean, electron or machenical including photocopy, recording, 01 any infomation storage end ratleva system, without the ovir writen permission of the publish ISBN 2-7108-0588-x Printed in France by Imprimerie Nouvelle, 45800 Saint-Jean-de-Braye CONTENTS ‘The Effect of Catalyst Feedstock and Operating Conditions on the Composition and Octane Number of FCC Gasoline ELT. Habib, Jr. ‘The Cracking Chemistry of Naphthenes A. Corma, F. Mocholi, V. Orchilles, GS. Koermer and RJ. Madon Increasing the RON and MON of the FCC Gas of Zeolite Y Synthesized with a High Si/Al Ratio .. M.A, Camblor, A. Corma, F. Mocholi, E. Iglesias and M. Pérez ine by Using Small Crystallites ‘Comparative Study of the Cracking of Alkanes and of Cycloalkanes on USHY Reaction Mechanisms .. L. Lin, N.S. Gnep and M.R. Guisnet FCC Carbenium-Ion Cracking Selectivity Effects on the Gasoline Structure and Octane Numbers. P. O'Connor, MB. Harikamp and H, Wijngaards FCC Matrix Effects on Catalytic Cracking... SJ. Yanik, A.P. Humphries, M.F. Brady, D.K. Abner and MC. Friedrich Reaction Pathways in n-Paraffin Cracking on Pure Catalysis D. Barthomeuf and C. Mirodatos Hydrogen Transfer on Zeolites with High Silica to Alumina Ratio... W.-C. Cheng, AW. Peters and K. Rajagopalan ZSM-S in Catalytic Cracking Gasoline Composition Analy D.A. Pappai and P.H. Schipper ‘The Origin and Formation of FCC Gasoline Aromatics AJ-EM. Van Klink, M.B. Hartkamp and P. O'Connor Aromatics Formation in FCC Catalysis CA. Yatsu and D.A. Keyworth ‘The Naphthenic Compounds in FCC Gasoline Characterization and Reactivity... AOL. Krause Effects of Gasoline Composition on Octane Number .... RL. Conerman and KW. Plumlee 19 25 4s 53 3 105 121 133 147 159 165 X Contents FCC Gasoline: What is Behind Octane .... C. Marcilly and M. Bourgogne The Effect of Feed Type and Operating Conditions on FCC Gasoline Properties: Pilot Plant Study... LM. Vislocky Octane Enhancement in Fluid Catalytic Cracking by Operation in the Overcracking Regime .. 3 J. Biswas and LE, Maxwell Chemical Aspects of Catcracking of Hydrotrested/Mild Hydrocracked FCC Feedstocks: “Experiences in Commercial Equipment” W. Ginzel, A. Buchsbaum and HJ. Lovink Optimization of FCC Gasoline Based on Laboratory Tests T. Mandy and L. Schmidt Catalytic Cracking of Hydrotreated Cycle Oils... M. Melin, R. Cahen, P. Bredael and J.Grootjans Catalytic Isomerization of n-Hexane and the Formation of Branched Paraffins in Industrial FCC Naphthas C..A. Mota and Lam Yiu Lau 173 183 191 199 217 223 229 image not available image not available image not available 4 BST. Habib i. Gasoline, Volume % Py ~ Naphthenes 7 0 7 o ° Olefins wo | Aromatics a w~ 0 ol ‘Mp 100 160 200 250 100 380 400 25 FOP Whole IOP I00 160 £00 50 300 $60 0 AIF FEP Whole Bailing Point Range, ° Figure 1. FCC gasoine molecular type distibution by boling point. Aniline Point, °F Bromine Number SS IBP 100 150 200 250 300 350 400 435 FBP Whole 150 200 250 300 350 400 435 FBP Whole Boiling Point, °F Figure 2. FOC gasoline composition by boiling point. image not available image not available image not available 8 ET. Habib Jr. sites, is the catalyst parameter controlling the octane number of the produced gasoline (14). Exchange cations influence performance only indirectly by causing the zeolite to equilibrate to a difierent unit cell size. itis postulated that both site density and acid strength impact catalyst performance, but both are controlled by the number of framework Al atoms and hence by the zeolite unit cell size. Low unit cell size means that there are only a few, highly isolated (and therefore strongly acidic) Al sites. Since hydrogen transfer reactions presumably require adjacent adsorbed molecules, hydrogen transfer should be greatly reduced when site density is low. Consequently, more gasoline olefins survive and the result is higher octane number (Figure 6). Increases of 2-3 RON have been reported for commercial FCC units which switched to catalysts that equilibrate to low (less than 24.28A) zeolite unit cell size (15). Compositional data on gasolines indicate that as catalyst unit coll size is reduced, Bromine Number increases, (more olefins) and Aniline Point declines (more aromatics) (16). The increase in aromatics is not easy to explain, since reduced rate of hydrogen transfer should result in a decrease in aromatics. Indeed, later work Shown in Table Ili, which included more detailed gasoline compositional (PONA) data showed that as catalyst unit cell size is reduced, gasoline aromatics decline slightly while olefins increase sharply (17). This latter result is consistent with the mechanism of decreased hydrogen transfer reactions at low unit cell size. Another effect which is somewhat complex is the impact of sodum. Figure 7 shows at constant unit cell size, lowering Na2O level to below 1.0% before steam deactivation sharply improves octane number (14) but lowers activity (17). Itis postulated that sodium neutralizes the strong acid sites. Lowering the Naz0 level exposes these strong sites which are believed to be more active for cracking than hydrogen transfer. However, there is a very different effect if the sodium is exchanged onto the catalyst after steam deactivation. In this latter case, sodium appears to only lower activity with little impact on selectivity or gasoline octane number (17). Thus, it appears that the first effect of sodium is to randomly neutralize sites. Preferential attack image not available image not available image not available 12 ETT. Habib. 5 8 4 3 Gasoline 4 viel, Wi%e 2 1 40 : 00 10 20 30 Catalyst ZSM-5 Content, % Figure 8. Effect of ZSN-5 content on FCC gasoline yield and octane number. Pilot plant data from Biswas et al. *9 10 Butylene + Propylene Change in Product €5-C10 Aromatics and Saturated Cyciies Yield, Wi% wesees ° €5-C10 Branched Olefins ~ €5-C10 Paraffins + 0.0 1.0 20 3.0 Catalyst ZSM-5 Content, % Figure 9. Effect of ZSN-5 content on component yields. Pitot plant data from Biswas et al. %9 Effect of Catalyst, Feedstock, Operating Conditions 13 gasoline paraffins and olefins to light olefins, because when gasolines produced by ordinary cracking (no ZSM-5) are contacted with ZSN-S, little conversion of paraffins occurs (25). Consequently, it appears that the ZSM-5 must act on paraffin precursors rather than the final gasoline paraffins. Conclusions Although octane number is an empirical measurement, most FCC gasoline octane number changes can be directly correlated to shifts in the molecular composition of the gasoline. Feedstock changes generally impact the aromaticity of the gasoline, with aromatic feeds producing high octane number aromatic gasolines. With the exception of feedstock, most variables that have a major impact on gasoline do so by changing the relative rates of cracking and hydrogen transfer, thus shifting the olefin content of the produced gasoline. Literature Cited 1. American Petroleum Institute Project 45, 1956. 2. Owen, H., Snyder, P. W. and Venuto, P. B., Proceedings of the Sixth International Conaress on Catalyisis, 2, 1071 (1977). 3. Creighton, J. E., Davison Catalagram, Z1, p. 14 (1984). 4. Habib, Jr., E, T., Owen, H., Snyder, P. W., Streed, C. W. and Venuto, P. B., Ind. Eng. Chem. Prod. Res. Dev., 16, 291 (1977). 5. Beech, J. H., Gross, B. and Ramage, M. P., U.S. Patent 4,459,203 (July 10, 1984). 6. Batchelder, M. L. (Editor), Krider, J. B., Spanger, Jr., C. W., Whittington, E. L. and Witoshkin, A., 1985 NPRA Q and A Session on Refining and Petrochemical Technology, Farrar and Associates, Tulsa, OK. p. 34 (1986). 7. Batchelder, M. L. (Editor), May, M. A. and Thiel, P. G., 1986 NPRA Q and A Session on Refining and Petrochemical Technology, Farrar and Associates, Tulsa, OK, p. 39 (1987). 4 10. "4 12. 13, 14, 15. 16. 17. 18. 19, 20. 21. ELT, Habib Jr. Magee, JS. and Ritter. R. E., Formation of High Octane Gasolina by Zeolite Cracking Catalysis, Presentation before the Division of Petroleum Chemistry, ACS Miami Beach, FL Meeting, Sept. 4-8, 1978. Venuto, P. B. and Habib, Jr., E. T., Fluid Catalytic Cracking with Zeolite Catalysts, Marcel Dekker, New York, 1979, (p. 7). Johnson, A. R., Recent Advances in Residual Oil Fluid Catalytic Cracking, Paper Presented at the 1986 NPRA Annual meeting in Los Angeles, CA, March 23-25, 1986. Ritter, R. E., Blazak, J. J., Wallace, D. N., Oil and Gas J., 72, p. 99 (1974). Magee, J. S., Ritter, R. E., Wallace, D. N. and Blazek, J. J., How Catalytic Cracker Feed Composition Atfects Octane Catalyst Performance, Paper Presented at the New Orleans, LA NPRA Meeting, March 23, 1980. Batchelder, M. L. (Editor), Leroy, C. F.,19686 NPRA Q and A Session on Refining and Petrochemical Technology, Farrar and Associates, Tulea, OK, p. 41 (1987). Pine, L.A., Maher, P. J. and Wachter, W. A., J. Catal, 85, 466, (1984). Habib, Jr., E. T., Davison Catalagram, 64, p. 1, (1984). Ritter, R. E., Creighton, J. E., Roberie, T. G., Chin, D. S. and Wear, C. C., Catalytic Octane from the FCC, Presented at the 1988 NPRA Annual Meeting in Los Angeles, CA, March 23-25, 1986. Edwards, G. C., Rajagopalan, K., Peters, A. W., Young, G. W. and Creighton, J. E., Strategies for Octane Enhancement in a Fluid Catalytic Cracking Unit, ACS ‘Symposium Series 375, p. 34, ACS, Washington, D. C. (1988). Addison, S. W., Cartidge, S., Harding, D. A. and McElhiney, G., Applied Catalysis, 45, 307 (1988). Pellet, R. J., Blackwell, C. S. and Rabo, J. A., J. Catalysis, 114, 71 (1988). Cotterman, R. L., Hickson, D. A. and Shatiock, M. P., Relationship Between ‘Structure and Catalytic Pertormance of Dealuminated Y Zeolites, Paper Presented atthe Los Angeles, CA ACS Meeting, September 1988. Magee, J. S. and Blazek J. J., Preparation and Performance of Zeolite Cracking Catalysts, ACS Monograph No. 171, 652 (1976). 22. 23. 24, 25. Effect of Catalyst, Feedstock, Operating Conditions 15 Deady, J., Young, G. W. and Wear, C. C., “Strategies for Reducing FCC Gasoline Sensitivity”, Presented at the 1989 Annual NPRA Meeting in San Francisco, CA, March 19-21, 1989. Donnelly, S. P., Mizrahi, S., Sparrell, P. T., Huss, Jr., A., Skipper, P. H. and Herbst, J. A., How ZSM-5 Works in FCC, Presented before the Division of Petroleum Chemistry, ACS New Orleans, LA Meeting, August 30 - Sept. 4, 1987. Biswas, J., Kak, H., van der Griend, J. and Maxwell, |., Laboratory Studies on ZSM-5 Addition in FCC Processing, Paper No. F-12 Presented at the Akzo Catalyst Symposium in Schereningen, Netherlands, May 29 - June 1, 1988. Rajagopalan, K. and Young, G. W., Hydrocarbon Cracking Selectivities with Dual Function Zeolite Catalysts, ACS Symposium Series 375, p 101, ACS, Washington, D.C. (1988). Table! Blending Octane Numbers of Selected Hydrocarbons?) Research Octane No. Motor Octane No. Sensitivity (RON) (WON) (RON-VON) Parattins n-Pentane 62 67 “5 n-Octane 19 15 “4 2-Methyl Heptane 13 24 ar) 2,3,4 Tri-Methy! Pentane 97 102 “5 Olefin 1-Heptene 66 53 13 2-Methyl-1-Hexene 118 108 10 2,3,4 Tri-Methyl-2-Pentene 142 130 12 Naphthenes Methyl-Cyclo-Pentane 107 99 8 Ethyl-Cyclohexane 43 39 4 Aromatics Benzene 99 a1 8 Toluene 124 112 12 Xylene (Para) 146 127 19 16 EFT. Habib J. Table II Effect of Feedstock Resid Content on FCC Gasoline Octane Vol.% 1000°F+ In Feed ° 17 27 Conversion, Vol.% 826 80.7 78.5 Gasoline Yield, Vol.% 65.7 61.5 59.5 Gasoline Octane, RON 907 91.6 94.7 Commercial Results Reported by A. R. Johnson (10). Table it 31 74.5 56.3 94.9 Gasoline Molecular Type Shifts Due To Unit Cell Size 10% REY Unit Cell Size, A 24.46 Gasoline Composition, WI% Paretfins 375 Olefins 219 Naphthenes 126 ‘Aromatics 22.4 Research Octane No. 88.6 Motor Octane No. TA Data trom Edwards ot al.(17) 20% USY 24.26 29.3 342 12.4 20.2 92.6 30.0 image not available image not available image not available image not available image not available image not available image not available 26 M.A. Camblor, A. Corma, F. Mocholi, E. Iglesias, M. Perez INTRODUCTION The actual tendency to produce gasoline from the FCC unit with higher Research and Motor octane number (RON and MON), has produced an intense research effort in traying to understand better the chemistry involved in the cracking and other accompanying reactions. In this way, the role of the different type of active sites present on the zeolite and their evolution with the different activation and deactivation procedures has been reviewed (1,2). Moreover, and from the point of view of both RON and MON , it is desirable to maximize the content of i- and n-olefins, i-paraffin and aromatics, while minimizing the presence of n-paraffins in the gasoline fraction. In order to do so one can work on the process increasing, for instance, the temperature and decreasing the contact time (3). It is also possible to influence the product distribution in the gasoline, by properly selecting the feed (4), and finally it is possible to modify the cracking pattern by controlling the chemical and physical characteristics of the catalysts (5). Looking into the third possibility, some authors have presented that vhen using Y zeolite as the active component, the gasoline yield and gasoline octane, depend exclusively on the framework 8i/Al ratio of the zeolite (6,7). Others, however, show that the presence of the different types of extraframework Al (EFAL) has a clear impact on the final behaviour of the catalyst (1,2,8-11). There are however image not available image not available image not available image not available image not available image not available image not available M.A. Camblor, A. Comma, F. Mocholi, E. Iglesias, M. Perez gasoline from heavier feed with a high content in aromatics should be higher. In other words, smaller crystallites synthesized with a high Si/Al ratio is a clear via to improve the RON and MON of the gasoline produced in a FCC process, as well as the octane per barril. REFERENCES 1. R.A. Beyerlein, G.B. Mc Vicker, L.N. Yacullo, and J.J. Ziemak, J. Phys. Chem., 92 (1988) 1967. 2. A. Corma, V. Fornés, A. Martinez, and J. Sanz, in "Fluid Catalytic Cracking: Role in Modern Refining", M.L. Occelli Ed., ACS Symp. Ser., 375, Chap. 2, (1988) p.17. 3. J. Biswas, and I.E. Maxwell, 8th ZC, "Zeolites Facts, Figures, Future", P.A. Jacobs and R.A. van Santen Eds., Stud. surf. Sci. Catal., B, 49 (1989) p. 1263. 4. HF. Henz, V.M. de Marco Meniconi, and J.M. Fusco, Ketjen Catalyst Symp., Scheveningen, (1986). A. Corma, in "Zeolites: Facts, Figures, Future", sth IZC, P.A. Jacobs and R.A. van Santen Eds., stud. Surf. Sci. catal., A, 49 (1989). 6. L.A. Pine, P.J. Maher, and W.A. Watcher, J. Catal., 85 (1984) 466. 7. G.C. Edwards, K. Rajalopagan, A.W. Feters, G.W. Young, and J.E. Creighton, in “Fluid catalytic image not available image not available image not available 38 M.A. Camblor, A. Corma, F, Mocholi, E. Iglesias, M. Perez Yield (0) ‘ao fil) 20 2 13] 19) os aa 0.) ope=— oC CRO 70 ‘Conversion (8) ‘Gomerson (4) Vela ie (3) ase ~ egrets) c7 ca 20 15] 15) 19) os| 70 Carrer) ons 010] 08 00! Cr ar ar ae ee) 0 8 80 4 OD conversion (6) Conversion (8) Figure 3a. Yield of isoparaffins versus conversion (Gases + Gasoline + Coke),(m) SK-4O (#) IC-1 image not available image not available image not available 42 MA. Camblor, A. Corma, F. Mocholi, E. Iglesias, M. Perez Yield (8) Vist) 0.00| 0.08 0.04 0.09} 0.02] a a ee ee, a Conversion (4) oo 66070 0 3040, Conversion (%) Figure Se. Yield of saturated naphthenes versus conversion (Gases + Gasoline + Coke).(#) SK-40(e)IC-1 image not available image not available image not available image not available image not available image not available image not available 50 L.Lin, NS. Grep, + Tsonerization + Primary cracking 1 n pS m =< + Secondary cracking ye 7 = « Alkylation eg. + Hydrogen transfer 3 Olefins + Gy" + Gt —e gt Big MBCer & e * ie — a oe O— samme & Figure 5. Mechanism of methylcyclohexane transformation () into the carbenium ion il which cracks through mode C, and (li) into the carbenium. ion Il which cracks more slowly through mode D (tertiary-primary). The cracking of the other certenium ions occurring through mode D (secondary-primary) is very slow and there- fore can be negjected. Alkylation reactions followed by B isomerization and A, B, C scis- sions are responsible for the formation of C4- Cs products. Two examples of alkylation are given in Figure 6. The low o/s ratio results from the secondary transformation of olefins into coke. Mechanism of Methylcyclohexane Trans- formation Methylcyclohexane isomerizes into dimethyl and ethyicyclopentane, which leads to Cy-Gg aliphatic hydrocarbons, and ul- timately to toluene and to coke. Figure § shows how these products can beformed. The mathylcyclohaxyl carbonium ions (I) are trans- formed through B isomerization into dimethy! of ethylcyclopentyicarbenium ions (II and Ill). |, land Ill crack through aC mode into olefinic Carbenium ions with a rheptyl or a methyl- hexyl skeleton. Through hydride transfer from methylcyclohexane, these olefinic carbenium ions are transformed into olefins. These olefins can undergo numerous secondary transfor- mations: ¢ cracking into propene and butenes ¢ alkylation followed by Isomerization and cracking leading to C3-C7 dlefins ¢ coke formation © saturation by hydrogen coming from methylcyclohexane with consequent- ly formation of toluene. It must be noted that all the olefinic products can participate in reactions 2 to 4. Moreover the transformation of the C7 com- Pounds (olefins or alkanes) is much more rapid than their formation since no C7 products are observed. Reactivity of Hydrocarbons Cracking Rates The classification of the reactivities in cracking that are discussed here in the ight of the reaction mechanisms is that determined for a 1 wt % coke content on catalyst. The modes of carbenium ion transformation (sur- face reactions} leading to the main light products are : A cracking for isooctane, B isomerization and C cracking or2 B isomeriza- tion and B cracking or 3 B isomerization and A cracking for r-octane, B isomerization + C cracking for n-hexane, C cracking or B isomerization + B or C cracking for methyl- cyclohexane. Consequently if the surface reactions the limiting step, isooctane must be image not available image not available image not available FCC Carbenium-Ion Cracking Selectivity 55 PILOT PLANT RISER DATA P+PN ‘wt on gsin 140 130 14 No metals, 65%wt Conv. No metals, 70%wt Conv. Metals, 65%wt Conv. post ICeince, 95 10.9 83 trat et ‘on gsin 142 175 173 tutA = light Aromatic: Benzene, Toluene, Xylenes. 4008 = Degree of Branching (8) depressors" and maximize the “octane boosters". The question becomes how to adapt FCC kinetics in orderto achieve these goals. The Effect of Cracking Reactions on Gasoline Structure Laboratory as well as commercial data show that the degree of branching and the aromaticity of FCC gasalines clearly increase with conversion (7). We have established that this is not solely a concentration effect. Furthermore, the data above (irom (7)) show that the presence of metals on FCC catalyst can resutt inan increase in aromatics in gasoline, however with some penalty in The “thermal activity” required for aromatics formation may be interpreted as energy required to shatter the large cyclic molecules in the feed. However, the similar effect of metals and reactor temperature may also indicate that some dehydrogenation ac- tivity is required forthe formation ofaromatics. ‘If we reflect on the formation of aromatics from saturated cyclic compounds, it is clear that a kind of dehydrogenation is required in ‘order to form unsaturated naphthenes, which can be converted to aromatics via hydrogen transfer reactions. The decrease in gasoline branching with metals and high reactor temperatures indi- NON-CARBENIUM ION CRACKING Olefin Thermal or Nono-molecular. Formation Pyrolytic cracking. (Radical) Dehydrogenation Mono-molecular. (Metal) CARBENIUM ION CRACKING Olefin Isomerization Mono-molecular. Reactions Cyclization Mono-motecular. Cracking-Branching Mono-molecular. Alkylation Bi-molecular Disproportionation Bi-molecular Hydrogen transfer Bi-molecular gasdline branching. ‘When we consider the effect of FCC reac- tor temperature, we also see that a higher Teactor temperature enhances aromatization, while this increase in “thermal” cracking wil result in a decrease of the gasoline branching (see Figure 7). cates a loss in carbenium-ion cracking selec- tivity. Obviously a well balanced combination of carbenium-lon cracking, hydrogen transfer, and even some dehydrogenation may be re- quired to obtain the optimum gasoline struc- ture. image not available image not available image not available FCC Carbenium-Ion Cracking Selectivity 59 TABLE 2. CRACKING SELECTIVITY WITH CETANE (nCt¢) t YIELDS: Conv. ica FGAS icsincs Sot in mol/100 mois cracked ‘THERMAL CRACKING 31.5 1 2.00 0 20.9 1 172 ° 26.6 0 59 0.09 68.4 1 54 0.06 CATALYTIC CRACKING Alumina (low SiO2) LHSV=05 61.7 3 2.72 02 Silica-Zirconia-Alumina LHS' 11.0 22 34 22 LHS\ 24.2 25 29 26 LHSV=4 40.0 32 34 22 LHSV=2 53.5 37 39 3.3 LHSV=1 68.0 44 4 3.8 ‘pata from B. S. Greensfolder ot al. (2). Ls TABLE 3. CRACKING SELECTIVITY OF SiO2/Al203 VS ZEOLITE CATALYSTSt CATALYST SiO2/Al203 REHX REHY in SiO2/Al203 matrix DURABEAD: ==———s(«CBTt—“<«t‘i«‘«éiU iS DB7~ DBs COMMERCIAL DATASET 1 Conversion, %wt 528 66.6 M2 724 ica ewe 44 42 46 54 FUELGAS — %wt 22 25 23 1.8 He, CHa, C2Ha, CoHe FUEL GAS/ ica 0.50 0.60 0.50 0.33 COMMERCIAL DATASET 2 (vol on gasoline) Paraffins 87 21 Naphthenes 10.4 19.3 Olefins 437 146 ‘Aromatics 37.3 45 Degree of Cyclization = (N+AV(O) 1.09 4.45 ‘pata from S. C. Eastwood et al. (6) Commercial data on moving bed TCC. ‘Data from B. S. Greensfelder et al, (2). image not available image not available image not available FCC Carbenium-lon Cracking Selectivity 63 GAS (C41+C2+C2=) VS CRACKING TEMPERATURE n-hexane pulse testing GAS, Xwt 70 60 50 40 30 20 10 AY zeolite FCC & 7SH-5 — Guartz ~®- $102-A1203 -LA --- “HA * -SA TEMPERATURE, degC Figure 2. RON VS GASOLINE P+PN Commercial FCC gasclines 42.14 #16 18 20 22 Paraffins + Paraffinic Naphthenes, wt Figure 3. image not available image not available image not available FCC Carbenium-Ion Cracking Selectivity 67 EXTRA FRAMEWORK ALUMINA DISTRIBUTION AL-ENAICHED Hign Surface EFAL *AES* Poor selectivity. AL-DEPLETED — Lack of FAL and EFAL “as" Poor Activity. HOMOGENEOUS = Balanced FAL - EFAL “HAS* Good Activity and Selectivity. Figure 10. VERY SIMPLIFIED ZEOLITE MODEL On Ss Si Si Si High Activity Low Activity High "Branching" Low "Branching" Low Fuelgas High Fuelges Poor Stability Good Stability Figure 11. image not available image not available image not available FCC Carbenium-Ion Cracking Selectivity 71 FCC OCTANES AND ZEOLITE TYPE FCC Pilot Riser Data WON - 0 B81 8 10 RE203 on Y, Xwt. Figure 18. Literature Cited (8) Pine, LA, Maher, PJ, and Wachter, (1) Oblad, A.G., Milliken, T.H., and Mills, G.A. W.A, J. Catalysis, 85, 466-476 (1984) in"The Chemistry of Petroleum (9) Long, GH. et al, 12th World Petroleum Hydrocarbons". Brooks, Kurts, Bord, Congress, Houston, 1987. and Schemeriing, Reinhold (1955). (10) Beyerlein, RA, McVicker, G.B., Yacullo, (2) Sroonatelder, B.S, Voge, H.H., and Pie and ao oe J.J., J. Phys. Chem, i, G.M., ind. Eng. Chem, 1949, 92, 1967-1970 (1988). 2573. (11) Sanz, J., Fomes, V. and Coma. A., J. (8) Wojciechowski, B.V., and Corma, A., Chem. Soc. Faraday, Trans 1, 84 (9), “Catalytic Cracking: Catalysis, 3113 (1966). Chemistry, Kinetics” Marcel Dekker (12) Pellet, Ru., Blackwell, C.S., and Rabo, Inc; New York, NY (1986). JA, J. Catalysis, 114, 71-89 (1988). (4) Gianetto, G., Sansare, $. and Guisnet, M., (13) Scherzer, J.. ACS Symposium Series, J. Chem. Soc. Chem. Commun, 1986, 248, 157 (1984). 1902. (14) Addison, 8.W., Cartidge, S., Harding, (6) Adams, C.E. and Kimberlin, GN. Jr., 3rd DA. and McElhiney, G., Applied Int. Congress on Catalysis, Amsterdam Catalysis, 45, 307-323 (1988). 1964. (18) Corma,A. and Orchiles, A.., J (©) Eastwood, S.C., Plank, C.J., and Weisz, Catalysis, 115, 551-566 (1989) P.B., 8th World Petroleum Congress, (16) O'Connor,P., and Hartkamp, M.B., London, 1971. ‘AKZO Catalysts Symposium 88, (7) O'Connor, P., AKZO Catalysts Sym- Scheveningen 1988. posium '88, Scheveningen 1988. image not available image not available image not available 76 S.J. Yanik, AP. Humphries Thus, once again it is the ratio of zeolite to active matrix surface area that determines catalytic performance. Figures 11 and 12 illustrate the magnitude of the changes in hydrothermal stability that occur as active matrix and/or zeolite content are increased. Note that increasing relative active matrix content from 0.4 to 1.6 improved hydrothermal stability by 2.8% = 3.5%. Increasing relative zeolite content from 0.75 to 1.75 reduced stability by 0.6 — 1.18 Thus, if attempts are made to compensate for the lack of activity in low active matrix content catalysts by increasing the amount of zeolite, hydrothermal stability suffers. Effect Of Matrix Composition On Product Yields And Selectivities Coke and Light Gases Figures 13 thru 16 show the MAT yields and selectivities for light gases (Cj) and coke. Because this study was designed to emphasize matrix composition effects, short deactivation time (2 hours) and high deactivation temperatures (1450 and 1540°F) were chosen. As mentioned previously, at these conditions, the zeolite will deactivate much more rapidly than the matrix. Thus, at constant conversion, the higher matrix activity catalysts produced more coke and light gases at the expense of liquid products. While these are expected effects for long catalyst contact time maT studies, these effects, are not generally observed during commercial operations.1/2 While beyond the scope of this study, future work will address the effects of deactivation time and temperature and MAT catalyst contact time on relative yields and selectivities of coke and light gases. LEG Figures 17 and 18 show LPG yields and selectivities vs. conversion. The data indicate that at constant conversion, the high active matrix content catalysts produce less LPG. This is consistent with matrix activity converting heavy oils to LCO and gasoline. Since much of LPG production comes trom secondary cracking of lighter molecules by the zeolite, it follows that high matrix activity should produce more LCO and gasoline, while low matrix activity produces more LPG and C2~ gases. image not available image not available image not available 80 S.J. Yanik, A.P, Humphries Future Work This is the first of a series of studies on effects of catalyst composition. Studies are underway to define the effects of deactivation conditions, and to compare results with those from testing equilibrium catalysts over the came conversion range. Sone matrix materials enhance metals tolerance. Tests will be performed on catalysts impregnated with nickel and vanadium. Since active natrix materials contribute relatively large pores to the catalyst, they are more effective in cracking heavy hydrocarbons. Tests with residual feed stocks will be done. Hydrocarbon composition of products will be defined. These studies will be published as the results becone available. References 1. Yanik, §.J., Brady, M.F., and Havey, J “optimizing Zeolite content in Fcc catalysts", to be published in Oi] and Gas Journal. 2. lLipinski, J.J. and Wilcox, J.R., "Octane Catalyst Boosts Octane, Resid-Charging Capability", Oi] and Gas_Journal, November 28, 1986. image not available image not available image not available S.1. Yanik, A.P. Humphries 84 e/ew vany aovauns YOY L/W SAS TAS BRS XIULWN FALOY 40 103449 » anita aD BLT SAS WS CN XIMLWA SALOY 40 103443 © aanSts Brew wan sovaune image not available image not available image not available S.J. Yanik, A.P. Humphries 88 nano an SL aN YE a oz gl on oF or omy or orD oro Py poi poesiiiiy a 3 sr i a4 = he FT “| «3 1 = ma i 2 ne Be 6 i sch = 4 Oa i wen = we seo = 24 OO SOU A WO ° RIT SOU ARO XIMLWN JAW 40 104444 XIMIWA BALDY 40 193413 zy eanfra tr asnsea o o coe—mave> NoUNmIaN LIALOY a2 vitae image not available S.J, Yanik, A.P. Humphries 90 SYSHSSHSSAARHRAANA a3 3 spe os a) OL OL OL OM OB OM OH Oe CH me Oe TL CH OL OM Om Ow OW rm Ow oe re perp iiiiiiiiiing poiriiirriiiiiiiiiis ci n eae ’ 0 a ye ng * sts NN non \ nf = SA S, Ye uy n a oy a" ist=Shi Rn C3 Vos § ANCES 30 * ose 93H) XIMLYN SALW 40 103443 XIMIYN SNLOV 40 193443 ot aansra st eins 1 FCC Matrix Effects 2 SEIN 7 WOSGANO OM OF 0% Oh OW OM 1 (079 Ow Hm OM OM Ch OU OU OO O99 OW 0 OM Om OM , dim 3 ° . gy © oF | f sm0=5M ce. ws wt, ‘NOSEANOO & AUALOTES S11 NOSE % 9 YIMLYW JALOY 40 193443 YIMLWN SAL 40 193443 1 ean8zz AT eansta o o om S.J. Yanik, A.P. Humphries 92. oa 328 3 WHT a Tw oO ‘Pe yosen a Osa 0% Oh OL CM ON OW HH Oy Om re OMS OX Oh OL OM OM 0M OW Ot OM Om ONS PPA eg PLE ig 1 ‘ me ” , SN —— 7 o 7 ; N, 7 s ™ we 8 a, ‘ age russ 0 ” Ne 0 ml . ON, ON ot nt mie, ft ot ‘NOSED & 90/001, ‘NOISAGINOD YO TIOKD XIV AWLOY 40 10343 XIMIWH SALO¥ 40 L037 oz arn8ra 61 east 93 FCC Matrix Effects sao un 3H ean Ra 3 SE A) Ce a 2) Popes iritritg Pa a 0 +90 0 a 0 0 60 o 270 We o o os noR " Or = ay a o§ 1 nn 8 in s sg $1 1 i a eel = aa sco 24 a oT a ° og 6 m sei 724 m u w a sci= 2H u % @ Ye " 7 sw ‘uno 9 on/en wad aI LW Sh 00/7 ETO NIMLWN SALOY 40 103443 XIMIWN SAV 40 193443 zz sandra ae onda ce-onmsve> cavoon viraa S.J. Yanik, A.P. Hurephries 94 spe ose YBN OL Oh OU OM Ot OM OO OX Oh OU ON OB OM OW Cy OM Oe OE pee rip erie ity boiiiirriiiiiiiiiit m0 (ro=sha ¢ wo . af wn a ate ° 4 \ N\ sro si j a ; Ne NY ‘50 j “ wo mes 0 NOSGNNCO $4 UNLESS NOSGANOO © IANO. XIMLYA SALOW 40 103449 XIMIWN ANOY 40 193443 oe asndta ex vans BRAS RRRARSRASSSSRILIS PARAFFINS. wi a e204 90.0 4 28.0 ~ 250 4 24.0 220 20.0 78.0 760 740 72.0 70.0 560 FCC Mawrix Effects 95 Figure 25 EFFECT OF ACTIVE MATRIX OCTANE INDEX va CONVERSION FON => ¢____. Mon ve S80 0.0 620 84.08 70.0 720478. CONVERSION, wi, na FA/se2.67 OP FA/S=0.23 Figure 26 EFFECT OF ACTIVE MATRIX GASOLINE COMPOSITION vs CONVERSON 80 600 520 64.008. 7.072 HO. 78 S.J. Yanik, AP. Humphries Figure 27 EFFECT OF ACTIVE MATRIX GASOLINE COMPOSITION va CONVERSION 3804 s704 380-4 3 8 ss04 2 g so 3304 s204 50.0 «20 6406068700720 74D 78.0 CONVERSION, wt Figure 28 BTX, BENZENE VS ZEOLITE CCl Feed 18.00 17.50 x 17.99 = 1650 x x x x 1600 * RAM © 1550 66 15.00 AGS. 14.50 100 1250130175 ogo ye 080 = 070 060 —— oso RAM 0.40 oe 88 O30 —S *042 0.20 100 125° #4150 «175 RELATIVE ZEOUTE CONTENT FCC Matrix Effects 97 Figure 29 BTX, BENZENE VS ACTIVE MATRIX ISO-PARAFFINS, wi CCl Feed 1800 1750 | ye 17.00 = 1650 a 16.00 7 5 1550 Se Se 15.00 150 14.50 [eave 042 158 1.00 030 g 080 0.70 060 RZ oso 150 0.40 — — 0.30 ee] 0.20 0.42 158 RELATIVE ACTIVE MATRIX CONTENT Figure 30 EFFECT OF ACTIVE MATRIX s CASOLNE cOuPOSTON ve CONVERSION 4407 410 ‘an: Fa/s = 2670 330 380 ee m0 nt o2 560 580 6008204 BBD BOOT 480 98 NAPHTHENES, wt S.J. Yanik, AP. Humphries Figure 31 EFFECT OF ACTIVE MATRIX GASOLINE COMPOSITION vs CONVERSION 12.04 18.04 vo 1604 iso 1404 304 1204 no 10.0 4 204 804 704 60 +FAS = 0.23 FNS = 2.670 360 10.0 seo 62.0 620700729. 740-780 CONVERSION, wt Figure 22 EFFECT OF ACTIVE MATRIX GASOLINE COUPOSTION vs CONVERSICN se 974 ee os a4 2 324 ai 204 as ae ac es as 324 80 | \ ne Fas = 023 58.0 380 62.0 62.0 GLOBO 700720740780 CONVERSION, wt REACTION PATHWAYS IN n-PARAFFIN CRACKING ON PURE CATALYSTS D. BARTHOMEUF", C. MIRODATOS? Introduction ‘The cracking of hydrocarbons gives rise to a large variety of products. Several classes of selectivities have been observed which can accountfor such results. The shape selectivity based on geometrical constraints explains for instance the higher rate of n-hexane cracking compared to that of 3-methyl-pentane in MFI (1) of the absence in erionite of branched isomers in the cracking of n-paraffins (2). For larger pore zeolites, for instance faujasit this selectivity directs the formation of products specific of the size of the windows present in the catalyst in the hydroconversion of n- decane (8). Another class of selectivity arises from the acidic properties of the zeolites, acid strength and site density. It modifies the for- mation of olefins over saturated products or the isomerization over the cracking and is ‘extensively studied (4). A third approach con- centrates On the factors which direct the C-C_ splitting, Le. how and where a molecule is attacked, since some zeolites favor the production of light products Cy, C2, Cg, over the heavier ones (5-8), Two reaction mechanisms have been invokedto explain the results besides the classical beta-scission, viz., the pentacoordinated carbonium ion mechanism (5-9) and a radical cracking mechanism particularly on steam dealuminated faujasites (4). Since inside the zeolites molecules are embedded in an electrostatic field and a fleld gradient which may modify thelr electron dis- tribution (10), the effects of those electrostatic parameters on the reactivity of the C-C bonds have to be considered. A long time ago Rabo (11-12) proposed the polarizing influence of the electrostatic field of cations to explain the high catalytic activity of some zeolites. There- 4, Labora 75252 Patls Ged 2, lnsitut do Rcherch fore, the system zeolke-hydrocarbon has in fact to be analyzed as a whole which leads to a charge density rearrangement and some modifications in interatomic distances and bond angles compared to pure components, catalyst and reactant (13). The charge on each atom of tho system depends on the external potential generated by the surrounding atoms (14). One may then expect a modification of the reactivity of hydrocarbons depending on their environment in the cages i.e. on the atomsand charge distribution in the wall ofthe cavity. This justifies that the field gradient inthe ‘cages has been proposed to explain the chan- ges in the production of light products in the cracking of n-paraffins in various zeolites ‘structures (7,8). The present paper gives an ‘over-viow of this approach. Experimental Materials The zeolites used have been described in detail previously (7,8). They consist of protonic forms of faujasite, MFI, erionite, offretite and mordenite, The cracking of n-heptane and n- octane were carried out at 723 K and that of n-decane at 793 K in a microreactor as described in (7). The actual conversion in steady-state conditions was limited to less than 10% by adjusting the weight of the catalyst. The results are expressed in molecules on n-paraffins converted per ‘second and per gram or as the porcont con- version calculated for 100 mg of catalyst. Results The total formation of the products with a given carbon number C2, C3, Cy... Is calcu- lated by adding the concentration of the re- lated products whether they are olefinic ornot, ‘Surface et Structure, Université Pars VI, 4 place Jussieu, £69826 Villeurbame Coder, 100. Barthomeuf, C. Mirodatos branched or not. It was observed that for the three hydrocarbons tested, significant chan- ges In product distribution occurred between Cgand Ca. The zeolites leading tohigh Caand C3 formation give low amounts of the heavier hydrocarbons Cq and C4. The ratios C2/C4 and Ca/C4 clearly appear to represent this trend and express a C-C spitting selectvity. They can be considered as indexes of this selectivity. Table 1 gives the Ca/C4 values for the cracking of the three n-paraffins on the different zeolites tested. It also reports values calculated from previous results on 2,2,4 trimethylpentane cracking at 738 K. Figure 1 gives the changes in the ratio C4/Cq as a function of the percent conversion of n-C7 of n-Cio for various zeoltes. The ex- perimental points are obtained from curves Discussion Table 1 shows that the different zeolites are ranked from high to low Cg/Cq values according to a similar sequence whatever the carbon number of the n-paraffin feed is. Only MFlanderioniteare different for n-Cy and n-Cg compared to n-Cjo. This may result from the well known window effect in erionite (18). The branched paraffin 2,2,4-trimethylpentane gives the same order as the n-paraffins for the four zeolites tested. The results indicate that the relative production of light products C2, C3 is not constrained by a mechanism depending on the length of the molecule or its linear or branched character. It rather suggests that an intrinsic zeolite property governs the C-C spiit- ting location. The dependence of Ca/C4 ratio on the percent conversion may give informa- 10 (A) or n-heptane (B) on HY: 4 20 30 Yoconversion MOR:0, ZSM-5: @ HERE A Figure 1. C3/C4 ratio as a function of per cent conversion in the cracking of n-decane giving activity as a function of time on stream or contact time. For each curve the high con- version values correspond to the beginning of the reaction before steady state conditions are reached. Table 2 reports the activity and the selectivity C3/C4 for faujasites with different Al content, in the framework or extraframework. The influence of cations at the same exchange level is described in Table 3 in relation to the polarizing powor of the cations expressed as Zejr2 (Z cation charge, r cation radius). tion on that point (Figure 1). It Is usually ob- served that the selectivities in acidic catalysis change with the conversion, selectivity involv- ing mainly modification of cracked fragments (isomerisation, hydrogenation...) with no change in carbon number. The previous studies have shown the expected changes in specific product distribution (olefin/paraffin, branched/linear) (7) which may be correlated with modification in acidity (4). By contrast with this type of selectivity, Figure 1 shows that the C-C splitting selectivity Is almost inde- pendent on the conversion level, in steady Reaction Pathways in n-Peraffin Cracking ‘101 Table 1. CHANGES IN THE RATIOS C3/C4 IN THE CRACKING OF PARAFFINS@) Feed: C3C4 Ratio Zeolite nc7 na nCio 18 MFI 1.22 og aa HERI 17 09 23 HERI 17 o8 HOFF 096 077 1.68 0.83 HOFF 0.83 048 127 0.13 MOR 0.86 043 1.39 0.12 HY os 0.48 133 014 HNeY 132 HNeX 1.03, @) Similar trends obtained for Ca/C4 state conditions, for a given reactant and a zeolite. It also shows no influence of coke ‘deposition which among other effects poisons the acid sites, and changes the conversion and the selectivity (16). The absence of cor- relationwith acid strength is seen also in Table 1 where H-ERI and H-Y materials have the same Ca/Cq ratios as H-K-ERI and H-Ne-Y which have a quite lower acid strength. Fora same cage geometry, ie. supercage in faujasite, Tables 2 and 3 show changes which cannot be ascribed to shape selectivity, and which should therefore describe the in- fluence of chemical factors. The results of the two dealuminated Y zeolites with framework SYAI ratios of 4.5 (Table 2) confirm that the Ca/Cq ratio is not related to the extent of conversion i. to the acidity which is very different for the two samples since the LZY-82 zeolite contains extraframework Al species. The similar Ca/C4 ratios clearly show that ‘what is important is the Al content, i.e., the number of charges in the lattice. The rise in Ca/C4 as the Al content deceases may be explained as an increase in localized disturb- ing influence of the field gradient on the molecule (see Figure 2). This would weaken C-Chonds at the end of the molecule, allowing Table 2. CHANGE IN C-C SELECTIVITY WITH AL CONTENT IN FAUJASITE IN THE CRACKING OF N-DECANE Activity Zeolte SiAlp@) suAtr@) moles! Ca/Ca HNax-a1) 42 12 2 4.03 H-NaY-41 24 24 43 1.33 H.Nay-87 24 24 347 1.32 HNaY-D4.5" 45 45 390 222 45 24 2600 2.23 H.NayD7) 7 7 sat 2.86 (@) Framework (F) ortotal ()Al (0) Cation exchange for proton (%) (c) Dealumination with EDTA (¢) Dealumination with (NH4)2SiF6, 102. Barthomeuf, C. Mirodatos Figure 2. Scheme of an adsorbed molecule on a) flat surface (1) or cavity (I) with a high charge density, ie. a rather homogeneous field\tow field gradient through the molecule) b) flat surface (I) or cavity (ll) with a discrete source of high field (Isolated Al st, polarizing cation...)(high field gradient). light products formation. The same explana- tions valid for the resulis of Table 3 Indicating a parallel change inthe C-C splitting selectivity (CalCa) and the polarizing power ofthe cation with no correlation with the reaction rato. The total number of hydrocarbon molecules adsorbed in NaX per supercage is 8.5 or n-C7, 3.05 for nCg and 2.83 for .Cg (17). It should be very close in Y zeolite. These low valves indicate that the molecules have to ‘accommodate to the cages. They may curve themselves as has been suggested for n- decane (18). Such a crowding should also occur in the other zeolite structures. studied here which implies strong zeolite - adsorbate interactions. In the equation expressing the energy of adsorption, the dispersion energy (coupling between instantaneous induced dipoles and quadruples of the adsorbate and ofthe adsorbent) has been shown to increase strongly with the surface curvature (19). For instance, an eightfold increase was calculated between a molecule on a flat surface or one tightly encaged ina spheric cavity. Experimen- tal determinations indicate a true effect of the cage curvature (20.21). What is presented here suggests that in addition to this curvature effect, the proper distribution of the charges, Positive and negative, creates a specific field and field gradient (Figure 2a, 2b). In a rather homogeneous field the molecule would be- Reaction Pathways in n-Paraffin Cracking 103 Table 3 (CHANGE IN C3/C4 RATIO IN THE CRACKING OF n-DECANE Activity Zeotite(@) Mol. st gt CaiCq Zeit? Hay «3 133 ww HLaY 650 1.76 2.27 H.MgY 347 2.00 476 (@) 33 to 41 percent cation exchange for protons have as if in gas or liquid phase, and the classical beta-scission mechanism would cour (Figure 2a). This will give no products smaller than Cg. In a heterogeneous field with Isolated sources of high field (Figure 2b) a part of the molecule Is strongly attracted and the other end ofthe chain s notdisturbed. The C-C bonds close to the charge would be polarized and broken more easily, for instance through @ pentacarbonium ion mechanism (5). This could explain the formation of specific light products reported in Tables 2 and 3. In summary, the fields and field gradient exist in cages. The present approach Is an attempt to quantify their directing influence on the selectivity for the cracking of molecules. In addition, it appears possible to consider that they would affect the activation of bonds near the end of a molecule not only for C-C or CH but possibly also or other bonds like C-O and. CN. Literature Cited (1) Chen, N.Y., Garwood, W.E., J. Catal, 52, 453 (1978). (2) Chen, NY, Luck, S.J, Mower, E.B., J. Catal., 13, 329 (1969). @) (4) Martens, J.A., Tielen, M., Jacobs, P.A., feolites, 4, 98 (1984). les, AV. In “Perspectives in Molecular Sieve Science" (Flank, W.H., Whyte, T.E., eds.), ACS Symposium Series, 968, 642 (1988) Heag, W.O., Dessau, R.M., Proceed. 8th Intern. Cong. Catal. (Bertin, 1984), Ver- lag Chemie, Weinheim, 2, 305 (1984), Corma, A., Monton, J.B., Orchilles, AV., Appl. Catal., 16, 59'(1985); Ind. Eng. Chem. Prod. Res. Divs, 23, 404 (1984). Mirodatos, C., Barthomeuf, D., J. Catal., 93, 246 (1985) and 114, 121 (1988); 6) 6 ” Fomes, V., Martinez, A, Orchil- @) 9) (10) (it) (2) (13) (14) (15) ) a) (18) (19) (20) et) AICHE Spring Meeting, Zeolte Sym- posium, 1986, paper 46b (microfilm). Mirodatos, C., Bioul, A., Barthomeul, D., dl, Chem, Soc. Chem. Comm., 149 (1987). Giamnetto, G., Sansare, S., Guisnet, M.. J. Chem. Soc. Chem. Comm., 1302 (1986). Barrer, LM, in “Zeolites and Clay Minerals", Academic Press, New York, 1978. Rabo, J.A., Pickert, P.E., Stamires, D.N., Boyle, J.E., Proceed. Intern. Cong. Catal., (2nd Paris 1980), Technip, Paris, 2055 (1961). Pickert, P.E., Rabo, J.A., Dempsey, E., ‘Schomaker, E., Proceed. Int. Cong, Catal., (ard, Amsterdam 1964), North Holland, Amsterdam, 714 (1965). Mortier, W.J., Proceed. Intern. Zeol. Cont, (Reno, 1983), Butterworths, Gull- ford, p. 734. Van Genechten, K.A,, Mortier, W.J., Zeoltes, 8, 273 (1988) Gorring, R.L, J. Catal., 31, 13 (1973). Rollmann, LD., Walsh, D.E., J. Catal 56, 199 (1979) Breck, D.W., in Zeolite Molecular Sieves, Wiley and Sons, New York, 1974, p. 430. Fiedler, K., Lohse, V., Sauer, H,, Thamm, H., Schirmer, W., in "Proceeding Sth Inter, Zeolite Cont.” (LV. Rees, ed.), Heyden, London, 490 (1980). DeBoer, J.H., Custers, JF.H., Zeit. Phys. Chem., B25, 225 (1934). Barrer, F.M., J. Colloid Intert. Scl., 21, 415 (1968) Derouane, E. 200 (1987) Stach, Chem, Phys. Lett., 142, image not available 106 W.C. Cheng, A.W. Peters INTRODUCTION Hydrogen transfer reactions alter FCC product distribution by converting olefins and naphthenes into paraffins and aromatics. These reactions have a net effect of decreasing gasoline octane with faujasite catalysts. In an earlier study, the conversion of cyclohexene on Y-zeolites was shown to be a good test reaction for hydrogen transfer [1]. The ratio of the products of isomerization to the products of hydrogen transfer of the cyclohexene reaction increased with increasing SiO2/Al20s ratio of the zeolite. This trend correlated well with the gasoline octane numbers from gas oil cracking. The effect of zeolite structure on the disproportioonation and alkylation of toluene and isomerization of xylenes have been examined previously (2- 4). In the present study, we have used the reaction of cyclohexene to study the hydrogen transfer properties of zeolites with high silica to alumina ratio. Hydrothermally dealuminated zeolite Y, zeolite beta and ZSM-5 with SiO2z/AlOs greater than 20 were used in this study. The effect of zeolite structure at constant SiOz/AlzOs as well as the effect of SiO2/Al2Os for a given zeolite structure were studied. Hydrogen Transfer on Zeolites with High SV/A1 Ratios 107 EXPERIMENTAL METHODS ‘The REUSY was prepared by rare earth exchange of a commercially available Z-14US zeolite (Davison Chemical Division, W. R. Grace & Co.-Conn.) using a procedure described by Rajagopalan and Peters [5]. Zeolite beta was prepared using the procedure of Wadlinger et al. (6], exchanged with ammonium sulfate and water washed. ZSM-5 of Si02/Al203 ratio between 23 and 65 were prepared following the procedure of Argauer and Landolt [7] exchanged in dilute H2SO« and water washed. X-ray analysis show the zeolite beta and ZSM-5's to be phase pure. Analysis by 27Al NMR showthe Alin zeolite beta and ZSM-5 to. be essentially all tetrahedral. The zeolites were mixed with an alumina sol and kaolin clay diluent ina high shear mixer and either spray dried or oven dried and sieved to between 100 and 325 mesh. The REUSY catalyst was steamed in a fluidized bed at 1088 K and a steam pressure of 101 kPa for 4 hours to further dealuminate the zeolite. The catalysts containing zeolite beta and ZSM-5 were not steamed, but only calcined in air at 810 K for two hours. It was assumed that under these conditions the zeolites did not dealuminate significantly. Properties of tho REUSY and the finished REUSY catalysts are shown in Table 1. The unit cell size of the zeolite after steaming is 2431 pm, which corresponds to a SiO2/Al2Os of 26.4 and a Al site density of 13.8 Al/unit cell, as calculated by the correlation of Breck and Flanigen [8], or a site density of 8.7, as calculated by the correlation of Fichtner- Schmittler et al. [9]. The properties of zeolite beta and ZSM-5's are shown in Table 2. The zeolite beta sample has a SiO2/Al2Os of 26.5, which is similar to that of the REUSY sample. In addition, one of the ZSM-5 samples has a Si02/Al203 of 23.0. These three samples will allow us to determine effect of zeolite structure {at comparable SiO2/AlzO3) on hydrogen transfer. Furthermore, ZSM-5 samples 108 W.C. Cheng, A.W. Peters with SiO2/Al2Oa of 23 to 65 will allow us to determine the effect of Al site density of ZSM-5 on hydrogen transfer. Cyclohexene reactions were carried out using a modified fixed-bed ASTM. D-3907-80 microactivity procedure. Details of the operation have been described earier [1]. The product mass balance was typically > 95%, and was normalized for reporting. RESULTS AND DISCUSSION Effect of Zeolite Structure The product distribution (Table 3) consisted of a hydrogenated product (cyclohexane), isomerized products (1-, 2- and 3- methylcyclopentene), isomerized and hydrogenated product (methylcyclopentane) and dimerized C1zs. Minor amounts of benzene, toluene, dimethyicyclopentane, dimethyicyclopentene, methylcyclohexane and methylcyclohexene were observed. The amount of dehydrogenated products, benzene and toluene, are much lower than the hydrogenated products, cyclohexane and methylcyclohexane. In order to satisty hydrogen balance, some of the hydrogen in the hydrogenated products must have come from the coke and C12’s fractions. The product yields of the REUSY, zeolite beta and ZSM-5 catalysts having comparable zeolite SiO2/Al2Os ratio are shown in Figures 1-5. Cyclohexane appears to be a primary stable product, as the yield curves are linear and pass through the origin. The yield of cyclohexane is highest for REUSY followed by beta and then ZSM-5. Since the SiO2/Al20s ratios are similar, the difference in selectivity must be attributed to zeolite structure. Faujasite has a 3 dimensional pore system with 7.4 A pores and large cavities at the intersections of the pores. The structure of zeolite beta has recently been identified by Treacy and Newsam [10] and Higgins et al. [11]. Its pore system consists of two orthogonal sets of linear elliptical channels of 7.5 by 5.7 Aand a set of tortuous channels having the dimensions of 6.5 by 5.6 A. Although the Hydrogen Transfer on Zeolites with High SWAIRatios ‘109 channels are relatively large in size, large cavities are not formed at pore intersections. ZSM-5 has a 2-dimensional pore system, consisting of linear channels of 5.3 by 5.6 A and sinusoidal channels of 5.1 by 5.5 A. Hydrogen transfer, being a bimolecular reaction, would be expected to involve a rather bulky transition state. ZSM-5, having smaller pores and beta, not having large cavities ‘would limit the rate of this reaction [12]. Additionally, the extra framework aluminum in hydrothermally dealuminated REUSY may contribute to hydrogen transfer. Methylcyclopentenes are also a primary product. They appear to be stable for ZSM-5 and zeolite beta, but unstable at high levels for REUSY. The yield of methylcyclopentene is highest for ZSM-5 and lowest for REUSY. This follows exactly the opposite ranking as that for the yield of cyclohexane. Olefin isomerization, being a unimolecular reaction [13] does not involve a large transition state, and therefore is not shape-selective. The high selectivity of ZSM-5 for isomerization could be attributed to the fact that most other reactions, such hydrogen transfer, dimerization and coking are suppressed. Alternatively, the acid strength of the sites on ZSM-5 could be stronger, giving them a higher activity for isomerization. The yield of methylcyciopentane is shown in Figure 3. Interestingly, methyleyclopentane appears to be a primary product on beta and a primary plus secondary product on ZSM-5 and REUSY. This suggests the possibility that the formation of methylcyclopentane on beta occurs via a different mechanism [14] ‘At low conversion, beta makes more methylcyclopentane than REUSY, however, at higher conversion the yield curves cross. The tendency to form dimerized products follow the same trend as hydrogen transfer. The low dimerization activity of ZSM-5 is attributed to both transition state shape selectivity and product shape selectivity [9]. The lower NO W.C. Cheng, A.W. Peters dimerization activity of zeolite beta as compared to faujasite is primarily attributed to transition state shape selectivity. Zeolite beta made the most carbonaceous deposit followed by REUSY and ZSM-5. The high coke forming tendency of zeolite beta compared to HY has been reported earlier [15]. The summary of the product distribution at 20% conversion is shown in Table 3. ZSM-6 is the most active for isomerization, producing roughly twice as much methylcyclopentenes as beta. REUSY is the least active for isomerization. The hydrogen transfer activity of ZSM-5 is low, as witnessed by the low levels of saturates. Following the work of Magnoux et al. [16], we will define hydrogen transfer products as the sum of cyclohexane and methylcyclopentane and isomerization products as the sum of the three methylcyclopentene isomers and methylcyclopentane. Using this comparison, zeolite beta produces ca. 25% more isomerization products and ca 25% less hydrogen transfer products than REUSY. Itis interesting to also note that on REUSY the yield of methycyclopentane is lower than that of cyclohexane, while on beta the yield of methyicyclopentane is higher than that of cyclohexane. This again demonstrates the higher isomerization activity of beta. The ratio of hydrogen transfer to isomerization is a measure of the relative importance of bimolecular to unimolecular processes. This parameter is the highest for REUSY, followed by beta and very much lower for ZSM-5. Effect of SiO2/Al203 Content in ZSM-5 The product distributions of cyclohexene reaction over ZSM-5 of varying SiO2/AI203 ratio are shown in Figures 6 - 10. The hydrogen transfer activity decreases and the isomerization activity increases as the SiO2/Al20s ratio of ZSM-5 increases from 23 to 43. This is expecied since a lower Al site density favors unimolecular reactions (isomerization) over bimolecular reactions Hydrogen Transfer on Zeolites with High Si/Al Ratios att (hydrogen transfer). Similar results were observed for Y-zeolites [1]. However, the product distribution and the ratio of hydrogen transfer products to isomerization products do not change as the SiOz/AlzOs further increases from 43 to 65. In this case the distance of site separation is great enough that further dealumination does not affect the local site density and site strength. The trend in C12's production follows that for hydrogen transfer. The yield of carbonaceous deposit increases with Al site density. CONCLUSIONS The reaction of cyclohexene was used to probe the effect of zeolite structure on the hydrogen transfer characteristics of faujasite, zeolite beta and ZSM-5. At the same silica to alumina ratio, the hydrogen transfer activity increased in the order ZSM-5 < zeolite beta < favjasite, while the isomerization activity followed the reverse order. The lower hydrogen transfer activity of the ZSM-5 and zeolite beta compared with faujasite can be explained by transition state shape selectivity and contribution from the extra-framework alumina of the steam-dealuminated faujasite. Over ZSM-5, the hydrogen transfer activity decreased as SiOz/Al20s increased from 23 to 43. This is consistent with the notion that hydrogen transfer increases with Al site density and hence decreases with increased in SiO2/Al2Os. The product distribution remained the same as ‘SiO2/Al20s increased further to 65. This indicates that the lower imit on hydrogen transfer rate is achieved at the Al site density corresponding to SiQ2/AlzOs of about 43. Further reduction in site density did not reduce hydrogen transfer rate. ACKNOWLEDGEMENTS: The authors are grateful to A. L. Wadsworth and C. R. Petr for the very capable laboratory assistance and to the Davison Chemical Division of W. R. Grace & Co.-Conn. for permission to publish this work. 2 W.C. Cheng, A.W. Peters LITERATURE CITED Cheng, W., and Rajagopalan, K., J. Catal. 119, 354 (1989). Ratnasamy, P., Bhat, R. N., Pokhriyal, S. K,, Hegde, S. G. and Kumar, R., v. Catal., 119, 65 (1989). Babu, G. P., Santra, M., Shiralkar, V. P. and Ratnasamy, P., J. Catal. 100, 458 (1986). Datka, J., Piwowarska, Z., Rakoczy, J. and Suikowski, B. Zeolites, 8, 199 (1988). Rajagopalan, K, and Peters, A. W., J. Catal. , 106, 410 (1987). Wadlinger, R. L., Kerr, G. T., Rosinski, E. J., U.S. Patent 3308069 (1967). Argauer, R. J. and Landolt, G. R., U.S. Patent 3702886 (1972). Breck, D. W. and Flanigen, E.M.,in “Molecular Sieves” (R. M. Barrer, Ed.), p. 47. Society of Chemical Industry, London, 1968. Fichtner - Schmittler, H. Lohse, U., Engelhardt, G. and Patzelova, V., Cryst. Res. Technol. 19 (1), K1 (1984). Treacy, M. M. J. and Newsam, J. M., Nature, 332:17, 249 (1988). . Higgins, J. B., LaPierre, R. B., Schlenker, J. L., Rohrman, A. C., Woods, J. D., Kerr, G. T. and Rohrbaugh, W. J., Zeolites, 8, 446 (1988). Hydrogen Transfer on Zeolites with High Si/Al Ratios 113 . Haag, W. O., Lago, R. M. and Weisz, Faraday Soc., 72, 317 (1981). . Chen, N. Y. and Haag, W. O., in “Hydrogen Effects in Catalysis Fundamentals and Practices” (Paal, Z. and P. G. Menon, Eds.), p. 695. Marcel Dekker, Inc., New York and Basel, 1988. . Peters, A. W., Cheng, W-C., Shatlock, M., Wormsbecher, R. F. and Habib, Jr., E. T., in “Physicochemical Properties of Zeolitic Systems and Their Low Dimensionaiity”, D. Barthomeuf, E. G. Derouane and W. Holderich, eds., NATO ASI Series, Plenum Press, NY (1989). . Corma, A., Fornes, V., Melo, F. and Perez-Pariente, J. in “Fluid Catalytic Cracking Role in Modern Refining” (Occelli, Ed.) ACS Symposium Series 375, p. 49 (1988). ; Magnoux, P., Gallet, A. and Guisnet, M. Bull. Soc. Chim. Fr. #5, 810 (1987). 114 W.C. Cheng, A.W. Peters Table 1: Properties of REUSY Catalyst. Zeolite Properties RE/Urit Cell 4.2 wt% Zeolite in Catalyst 27 Catalyst Properties, wt% Ales 42.6 wi% E203 1.77 wt% NazO 0.19 BET Surface Area/m?g” 200 Nitrogen Pore Volume/omg 0.21 Bulk Density/gem? 0.78 Average Particle size/um 54 Atter Hydrothermal Treatment BET Surface Area/m?gt 128 Unit Coll Constantipm 2431 1SiO2/Al209 26.4 *AVUnit Cell 13.8 2AvUnit Cell 87 ‘Calculated as 1.152 x (unit cell - 2419) [5] 2Calculated as 1.124 x (unit cell - 2423) [6] Hydrogen Transfer on Zeolites with High Si/Al Ratios Table 2: Properties of Zeolite Beta and ZSM-5's Beta ZSM-5-23 ZSM5-43_ZSM-5-65 Zagiite Properties wt%NazO 0.023, 0.021 0.019 0.023, wi% Al203 6.01 6.82 3.70 2.55 wt%SiO2 93.7 92.5 937 97.3 wt% SiO2/Al203 265 23.0 43.3 65.0 BET/m2g* 623 420 412 354 Ne Pore Volume/em3g 0.34 0.21 0.22 0.26 wt% zeolite in catalyst 40 15, 15 15 Table 3: Product Distribution at 20% Cyclohexene Conversion. REUSY Bela ZSM-5 mcyCs" 236 29.7 58.0 mCyCs 67 75 34 CyCs 103 55 20 Crs 37.9 22.6 18.6 Coke 18.5 31.2 15.6 Light Gases 18 25 44 Miso. 15 1.0 1.0 Hydrogen Transfer 17.0 13.0 54 Isomerization 303 37.2 61.4 HT/Asomerization 0.56 0.35 0.088 us 6 W.C. Cheng, A.W. Peters w ¥ + * é £ x c & 9 4 E x A $ N E Le — 10 20 30 WT% CYCLOHEXENE CONVERSION Figure 1. Cyclohexane yield as a function of cyclohexene conversion. REUSY (+), Beta( *), ZSM5 (0). 16 14 12 10 mzmazmvoro4zmv0ro

You might also like