Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-26T14:41:14.042Z Has data issue: false hasContentIssue false

Synthesis of Highly Porous Aluminas Mediated by Cationic Surfactant: Structural and Textural Properties

Published online by Cambridge University Press:  03 March 2011

J.A. Toledo
Affiliation:
Instituto Méxicano del Petróleo, Prog. de Ingeniería Molecular, Eje Central Lázaro Cárdenas # 152, A.P. 07730 México, D.F. México
X. Bokhimi
Affiliation:
Institute of Physics, The National University of Mexico, A.P. 20-364, 01000 México D.F., Mexico
C. Lopez
Affiliation:
Instituto Méxicano del Petróleo, Prog. de Ingeniería Molecular, Eje Central Lázaro Cárdenas # 152, A.P. 07730 México, D.F. México
C. Angeles
Affiliation:
Instituto Méxicano del Petróleo, Prog. de Ingeniería Molecular, Eje Central Lázaro Cárdenas # 152, A.P. 07730 México, D.F. México
F. Hernandez
Affiliation:
Instituto Méxicano del Petróleo, Prog. de Ingeniería Molecular, Eje Central Lázaro Cárdenas # 152, A.P. 07730 México, D.F. México
J.J. Fripiat
Affiliation:
Instituto Méxicano del Petróleo, Prog. de Ingeniería Molecular, Eje Central Lázaro Cárdenas # 152, A.P. 07730 México, D.F. México
Get access

Abstract

The cationic surfactant cetyltrimetylammonium bromide was used to synthesize mesostructured γ–Al2O3. The effects of the surfactant concentration and of the sol aging (at 95 °C for 24 h) were studied by x-ray powder diffraction, nuclear magnetic resonance, transmission electron microscopy, and analysis of the low-temperature nitrogen adsorption-desorption isotherms. Mesostructured alumina with wormhole morphology and amorphous walls was obtained through the precipitation by ammonium hydroxide of a 0.1 M aluminum nitrate aqueous solution in presence of 0.1 M surfactant. The pore size was smaller than 5 nm. After digesting the milky suspension under atmospheric pressure at 95 °C, a crystallized boehmite-surfactant phase, with fiber morphology, is formed which at 550 and 700 °C is transformed into a highly porous γ−Al2O3. A similar evolution was observed using 0.01 M CTAB solution and aging. Pore volume up to 1.1 cm3/g and pore size up to 16 nm were obtained. Without surfactant, the same aging treatment led to aggregated fibers: the pore size is less than 8 nm and the pore volume is smaller than 0.6 cm3/g. The γ-alumina surface area is determined mainly by the organization generated by the surfactant and to a lesser extent by the boehmite precursor particle size. From the point of view of catalyst preparation, the surfactant at the concentration of 0.01 M in 0.1 M aluminum nitrate and the aging treatment in solution play a beneficial role.

Type
Articles
Copyright
Copyright © Materials Research Society 2005

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

1Grange, P.: Catalytic hydrodesulfurization. Cata. Rev.-Sci. Eng. 21, 135 (1980).CrossRefGoogle Scholar
2Gates, B.C.: Supported metal clusters: Synthesis, structure, and catalysis. Chem. Rev. 95, 511 (1995).CrossRefGoogle Scholar
3Xu, Z., Xiao, F-S., Purnell, S.K., Alexeev, O., Kawai, S., Deutsch, S.E. and Gates, B.C.: Size-dependent catalytic activity of supported metal clusters. Nature 372, 346 (1994).CrossRefGoogle Scholar
4Kubicek, N., Vaudry, F., Chiche, B.H., Hudec, P., Di Renzo, F., Schulz, P. and Fajula, F.: Stabilization of zeolite beta for fcc application by embedding in amorphous matrix. Appl. Catal. A 175, 159 (1998).CrossRefGoogle Scholar
5Furimsky, E. and Massoth, F.E.: Deactivation of hydroprocessing catalysts. Catal. Today 52, 381 (1999).CrossRefGoogle Scholar
6De Jong, A.M., De Beer, V.H.J., Van-Veen, J.A.R. and Niemantsverdriet, H.W.: Surface science model of a working cobalt-promoted molybdenum sulfide hydrodesulfurization catalyst: Characterization and reactivity. J. Phys. Chem. B 100, 17722 (1996).CrossRefGoogle Scholar
7Nielsen, L.P., Christensen, S.V., Topsoe, H. and Clausen, B.S.: Changes in metal-sulfur bond energy in promoted and unpromoted molybdenum catalysts. Catal. Lett. 67, 81 (2000).CrossRefGoogle Scholar
8Sharma, L.D., Kumar, M., Saxena, A.K., Chand, M. and Grupta, J.K.: Influence of pore size distribution on Pt dispersion in Pt–Sn/Al2O3 reforming catalyst. J. Mol. Catal. A 185, 135 (2002).CrossRefGoogle Scholar
9Farbotko, J.M., Garin, F., Girard, P. and Maire, G.: Reactions of labeled hexanes on Pt–WO3/Al2O3 catalysts. J. Catal. 139, 256 (1993).CrossRefGoogle Scholar
10Hua, W. and Sommer, J.: Alumina-doped Pt/WOx/ZrO2 catalysts for n-heptane isomerization. Appl. Catal. A 232, 129 (2002).CrossRefGoogle Scholar
11Trombetta, M., Busca, G., Rossini, S.A., Piccoli, V. and Cornaro, U.: FT-IR studies on light olefin skeletal isomerization catalysis: I. The interaction of C4 olefins and alcohols with pure γ-alumina. J. Catal. 168, 334 (1997).CrossRefGoogle Scholar
12Elmasides, C., Kondarides, D.I., Grünert, W. and Verykios, X.E.: XPS and FTIR study of Ru/Al2O3 and Ru/TiO2 catalysts: Reduction characteristics and interaction with a methane-oxygen mixture. J. Phys. Chem. B 103, 5227 (1999).CrossRefGoogle Scholar
13Taylor, K.C.: Nitric-oxide catalysis in automotive exhaust systems. Cat. Rev.-Sci. Eng. 35, 457 (1993).CrossRefGoogle Scholar
14Lerner, B.A.: U.S. Patent No. 5 559 067 (1996).Google Scholar
15Cheng, W-Ch. and Rajagopalan, K.: U.S. Patent No. 5 547 564 (1966).CrossRefGoogle Scholar
16Bokhimi, X., Toledo-Antonio, J.A., Guzman-Castillo, M.L., Mar, B. Mar, Hernandez, F. and Navarrete, J.: Dependence of boehmite thermal evolution on its atom bond lengths and crystallite size. J. Solid State Chem. 161, 319 (2001).CrossRefGoogle Scholar
17Tsukada, T., Segawa, H., Yasumori, A. and Okada, K.: Crystallinity of boehmite and its effect on the phase transition temperature of alumina. J. Mater. Chem. 9, 549 (1999).CrossRefGoogle Scholar
18Vaudry, F., Khodabandeh, S. and Davis, M.E.: Synthesis of pure alumina mesoporous materials. Chem. Mater. 8, 1451 (1996).CrossRefGoogle Scholar
19Bokhimi, X., Sanchez-Valente, J. and Pedraza, F.: Crystallization of sol-gel boehmite via hydrothermal annealing. J. Solid State Chem. 166, 182 (2002).CrossRefGoogle Scholar
20Zhu, H.Y., Riches, J.D. and Barry, J.C.: Gamma-alumina nanofibers prepared from aluminum hydrate with poly(ethylene oxide) surfactant. Chem. Mater. 14, 2086 (2002).CrossRefGoogle Scholar
21Bangshaw, S.A. and Pinnavaia, J.T.: Mesoporous alumina molecular sieves. Angew. Chem. Int. Ed. Engl. 35, 1102 (1996).CrossRefGoogle Scholar
22Cabrera, S., Haskouri, J.E., Alamo, J., Beltran, A., Beltran, D., Mendioroz, S., Marcos, M.D. and Amoros, P.: Surfactant-assisted synthesis of mesoporous alumina showing continuously adjustable pore sizes. Adv. Mater. 11, 379 (1999).3.0.CO;2-6>CrossRefGoogle Scholar
23Yada, M., Ohya, M., Machida, M. and Kijima, T.: Synthesis of porous yttrium aluminium oxide templated by dodecyl sulfate assemblies. Chem. Commun. 18, 1941 (1998).CrossRefGoogle Scholar
24Zhang, Z., Hicks, R.W., Pauly, T.R. and Pinnavaia, T.J.: Mesostructured forms of gamma-Al2O3. J. Am. Chem. Soc. Commun. 124, 1592 (2002).CrossRefGoogle Scholar
25Zhang, Z. and Pinnavaia, T.J.: Mesostructured gamma-Al2O3 with a lathlike framework morphology. J. Am. Chem. Soc. 124, 12294 (2002).CrossRefGoogle Scholar
26Lee, H.Ch., Kim, H.J., Chung, S.H., Lee, K.H., Lee, H.Ch. and Lee, J.S.: Synthesis of unidirectional alumina nanostructures without added organic solvents. J. Am. Chem. Soc. 125, 2882 (2003).CrossRefGoogle ScholarPubMed
27Rodríguez-Carbajal, J.: Laboratoire Leon Brillouin (CEA-CNRS), France .Google Scholar
28Thompson, P., Cox, D.E. and Hasting, J.B.: Rietveld refinement of Debye-Scherrer synchrotron x-ray data from Al2O3. J. Appl. Crystallogr. 20, 79 (1987).CrossRefGoogle Scholar
29Coster, D.J. and Fripiat, J.J.: Memory effects in gel solid transformations - coordinately unsaturated Al sites in nanosized aluminas. Chem. Mater 5, 1204 (1993).CrossRefGoogle Scholar
30Engelhardt, G. and Michel, D.: High-Resolution Solid-State NMR of Silicates and Zeolites (John Wiley and Sons, New York, 1987).Google Scholar
31Bottero, J.Y., Axelos, M., Tchoubar, D., Cases, J.M., Fripiat, J.J. and Fiessinger, F.: Mechanism of formation of aluminum trihydroxide from Keggin Al13 polymers. J. Colloid Interface Sci. 117, 47 (1987).CrossRefGoogle Scholar
32Sanchez-Valente, J., Bokhimi, X., Guzman-Castillo, M.L., Hernandez, F. and Fripiat, J.J.: Quantitative relationships between boehmite and γ-alumina crystallite sizes. J. Mater. Res. 19, 1499 (2004).CrossRefGoogle Scholar
33Hicks, R.W. and Pinnavaia, T.J.: Nanoparticle assembly of mesoporous AlOOH (boelamite). Chem. Mater. 15, 78 (2003).CrossRefGoogle Scholar
34Galarneau, A., Desplantier, D., Dutartre, R. and Di Renzo, F.: Micelletemplated silicates as a test bed for methods of mesopore size evaluation. Microporous Mesoporous Mater. 27, 297 (1999).CrossRefGoogle Scholar
35Pacheco, G., Zhao, E., Garcia, A., Sklyarov, A. and Fripiat, J.J.: Synthesis of mesoporous zirconia with anionic surfactants. J. Mater. Chem. 8, 219 (1998).CrossRefGoogle Scholar
36Blumenfeld, A.L. and Fripiat, J.J.: 27Al-1H REDOR NMR and 27Al spin-echo editing: A new way to characterize Brönsted and Lewis acidity in zeolites. J. Phys. Chem. B 101, 6670 (1997).CrossRefGoogle Scholar
37Blumenfeld, A.L. and Fripiat, J.J.: Characterization of Brønsted and Lewis acidity in zeolites by solid-state NMR and the recent progress in the REDOR technique. Magn. Reson. Chem. 37, S118 (1999).3.0.CO;2-K>CrossRefGoogle Scholar
38Coster, D.J., Blumenfeld, A.L. and Fripiat, J.J.: Lewis-acid sites and surface aluminum in aluminas and zeolites: A high-resolution NMR-study. J. Phys. Chem. B 98, 6207 (1994).Google Scholar