Structure Analysis of Si-Atom Pillared Lamellar Silicates Having

Feb 4, 2010 - Chinese Journal of Chemistry 2018 36 (3), 227-232 ... Chinese Chemical Letters 2014 , ... Two-dimensional zeolites: dream or reality? Wi...
0 downloads 0 Views 4MB Size
3466

J. Phys. Chem. C 2010, 114, 3466–3476

Structure Analysis of Si-Atom Pillared Lamellar Silicates Having Micropore Structure by Powder X-ray Diffraction Takuji Ikeda,*,† Syunsuke Kayamori,† Yasunori Oumi,‡ and Fujio Mizukami† Research Center for Compact Chemical Process, National Institute of AdVanced Industrial Science and Technology (AIST), 4-2-1 Nigatake, Sendai, Japan, and Department of Engineering, Hiroshima UniVersity, 1-4-1 Kagamiya, Higashi-Hiroshima, Japan ReceiVed: December 21, 2009

Four kinds of Si-atom pillared lamellar silicates (APZ-1-APZ-4) were prepared by thermal acid treatment in an autoclave at 443 K for 1-24 h using aqueous HCl. Layered silicates PLS-1, PLS-3, PLS-4, and PREFER, which consist of ferrierite layer sheets, could be converted into new open-framework microporous materials by pillaring the interlayers with SiO2(-OH)2 fragments. The fragments probably originated from the collapse of the crystalline structure, and interlayer silylation occurred with the migration of the fragments (dissolution-condensation). The four APZ crystal structures were investigated by means of powder X-ray diffraction, revealing topotactic microporous structures that derived from each original layered silicate. APZ-2 and APZ-4 have the large pore openings of a 12-membered ring, whereas APZ-1 and APZ-3 have the slightly smaller pore openings of a 10-membered ring. All the APZs are hydrophilic to approximately the same degree as the common hydrophilic zeolite, owing to hydroxyl groups of the Si atom pillar or the terminal silanols in the interlayer. Thus, large numbers of water molecules are adsorbed in the micropores. 29Si MAS NMR spectra showed Q2, Q3, and Q4 resonance peaks, whose Q2 signal indicated silanol groups of the Si atom pillar. All the APZ crystals showed high thermal stability and did not collapse upon heating to at least 873 K. 1. Introduction Layered silicates are potential candidates for various applications such as catalysis, adsorption, and separation and are themselves widely used as silica sources for the synthesis of other silicates. For the past several years, the topotactic conversion of a crystalline layered silicate into a novel highsilica zeolite has attracted attention as a new zeolite synthesis method. It has been more than one decade since the first report on the structural transformation of layered silicates PREFER1 and MCM-22(P)2 (isomorphic material: ERB-1,3 PSH-3,4 ITQ1,5 SSZ-256) into FER- and MWW-type zeolites, respectively. Subsequently, the synthesis of novel zeolites by topotactic conversion via a phase transition was attempted by many researchers. To date, zeolites such as CDO,7 NSI,8 CAS-NSI,9 RWR,10-12 and RRO13 have been prepared by using layered silicates PLS (Pentagonal cylinder Layered Silicate)-17 (isomorphic materials: PLS-4,14 RUB-36,15,16 MCM-47,17 MCM65,18 UZM-13,19 UZM-17,19 UZM-1919), Nu-6(1),8 EU-19,20,21 RUB-18,22 and RUB-39,23 respectively. In these zeolites, micropores are formed in the interlayer by dehydration condensation of terminal silanol groups facing each other. These novel zeolites typically have a one- or two-dimensional (2D) pore structure with an eight- or ten-membered ring (MR). Recently, Fan et al.24 and Ruan et al.25 reported the interesting structural change of Ti-MWW zeolite to Ti-YNU-1. From HR-TEM observations,25 they suggested that a new T-site, intercalated between the layers, might connect adjacent MWW silicate layers as a Q2-local structure by reflux in HNO3 solution. * Corresponding author. Tel.: +81-22-237-3016. Fax: +81-22-237-5217. E-mail: [email protected]. † National Institute of Advanced Industrial Science and Technology. ‡ Hiroshima University.

Ti-YNU-1 has 12-MR pore openings, which are larger than the 10-MR openings of Ti-MWW. Subsequently, Inagaki et al.26 prepared the organic-inorganic hybrid zeolite IEZ (Interlayer Expanded Zeolite)-1 by interlayer silylation of the layered silicate PLS-1.7 In IEZ-1, neighboring silicate layers are bridged by SiO2(CH3)2 fragments by an intentional silylation with dichlorodimethylsilane [Si(CH3)2Cl2] as an additive Si source. An open-framework 2D microporous structure with a 10-MR opening was formed in the interlayer, whose structural model was predicted by the change in the basal spacing in the X-ray diffraction (XRD) pattern. Furthermore, the methyl groups of the fragment are exchanged for hydroxyl groups by calcination only, yielding IEZ-2. The same interlayer expansion technique was investigated further using various layered silicates by Wu et al.27 The application of this silylation technique to various layered silicates has the potential to make new functional microporous materials. Mochizuki et al.28-30 and Ishii et al.31,32 reported an artful bridge formation by silylation with alkoxychlorosilanes on well-known layered silicates (octosilicate, magadiite, and kenyaite). In their work, large flexible micropores were formed in the interlayers by pillaring with bulky organic molecules, while their crystal structures appeared disorderly owing to incomplete bridging by soft molecule structures. More recently, Ruan et al. reported the IEZ-FER with 12-MR openings prepared by interlayer silylation as well as IEZ-1, whose crystal structure was elucidated by the HR-TEM observations.33 Both IEZ-1 and IEZ-FER consist of the same ferrierite layer with a different layer stacking sequence. In both IEZ-2 and IEZ-FER, a Q2 local environment (dO2Si(-OH)2) was observed in the 29Si MAS NMR, with the signal originating from hydroxyl groups in the Q2 bridge. However, the crystal structures of these new interlayer expanded porous materials in air have to date not been

10.1021/jp912026n  2010 American Chemical Society Published on Web 02/04/2010

Analysis of Si-Atom Pillared Lamellar Silicates

J. Phys. Chem. C, Vol. 114, No. 8, 2010 3467

TABLE 1: XRD Experimental Conditions and Crystallographic Data for APZ-1, APZ-2, APZ-3, and APZ-4 compound name

APZ-1

APZ-2

APZ-3

APZ-4

estimated chemical formula FW space group

Si18.8O38(OH)1.5 · 6.7(H2O) 1282.3 P21/m (No. 11) 1.09685(8) 1.40153(4) 0.74063(6) 98.307(11) 1.127(2) Cu KR1 5-90 0.017473 7 12 4865 962

Si38.4O76(OH)4.8 · 23.9(H2O) 2806.9 Pnnm (No. 58) 1.40462(9) 0.74172(2) 2.39378(14) 2.4939(2) Cu KR1 5-90 0.017473 7 10 4865 1050

Si39.6O76(OH)7.2 · 16.2(H2O) 2742.7 Pnma (No. 62) 2.3339(2) 1.39582(6) 0.73906(5) 2.4077(3) Cu KR1 5-90 0.017473 6 12 4865 1029

Si38.2O76(OH)4.5 · 24.5(H2O) 2807.0 Pnnm (No. 58) 1.40274(3) 0.741310(13) 2.3900(12) 2.48528(14) Cu KR1 5-90 0.017473 5 14 4865 1044

70

71

71

73

0.0385 0.0114 0.0233

0.0276 0.0100 0.0201

0.0360 0.0087 0.0218

0.0399 0.0097 0.0253

a/nm b/nm c/nm β/° V/nm3 wavelength λ/Å 2θ range/° step size (2θ)/° counting time per step/s profile range in fwhm number of observations number of contributing reflections number of refined structural parameters Rwp RF Re

determined quantitatively by XRD. In previous HR-TEM observations, the expansion of the interlayer in IEZ was evidently suggested by an atomic-scale image; however, detailed distribution of the Q2-pillar and guest atoms or molecules in micropores was not clear. We have attempted to prepare crystalline nanosheet materials by the delamination of a crystalline layered silicate. During a high-temperature acid treatment (aqueous HCl at 170 °C for 0.9 in APZ-1, APZ-2, and APZ-3, suggesting the presence of macropores due to interparticle voids between small crystalline particles. This finding is compatible with the adsorption properties of PLS-3 and PLS-4.14 The micropore diameters were estimated to be 0.51-0.57 nm (Table 3), which is almost consistent with a size of a 10-MR or 12MR pore opening. In APZ-2, two pore diameter sizes were estimated to be 0.51 and 0.57 nm, corresponding to the pore size of a 10-MR and 12-MR, respectively. However, a single pore with a diameter of 0.55 nm was confirmed in APZ-4, despite it having an essentially identical crystal structure, as described later. Moreover, water vapor adsorption measurements revealed interesting adsorption properties of APZs. The observed water vapor adsorption isotherms in Figure 8 indicate high hydrophilicity of APZs, similar to a hydrophilic zeolite. Figure 8 also shows the isotherm of ZSM-5 (Si/Al ≈ 100) for comparison. The surface areas SBET are listed in Table 3. It is likely that the

Analysis of Si-Atom Pillared Lamellar Silicates

Figure 5.

29

J. Phys. Chem. C, Vol. 114, No. 8, 2010 3471

Si DDMAS NMR and 1H-29Si CP/MAS NMR spectra of (a) APZ-1, (b) APZ-2, (c) APZ-3, and (d) APZ-4.

TABLE 2: Relative Intensity Ratios of Q2, Q3, and Q4 Resonance Peaks Estimated by Deconvolution of 29Si DDMAS NMR Spectra Obtained APZ-1 APZ-2 APZ-3 APZ-4

Q2

Q3

Q4 (%)

2.5 2.3 2.4 2.2

22.3 14.3 16.5 22.4

75.2 83.4 81.1 75.4

hydrophilicity of APZs originates from a large number of silanol groups of the Q2 and Q3 sites in the micropores. The estimated value of SBET(H2O) for APZ-1 is large, whereas its SBET(N2) value is the smallest among all APZs. Probably, the large SBET(H2O) of APZ-1 is strongly correlated with the number of silanol groups in the crystal, as supported by the 29Si NMR measurements; viz., (Q2 + Q3)/Q4 for APZ-1 is the highest compared to other APZs. 3.4. Crystal Structure Analysis. Crystal structures of APZs were revealed by the Rietveld analyses; however, those analytical accuracies are not so high owing to crystallinity deterioration by acid treatment (except for APZ-4). Furthermore, the XRD

pattern obtained for APZ-1 suggests the presence, to some degree, of stacking faults in the layer framework because of asymmetric peak profiles with a long tail. We applied the space groups of layered silicate precursors to each APZ, and then the systematic absences of reflections were satisfied basically. At first, unit cells and space groups of layered silicates, APZ, and zeolites obtained by direct calcination of the precursors are summarized in Table 4. Table 4 indicates that lattice constants along the stacking direction of all APZs are larger than those of layered silicates (except PREFER) and related zeolites by direct calcination of layered silicates, indicating that interlayer expansion of layered silicates took place by the simple acid treatment. From preliminary structure analysis,1 the interlayer distance of PREFER is considerably large because a bulky SDA molecule is accommodated in the interlayer. On the other hand, the lattice constant along the layer stacking direction and unitcell volume were remarkably decreased by direct calcination of these layered silicates (i.e., CDO- and FER-type zeolites). Figure 9 shows the crystal structure model for APZ-1, as revealed by the Rietveld refinement. Refined values for the

3472

J. Phys. Chem. C, Vol. 114, No. 8, 2010

Ikeda et al. TABLE 3: BET Specific Surface Area SBET(N2), Micropore Volume Vmicro(N2), and Total Pore Volume Vtotal(N2) of APZs and Directly Calcined Precursors Estimated by Nitrogen Gas Adsorption, Pore Diameter D(Ar) Estimated by Argon Gas Adsorption, and SBET(H2O) Estimated by the Water Vapor Adsorption SBET(N2) Vmicro(N2) Vtotal(N2)

APZ-1 APZ-2 APZ-3 APZ-4 calcined PLS-1 (CDO) calcined PLS-3 (FER) calcined PLS-4 (CDO) calcined PREFER (FER)

Figure 6. 1H MAS NMR spectra for (a) APZ-1, (b) APZ-2, (c) APZ3, and (d) APZ-4.

D(Ar)

SBET(H2O)

(mL/g)

(nm)

(m2/g, S.T.P)

0.052 0.143 0.123 0.108 0.053

0.218 0.320 0.159 0.164 0.178

0.52 0.51, 0.57 0.52 0.55

196 127 125 162

523

0.147

0.331

318

0.154

0.172

366

0.133

0.154

(m2/g, S.T.P)

(mL/g)

235 514 346 332 284

parameter of the Q2 site, g(Si7), was approximately 0.19, signifying that adjacent terminal silanols, around 38% of which are facing each other, were bridged by the Q2 Si-tetrahedra. Consequently, a two-dimensional 2D micropore with a 10 × 10-MR pore opening was formed, although most of the interlayer space does not become micropores. Furthermore, structure refinement revealed that some water molecules are located around the Q2 Si-tetrahedra. Adsorbed water molecule sites WO1 and WO2 are located around the center of the micropore. The possible atomic distances between site WO and the silanol site (O13 and O14), l(WO1-O14), l(WO2-O13), and l(WO2-O14) were 3.13, 2.28, and 2.87 Å, respectively. This estimation is reasonable while considering the low occupancies of these sites. Site WO3 is somewhat close to site Si4 with an atomic distance of 1.73 Å. The number of adsorbed water molecules was approximately 7.2 per unit cell (viz., 10 wt %), which is slightly larger than the TG-DTA analysis, which gave about 8 wt %. Consequently, the presence of the additive Si site with a Q2-structure is confirmed by 29Si MAS NMR and XRD measurements. These findings suggest that Si fragments

Figure 7. Nitrogen gas adsorption isotherms for (a) APZ-1, (b) APZ2, (c) APZ-3, and (d) APZ-4. They are offset by 0, 100, 200, and 300 mL · g-1, respectively.

lattice constants b and c are compatible with those of PLS-1. A lattice parameter a along the layer stacking direction was increased from 1.057 to 1.097 nm (Table 4), although the interlayer distance slightly increased ca. 0.02 nm. An average stacking sequence of ferrierite layers is almost the same as PLS1, although some stacking fault might have occurred. Each ferrierite layer is stacked with a shift of approximately -0.16c along the [001] direction. Because elongated electron density distributions were observed between adjacent terminal silanols by the MEM/Rietveld analysis, the presence of a new Q2 pillar site (site Si7) was suggested as if a tSi-O-Si(OH)2-O-Sit linkage was formed in the interlayer. Two equivalent sites Si7 were positioned between adjacent O8 sites along the [010] direction (Figure 9), which is regarded as a split atom model by restriction of symmetric operation. The interatomic distance between the oxygen atom of the Q2 silanol group and nearest oxygen atom of the framework is approximately 2.2 Å. The occupancy

Figure 8. Water vapor adsorption isotherms for (a) APZ-1, (b) APZ2, (c) APZ-3, (d) APZ-4, and (e) ZSM-5 (Si/Al ≈ 100). They are offset by 0, 0.05, 0.10, 0.15, and 0.20 g · g-1, respectively.

Analysis of Si-Atom Pillared Lamellar Silicates

J. Phys. Chem. C, Vol. 114, No. 8, 2010 3473

TABLE 4: Comparison of Unit-Cell Parameters Among APZs, Layered Precursors, and Zeolites after Direct Calcination of Precursorsa APZ-1 PLS-1 CDO zeolite (cal-PLS-1) APZ-2 PLS-3 FER zeolite (cal-PLS-3) APZ-3 PLS-4 CDO zeolite (cal-PLS-4) APZ-4 PREFER FER zeolite (cal-PREFER)

a/nm

b/nm

c/nm

β/°

V/nm3

S.G.

N.L.

1.097* 1.057* 1.835*

1.402 1.401 1.378

0.741 0.742 0.737

98.3 98.0 -

1.127 1.088 1.863

P21/m P21/m Pnma

single single double

1.405 1.399 1.406

0.742 0.742 0.742

2.394* 2.332* 1.870*

-

2.494 2.419 1.949

Pnnm Pnnm Pnnm

double double double

2.334* 1.398 1.843*

1.396 0.738 1.388

0.739 2.220* 0.739

89.8 -

2.408 2.288 1.888

Pnma P21/m Pnma

double double double

1.403 2.635* 1.406

0.741 1.397 0.742

2.390* 0.743 1.874*

-

2.485 2.734 1.949

Pnnm Pnnm Pnnm

double double double

a S.G., N.L., and asterisk indicate space group, number of layer, and stacking direction of silicate layers, respectively.

Figure 10. Crystal structure model of APZ-2 and APZ-4 along the (upper) [010] direction and the (lower) [100] direction.

Figure 9. Crystal structure model of APZ-1 along the (upper) [001] direction and the (lower) [010] direction.

were yielded by dissolution of the silicate layer during the acid treatment. The Si fragment migrates into the interlayer and forms a new linkage between terminal Q3 silanol and the Si fragment

by dehydration-condensation. Therefore, we called the phenomenon of this structural change “self-reassembly”. Figure 10 shows the crystal structure model of APZ-2. Interlayer distances were expanded by about 0.06 nm through additional Q2 pillars, forming a 2D micropore structure with 12 × 10-MR and 8 × 10-MR pore openings. The structural viewing of APZ-2 along the [010] direction was almost compatible with the IEZ-FER proposed by Ruan.33 The stacking sequence of ferrierite layers is consistent with that of the FERtype zeolite without a layer shift parallel to the a-b plane. In the proposed model,33 the nearest hydroxyl groups of neighboring Q2 pillars are close to facing each other and are bonded by dehydration-condensation during calcination. However, the two hydroxyl groups of the Q2 pillars are located inside large 12MR pores in APZ-2. In this study, as-prepared samples were used for structure analysis. In APZ-2, a large quantity of adsorbed water molecules distributed in large pores is detected by the powder XRD measurement. The atomic distances

3474

J. Phys. Chem. C, Vol. 114, No. 8, 2010

between site WO and the silanol site (O11 and O12), l(WO1-O12), l(WO1-O12), and l(WO3-O11) were 1.87, 2.50, and 2.35 Å, respectively. It is considered that hydroxyl groups of the Q2 pillars are associated with adsorbed water molecules by strong hydrogen bondings, which are supported by our 1H MAS NMR measurements as shown in Figure 6. The occupancy parameter of the Q2 site, g(Si6), was approximated as ca. 0.60 by the refinement, which means that adjacent terminal silanols facing each other of ca. 60% were bridged by the Q2 Si-tetrahedra. The crystal structure of APZ-4 was almost compatible with that of APZ-2. The position of Q2 Si-tetrahedra was determined more precisely because the crystalline size of PREFER is adequately large and the crystallinity of APZ-4 did not deteriorate very much by the acid treatment. The occupancy parameter, g(Si6), of the Q2 pillar site was approximately 0.57 for APZ-4. The total amount of adsorbed water molecules in APZ-2 and APZ-4 was around 16 and 14 wt %, respectively, as calculated from refined occupancy parameters. This figure is larger than the results of TG-DTA measurement (ca. 6-8 wt %). Structural viewing of APZ-3 along the [001] direction (Figure 11) is similar to that of APZ-1, although their lateral viewing [010] direction resembles that of APZ-2. The lattice parameter along the stacking direction and the interlayer distance were clearly increased from 2.220 to 2.389 nm and from 0.17 to 0.23 nm, respectively. Consecutive ferrierite layers are stacked in a sequence ABABAB as well as PLS-4. The magnitude of the layer shift is 0.55c along the [001] direction, which is larger than that of PLS-4 (approximately 0.35c). In other words, the positions of two neighboring oxygen atoms (site O5) get closer to each other (Figure 11). The distribution of additional Q2 sites is similar to that of APZ-1, which connects with adjacent layer frameworks. The occupancy parameter of the Q2 site, g(Si7), was approximately 0.45 by the refinement; viz., adjacent terminal silanols facing each other of ca. 89% were bridged by the Q2 tetrahedra. Consequently, 2D micropores with 10 × 10MR pore opening are formed the same as APZ-1. The neighboring atomic distances between site WO and the silanol site (O12 and O13) l(WO1-O13) and l(WO2-O12) were 2.28 and 2.30 Å, respectively. The total amount of adsorbed water molecules was estimated to be around 11 wt %, a larger figure than that of the TG-DTA measurement (about 5 wt %). In the structure refinement of all APZs, the number of adsorbed water molecule was overestimated relative to the result of TG-DTA measurement. At present, this contradiction is not clarified. However, a part of the water molecule detected in micropores may be substituted for residual SiO4 fragments or hydrocarbon species. Furthermore, the number of Q2 Si atoms calculated from g(Si) values was larger than that an estimated value 29Si DDMAS NMR measurements (20%-23%) in all APZs. Because the observed relative intensity ratio of the Q2 resonance peak is almost the same (ca. 2.3%) among all APZs (Table 2), we guess that this discrepancy is due to insufficiency of recycle delay time (100 s) for a quantitative estimate of Q2resonance in 29Si DDMAS NMR measurements; namely, relaxation time T1 may be abnormally long. It is assumed that the motion of the Q2 Si-tetrahedra is strongly affected by the coupling with adsorbed water molecules. Additionally, the final MEM analyses for all APZs clearly show Q2-pillar atoms between layers and framework structures with covalent bonding in an electron density level as depicted

Ikeda et al.

Figure 11. Crystal structure model of APZ-3 along the (upper) [001] direction and the (lower) [010] direction.

in Figures S7-S10 (see Supporting Information). The presence of the Q2-pillar was surely proved by the electron density images. Refined structural parameters for each APZ are listed in Tables S1-S4 (see Supporting Information). A bond length, l, l(Si-O), and bond angles, φ, φ(O-Si-O), calculated from refined atomic coordinates and lattice parameters were summarized in Table S5 and Table S6 (Supporting Information). Observed, calculated, and difference patterns obtained by the Rietveld refinement for all APZs are shown in Figures S3-S6 (Supporting Information). R-factors were adequately low (Table 1). 3.5. Thermal Stability of APZ. The thermal stability of APZ was investigated by XRD after thermal treatment. Figure 12 shows powder XRD patterns of APZ-1 after calcination up to 873 K. No structural change of APZ-1 is observed up to 473 K, but the crystal structure was gradually collapsed and changed in the higher-temperature region. The d-spacing of the basal plane was shrunk from 1.089 to 1.009 nm with increasing calcination temperature. The unit cell volume was decreased by about 3.8%. Furthermore, profile broadening of the basal plane (the 100 reflection) took place considerably, indicating

Analysis of Si-Atom Pillared Lamellar Silicates

J. Phys. Chem. C, Vol. 114, No. 8, 2010 3475 Si fragments migrate with little efficiency and disperse into the interlayer space; (3) finally, adjacent silica layers are bridged by interlayer silylation using the Si fragments. It is assumed that the Si fragments are generated everywhere in the crystal, probably even on the outer surface of the crystal, by the acid treatment. The APZs obtained show high gas adsorption properties and narrow pore size distributions, although the micropore structure is formed incompletely owing to the lack of Q2 pillars. Furthermore, APZs have a high hydrophilicity due to the existence of a large number of hydroxyl groups in micropores. Nevertheless, APZs show high thermal stability up to 873 K without losing porosity. Supporting Information Available: Additional experimental details and CIF files. This material is available free of charge via the Internet at http://pubs.acs.org. References and Notes

Figure 12. Powder XRD patterns of APZ-1 and its calcined form at each heating temperature. A vertical dot line indicates the 100 peak position of APZ-1.

disorderly stacking arrangement of ferrierite layers. From a preliminary structure refinement, interlayer distance was slightly shrunk, and the ferrierite layers moved in the opposite direction from each other along the c-axis. However, the micropore structure was maintained after calcination, and the BET surface area after calcination at 873 K was 180 m2/g, as measured by a N2 gas adsorption measurement. On the other hand, structural changes were small in other APZs as shown in Figure S2 (see Supporting Information), although the basal d-spacing was slightly decreased. The crystallinity of APZ-2 and APZ-3 was slightly degraded during calcination. These facts indicate that all APZs have high thermal stability up to 873 K at least. 4. Conclusion Four kinds of Si-atom pillared lamellar silicates, APZ-1, APZ-2, APZ-3, and APZ-4, were prepared by the thermal acid (HCl) treatment of layered silicates PLS-1, PLS-3, PLS4, and PREFER, respectively, at 443 K for 24 h. The simple preparation method without using alkoxysilane is large merit in comparison with previous works.26,27,33 All APZs have 2D microporous structures with 10 × 10-MR and/or 12 × 10MR pore openings, whose structures are constructed of a SiO2(-OH)2 bridge (Q2-structure) between terminal oxygen atoms (Q3 sites) of neighboring silicate layers. The stacking sequence of the ferrierite layer for each APZ is similar to that of the layered silicates used. The presence of Q2 pillars was detected by the 29Si MAS NMR and powder X-ray diffraction techniques. The SDA molecules accommodated into the interlayer of each layered silicate were almost removed during the acid treatment, which is elucidated by the 13C MAS NMR and TG-DTA measurements. From all of the above experimental results, we argue that the new micropore structure is formed by the following process: (1) Si fragments originate from the structural collapse of silica layers by the thermal acid treatment; (2)

(1) Schreyeck, L.; Caullet, P. H.; Mougenel, J. C.; Guth, J. L.; Marler, B. Microporous Mater. 1996, 6, 259–271. (2) Leonowicz, M. E.; Lawton, J. A.; Lawton, S. L.; Rubin, M. K. Science 1994, 264, 1910–1913. (3) Bellussi, G.; Perego, G.; Clerici, M. G.; Giusti, A. European Patent Application 293,032, 1988. (4) Puppe, L.; Weisser, J. U.S. Patent 4,439,409, 1984. (5) Camblor, M. A.; Corma, A.; Dı´az-Caban˜as, M. -J; Baerlocher, Ch. J. Phys. Chem. B. 1998, 102, 44–51. (6) Zones, S. I.; Holtermann, D. I.; Innes, R. A.; Pecoraro, T. A.; Santilli, D. S.; Ziemer, J. N. U.S. Patent 4,826,667, 1989. (7) Ikeda, T.; Akiyama, Y.; Oumi, Y.; Kawai, A.; Mizukami, F. Angew. Chem., Int. Ed. 2004, 43, 4892–4896. (8) Zanardi, S.; Alberti, A.; Cruciani, G.; Corma, A.; Fornes, V.; Brunelli, M. Angew. Chem., Int. Ed. 2004, 43, 4933–4937. (9) Marler, B.; Stro¨ter, N.; Gies, H. Microporous Mesoporous Mater. 2005, 83, 201–211. (10) Marler, B.; Camblor, M. A.; Gies, H. Microporous Mesoporous Mater. 2006, 90, 87–101. (11) Oumi, Y.; Takeoka, T.; Ikeda, T.; Yokoyama, T.; Sano, T. New J. Chem. 2007, 31, 593–597. (12) Ikeda, T.; Oumi, Y.; Takeoka, T.; Yokoyama, T.; Sano, T.; Hanaoka, T. Microporous Mesoporous Mater. 2008, 110, 488–500. (13) Wang, Y.; Marler, B.; Gies, H.; Mu¨ller, U. Chem. Mater. 2005, 17, 43–49. (14) Ikeda, T.; Syunsuke, K.; Mizukami, F. J. Mater. Chem. 2009, 19, 5518–5525. (15) Song, J.; Gies, H. Recent Advances in the Science and Technology of Zeolites and Related Materials. Stud. Surf. Sci. Catal. 2004, 154A, 295– 300. (16) Marler, B.; Wang, Y.; Song, J.; Gies, H. Abstr. 14th Int. Zeolite Conf. 2007, 599. (17) Burton, A.; Accardi, R. J.; Lobo, R. F.; Falcioni, M.; Deem, M. W. Chem. Mater. 2000, 12, 2936–2942. (18) Dorset, D. L.; Kennedy, G. J. J. Phys. Chem. B 2004, 108, 15216– 15222. (19) Knight, L. M.; Miller, M. A.; Koster, S. C.; Gatter, M. G.; Benin, A. I.; Willis, R. R.; Lewis, G. J.; Broach, R. W. Stud. Surf. Sci. Catal. 2007, 170A, 338–346. (20) Blake, A. J.; Franklin, K. R.; Lowe, B. M. J. Chem. Soc., Dalton Trans. 1988, 2513–2517. (21) Andrews, S. J.; Papiz, M. Z.; McMeeking, R.; Blake, A. J.; Lowe, B. M.; Franklin, K. R.; Helliwell, J. R.; Harding, M. M. Acta Crystallogr. B 1988, 44, 73–77. (22) Vortmann, S.; Rius, J.; Siegmann, S.; Gies, H. J. Phys. Chem. B 1997, 101, 1292. (23) Wang, Y. X.; Gies, H.; Lin, J. H. Chem. Mater. 2007, 19, 4181– 4188. (24) Fan, W.; Wu, P.; Namba, S.; Tatsumi, T. Angew. Chem., Int. Ed. 2004, 43, 236–240. (25) Ruan, J.; Wu, P.; Slater, B.; Terasaki, O. Angew. Chem., Int. Ed. 2005, 44, 6719–6723. (26) Inagaki, S.; Yokoi, T.; Kubota, Y.; Tatsumi, T. Chem. Commun. 2007, 5188–5190. (27) Wu, P.; Ruan, J.; Wang, L.; Wu, L.; Wang, Y.; Liu, Y.; Fan, W.; He, M.; Terasaki, T.; Tatsumi, T. J. Am. Chem. Soc. 2008, 130, 8178– 8187. (28) Mochizuki, D.; Shimojima, A.; Imagawa, T.; Kuroda, K. J. Am. Chem. Soc. 2005, 127, 7183–7191. (29) Mochizuki, D.; Kuroda, K. New J. Chem. 2006, 30, 277–285.

3476

J. Phys. Chem. C, Vol. 114, No. 8, 2010

(30) Mochizuki, D.; Kowata, S.; Kuroda, K. Chem. Mater. 2006, 18, 5223–5229. (31) Ishii, R.; Ikeda, T.; Itoh, T.; Ebina, T.; Yokoyama, T.; Hanaoka, T.; Mizukami, F. J. Mater. Chem. 2006, 16, 4035–4043. (32) Ishii, R.; Ikeda, T.; Mizukami, F. J. Colloid Interface Sci. 2009, 331, 417–424. (33) Ruan, J.; Wu, P.; Slater, B.; Zhao, Z.; Wu, L.; Terasaki, O. Chem. Mater. 2009, 21, 2904–2911. (34) Ikeda, T.; Kayamori, S.; Mizukami, F.; Oumi, Y. Japanese Patent JP2008137857A, 2008. (35) Izumi, F.; Momma, K. Solid State Phenom. 2007, 130, 15–20. (36) Sakata, M.; Sato, M. Acta Crystallogr., Sect. A 1990, 46, 263– 270. (37) Izumi, F.; Dilanian, R. A. Recent Research DeVelopments in Physics; Transworld Research Network: Trivandrum, 2002; Vol. 3, Part II, pp 699-726. (38) Momma, K.; Izumi, F. J. Appl. Crystallogr. 2008, 41, 653–658.

Ikeda et al. (39) Bennett, P. C.; Melcer, M. E.; Siegel, D. I.; Hasset, J. P. Geochem. Cosmochem. Acta 1988, 52, 1521–1530. (40) Terry, B. Hydrometallurgy 1983, 10, 135–150. (41) Eckert, H.; Yesinowski, J. P.; Silver, L. A.; Stolper, E. M. J. Phys. Chem. 1988, 92, 2055–2064. (42) Xue, X.; Kanzaki, M. J. Phys. Chem. B 2007, 111, 13156–13166. (43) Paris, M.; Fritsch, E.; Aguilar Reyes, B. O. J. Non-Cryst. Solids 2007, 353, 1650. (44) d’Espinose de la Caillerie, J. B.; Aimeur, M. R.; El Kortobi, Y.; Legrand, A. P. J. Colloid Interface Sci. 1997, 194, 434–439. (45) Brunauer, S.; Emmett, P. H.; Teller, E. J. Am. Chem. Soc. 1938, 60, 309–319. (46) Neimark, A. V.; Ravikovitch, P. I.; Grune, M.; Schueth, F.; Unger, K. K. J. Colloid Interface Sci. 1998, 207, 159–169. (47) Brunauer, S.; Deming, L.; Deming, W.; Teller, E. J. Am. Chem. Soc. 1940, 62, 1723–1732.

JP912026N