You are on page 1of 9

Microporous and Mesoporous Materials 142 (2011) 292300

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


journal homepage: www.elsevier.com/locate/micromeso

i-Silica: Nanostructured silica hybrid materials containing imidazolium groups by hydrolysis-polycondensation of disilylated bis-N,N0 -alkyl-imidazolium halides
Thy Phuong Nguyen, Peter Hesemann , Jol J.E. Moreau
Institut Charles Gerhardt de Montpellier, Ecole Nationale Suprieure de Chimie de Montpellier, 8 rue de lEcole Normale, 34296 Montpellier cedex 05, France

a r t i c l e

i n f o

a b s t r a c t
We report the synthesis of silica hybrid materials from the ionic bis-N,N0 -(3-triethoxysilyl)propyl imidazolium chloride precursor. This precursor shows very particular behaviour in template directed hydrolysis polycondensation procedures due to its ionic nature. Nanostructured phases were obtained by solgel transformation of this precursor in the presence of anionic surfactants. In contrast, hydrolysis-polycondensation under standard reaction conditions using usually applied templating methods with cationic or neutral surfactants led to the formation of amorphous materials. Detailed insight in the hydrolysis polycondensation behaviour of this precursor was obtained via co-condensation reactions of the ionic precursor with a silica network former tetraethoxysilane (TEOS). It appears that the formation of nanostructured phases in the presence of anionic surfactants can be explained by ionic interactions between the template and the organo-cationic substructure of the imidazolium precursor in the hydrolysis polycondensation mixture. These results point a new templating mechanism for the formation of nanostructured silica hybrid materials. 2010 Elsevier Inc. All rights reserved.

Article history: Received 29 October 2010 Received in revised form 9 December 2010 Accepted 10 December 2010 Available online 17 December 2010 Keywords: Periodic mesoporous organosilica Nanostructured silica hybrid materials Template synthesis Ionic liquids Anionic surfactant

1. Introduction In the eld of silica hybrid materials, periodic mesoporous organosilicas (PMOs) recently emerged as a new class of versatile functional materials with dened architectures on a mesoscopic length scale [13]. These materials can be obtained via template directed syntheses [4]. PMO type materials, typically synthesized by hydrolysis-polycondensation of bridged bis-trialcoxysilylated precursor molecules [(R0 O)3Si-R-Si(OR0 )3] in the presence of structure directing agents [5], display two main features which distinguish them fundamentally from other silica-based materials, which are (i) dened pore architecture on a mesoscopic level, generating high porosity and narrow pore size distribution, and (ii) organic functional groups incorporated in the pore walls, conferring specic surface properties to the materials. PMOs containing various types of organic substructures have been reported, including small aliphatic [6], vinylic [7] or aromatic moieties [8] but also more complex structures such as isocyanurates [9], dendrimers [10], fullerenes [11] or chiral groups [1214]. PMO type materials attracted great interest in the last few years owing to their large potential in catalysis [15], sorption [16] and optics [17,18]. The chemical immobilization of ionic species on solid silica support was introduced by Holderich et al. [1921] and Mehnert et al. [22,23] in 2001/2002. This approach gives rise to materials with
Corresponding author. Tel.: +33 (0)4 67 14 72 17; fax: +33 (0)4 67 14 43 53.
E-mail address: peter.hesemann@enscm.fr (P. Hesemann). 1387-1811/$ - see front matter 2010 Elsevier Inc. All rights reserved. doi:10.1016/j.micromeso.2010.12.014

large potential in heterogeneous catalysis, chromatography and separation. Most of the materials are surface functionalized silicates, synthesized via template directed co-condensation reactions of silylated precursors with silica network formers or post-grafting reactions. In contrast, reports on silica hybrid materials, so-called polysilsesquioxanes, bearing ionic entities obtained by hydrolysis polycondensation of bridged ionic precursors of the type [(R0 O)3Si-R+-Si(OR0 )3 X] are relatively scarce. Garci described MCM-41 type materials bearing viologen and 1,2-bis-(4-pyridyl)ethylene groups embedded in a silica matrix [24,25]. Large excess of silica precursors (tetraethoxysilane, TEOS) were necessary in order to obtain materials with dened architectures on a mesoscopic scale. Inagaki and co-workers recently reported PMO-type materials bearing cationic methylacridinium substructures [26]. Most studies in this eld are related to disilylated imidazolium or dihydroimidazolium precursors 1 and 2 (Scheme 1). Firstly reported by Veith, Braunstein et al. [27], these compounds were already used for the elaboration of anion exchange materials [28] or as ionic linker to connect silica nanoparticles [29]. Very recently, Karimi et al. reported porous silica-based materials obtained from the imidazolium precursor 1 and TEOS and their use as heterogeneous precursors for the formation for PMO supported Pd-NHC species or as ionic support for the immobilization of rare earth catalysts [30,31]. Our special concern is the synthesis of silica-based materials containing ionic species [3237]. These i-silica type materials have large potential for applications in catalysis [38], and

T.P. Nguyen et al. / Microporous and Mesoporous Materials 142 (2011) 292300

293

(EtO)3Si

N Cl
-

Si(OEt)3

(EtO)3Si

N Cl-

Si(OEt)3

1
1,3-bis(3-(triethoxysilyl)propyl)-1H-imidazol-3-ium chloride

2
1,3-bis(3-(triethoxysilyl)propyl)-4,5-dihydro-1H-imidazol-3-ium chloride

Scheme 1. Disilylated imidazolium and dihydroimidazolium precursors 1 and 2.

separation [3941]. We are particularly interested if the ionic character of ammonium or imidazolium precursors has direct inuence on the generation of structured phases and if materials featuring particular architectures can be obtained from these compounds. The inuence of ionic species in the hydrolysis polycondensation mixture on the architecture of nanostructured silica materials was already monitored in the synthesis of highly ordered silica [42] or large-pore hexagonal PMO materials with long-range structural order [43] which were obtained with the aid of inorganic salts. We especially focus on the elaboration of silica supported ionic phases with dened architectures. For example, we studied nanostructured hybrid silica bearing ammonium [32] or di-arylimidazolium entities [34], which were obtained by template directed hydrolysis-polycondensation from related trialkoxysilylated precursors. These original materials show interesting composition, textural and morphological features such as homogeneously distributed ionic species over the whole silica material, high porosity and narrow pore size distribution. We already studied the disilylated imidazolium precursor 1 in view of the generation of hybrid supported NHC catalysts [44]. The precursor was successfully reacted with palladium acetate in order to yield the corresponding tetrasilylated Pd-NHC complex. This molecular carbene complex was converted to the Pd-NHC containing silica hybrid via uoride catalyzed hydrolysis-polycondensation reaction, and the resulting material was successfully used as heterogeneous catalyst for HeckMirozoki cross-coupling reactions. The formation of nanostructured silica materials from the di-alkyl imidazolium precursor 1 or the related dihydroimidazolium derivative 2 is still subject of debate. Dai et al. reported mesoporous materials from the dihydroimidazolium precursor 2 by templating with cationic surfactants in alkaline reaction media [28]. Higher porosity was found in materials synthesized in the presence of imidazolium derived surfactants compared to solids prepared in the presence of ammonium salts (CTAB). The author suggest that this result can be attributed to favorable interactions between the dihydroimidazolium precursor and the used long-chain substituted imidazolium surfactant. Other works indicate that porous phases can only be obtained by co-condensation approaches using the imidazolium precursor 1 or the dihydroimidazolium precursor 2 and considerable amounts of silica network formers such as TEOS [30,45]. Yin et al. systematically studied the hydrolysis-polycondensation behaviour of the dihydroimidazolium precursor 2 in template directed hydrolysis polycondensation reactions using the neutral triblock copolymer P123 as structure directing agent. Structured phases were only obtained from hydrolysis polycondensation mixtures containing a molar TEOS/precursor 2 ratio of at least 18/1 [45]. Karimi et al. reported that the addition of mineral salt to the hydrolysis polycondensation mixtures allowed to reduce signicantly this ratio. Porous phases were obtained from a molar TEOS/precursor 1 ratio of 9/1 [30]. Although these studies indicate the chemical stability of the ionic precursors 1 and 2 under the different reaction conditions, it is of prime importance to give scientic insights in the knowledge on the ordering and on the behaviour of the resulting materials. Anionic surfactants have already been used for the synthesis of nanostructured silica by the aid of silylated ammonium species

[4648]. More recently, we described the synthesis of PMO type materials containing ammonium substructures by anionic templating [32]. Here, we focused on a better understanding of the mechanisms which govern the formation of nanostructured PMO type materials bearing ionic imidazolium substructures both using classical templating conditions using cationic or non-ionic templates and anionic templating strategies. 2. Experimental 2.1. Materials The used anionic sulphate surfactants were purchased at ABCR. Four different anionic surfactants were used: (1) sodium dodecylsulphate (SDS), (2) sodium hexadecylsulphate (SHS), (3) sodium cetyl stearyl sulphate (mixture of 60% sodium hexadecylsulphate/40% sodium octadecylsulphate) and (4) sodium octadecylsulphate. TEOS, Brij76, P123 and CTAB were purchased at SigmaAldrich. All chemicals were used as received except TEOS which was distilled before utilization. The synthesis of precursor 1 was performed according to published procedures [27,44]. 2.2. General methods Solid state 13C and 29Si CP MAS NMR experiments were recorded on a Varian VNMRS 400 MHz solid spectrometer using a two channel probe with 7.5 mm ZrO2 rotors. The 29Si Solid NMR spectra were recorded using both CP MAS and One Pulse (OP-) sequences with samples spinning at 6 kHz. CP MAS was used to get high signal to noise ratio with 5 ms contact time and 5 s recycling delay. For OP experiments, p/6 pulse and 60 s recycling delay were used to obtain quantitative information on the silanesilanol condensation degree. The number of scans was in the range 1000 3000. Nitrogen sorption isotherms at 77 K were obtained with a Micromeritics ASAP 2020 apparatus. Prior to measurement, the samples were degassed for 18 h at 100 C. The surface areas (SBET) were determined from BET treatment in the range 0.040.3 p/p0 and assuming a surface coverage of nitrogen molecule estimated to 13.5 2. Pore size distributions were calculated from the adsorption branch of the isotherms using the BJH method. The pore width was estimated at the maximum of the pore size distribution. TEM images were obtained using JEOL 1200 EX II (120 kV). XRD experiments were carried out with a Xpert-Pro (PanAnalytical) diffractometer equipped with a fast Xcelerator detector using Cu Ka radiation. 2.3. Materials syntheses 2.3.1. Syntheses of pure hybrid materials from the precursor 1 in the presence of anionic surfactants These materials were synthesized according to a general procedure as follows: Surfactant (1.95 mmol) was solubilized in a solution containing 46 mL of water and 0.53 mL of hydrochloric acid (1 M). This mixture was heated to 60 C during 30 min before a solution of 1.00 g (1.95 mmol) of precursor 1 in 4 mL of ethanol 95% was

294

T.P. Nguyen et al. / Microporous and Mesoporous Materials 142 (2011) 292300

rapidly added. The hydrolysis-polycondensation mixture was vigorously stirred at 60 C. A white precipitate formed after approximately 30 s. After 1 h, the precipitate was ltered and washed three times with distilled water. Finally, the nanocomposite was dried at 70 C for 24 h. 2.3.2. Syntheses of silica materials by co-condensation reactions of precursor 1 and TEOS in the presence of anionic surfactants These materials were synthesized according to a general procedure as representatively described for the preparation of material 1-A16A18-1/4: 1.95 mmol of the surfactant was solubilized in a solution containing 41.7 mL of water and 5.3 mL of hydrochloric acid (1 M). This mixture was heated to 60 C during 30 min before a solution of 1.00 g (1.95 mmol) of precursor 1 and 1.62 g (7.80 mmol) of TEOS, dissolved in 4 mL of ethanol 95%, was rapidly added. The hydrolysis-polycondensation mixture was vigorously stirred at 60 C. The molar composition of the reaction mixture was as follows: precursor 1/TEOS/A1618/HCl/H2O = 0.75/3.0/0.75/ 2/1000. After sufcient reaction time, the precipitate was ltered and washed three times with distilled water. Finally, the obtained nanocomposite 1-A16A18-1/4 was dried at 70 C for 24 h. 2.3.3. Syntheses of silica materials by co-condensation reactions of precursor 1 and TEOS in the presence of cationic surfactants (CTAB) These materials are synthesized according to a general procedure as representatively described for the preparation of material 1-CTAB-1/9. 1.70 g of CTAB were dissolved in a mixture containing 79.8 mL of distilled water and 5.3 g of an aqueous ammonia solution (25%). Then, 2.00 g of precursor 1 and 7.31 g of TEOS were simultaneously added with stirring. The molar composition of the reaction mixture was as follows: precursor 1/TEOS/CTAB/ NH3/H2O = 0.1/0.9/0.12/8/114. The hydrolysis-polycondensation mixture was vigorously stirred at 60 C for 3 h. Then, the precipitate was kept at 90 C without stirring for 48 h. After this time, the precipitate was ltered and washed three times with distilled water. Finally, the obtained nanocomposite 1-CTAB-1/9 was dried at 70 C for 24 h. 2.3.4. Syntheses of silica materials by co-condensation reactions of precursor 1 and TEOS in the presence of non-ionic surfactants (P123) These materials are synthesized according to a general procedure as representatively described for the preparation of material 1-P123-1/9. P123 (1.64 g) were dissolved in 69 mL of aqueous solution of hydrochloric acid (2 M) at 45 C. After 3 h, 3.00 g of TEOS were added and the resulting solution was stirred for further 60 min at the same temperature. Then, 616 mg of precursor 1 were added. The molar composition of the reaction mixture was as follows: precursor 1/TEOS/P123/HCl/H2O = 0.5/4.5/0.08/35/1100. The reaction mixture was stirred for further 15 h and then kept without stirring at 90 C during 48 h. After this time, the precipitate was ltered and washed three times with distilled water. Finally, the obtained nanocomposite 1-P123-1/9 was dried at 70 C for 24 h. 2.3.5. Removal of the surfactants by washing 2.3.5.1. Anionic surfactants. Nanocomposites (500 mg) were suspended in a solution containing 100 mL of ethanol 95%, 3 mL of

an aqueous ammonia solution (25%) and 500 mg of NH4Cl during 24 h. Finally, the solids were isolated by ltration and dried at 70 C during 24 h. 2.3.5.2. Cationic and non-ionic surfactants. Nanocomposites (500 mg) were suspended in a solution containing 100 mL of ethanol 95% and 2 mL of hydrochloric acid (37%) during 24 h. Finally, the solids were isolated by ltration and dried at 70 C during 24 h. 3. Results and discussion 3.1. Pure hybrid materials from precursor 1 We rstly studied the formation of i-silica from the di-alkyl imidazolium precursor 1. We realized hydrolysis polycondensation reactions of the pure precursor 1 (Scheme 2, y = 0) under different reaction conditions, in particular in the presence of various types of surfactant (cationic, non-ionic, anionic). The chemical stability of the di-alkyl-imidazolium substructure under both acidic and alkaline hydrolysis-polycondensation conditions has already been demonstrated in former studies [28,45]. Here, we monitored the incorporation of the di-alkyl-imidazolium substructure via 29Si solid state NMR spectroscopy. The 29Si OP MAS spectrum of material 1-CTAB is shown in Fig. 1, the spectra of materials 1-A16A18 and 1-P123 are given in Supplementary information. All spectra showed exclusively the signals of the T-series characteristic for alkyl substituted silicon centers R-SiO3 in the materials. The absence of signals of the Q-series (100 ppm < d < 120 ppm) conrmed that no SiC bond cleavage occurred during hydrolysis-polycondensation process. In all spectra, the most intense signals correspond to T3 environments indicating that completely condensed R-Si(OSi)3 species are predominant in all three materials. These results are in agreement with former results and indicate that the precursor 1 can be converted to the corresponding silica hybrid materials without chemical decomposition. This result also reects the higher chemical stability of the di-alkyl substituted precursors 1 and 2 compared to the related di-aryl-imidazolium precursor which decomposed in alkaline media [34].

Fig. 1. Solid state

29

Si OP MAS NMR spectra of the material 1-CTAB.

(EtO)3Si

N
+

Cl-

Si(OEt)3

hydrolysispolycondensation

O O O

Si

N Cl-

Si

O O

surfactant

Scheme 2. Template directed hydrolysis polycondensation of the disilylated imidazolium precursor 1.

T.P. Nguyen et al. / Microporous and Mesoporous Materials 142 (2011) 292300

295

Fig. 2. Small angle X-ray diffractograms of the nanocomposites 1-A16A18, 1-Brij76, 1-A16A18-P123 and 1-CTAB.

Fig. 3. Small angle X-ray diffractograms of the nanocomposites 1-A12, 1-A16, 1A16A18 and 1-A18.

The X-ray diffractograms of the as-synthesized nanocomposites materials, before surfactant elimination, are shown in Fig. 2. Much to our delight, nanostructured composites were formed in the presence of anionic surfactant. The diffractogram of the material obtained in the presence of the anionic surfactant sodium cetyl stearyl sulphate, containing 60% sodium hexadecylsulphate and 40% sodium octadecylsulphate), shows the three diffraction rays (1 0 0), (1 1 0) and (2 0 0) characteristic for materials with 2d-hexagonal architectures (Fig. 2, material 1-A16A18). In contrast, hydrolysis-polycondensation reactions of the precursor in the presence of commonly used neutral or cationic surfactants yielded completely amorphous materials (Fig. 2 and 1-Brij76, 1-P123 and 1-CTAB). These results prompted us to investigate the utilization of anionic surfactants as templates for hydrolysis-polycondensation processes with cationic imidazolium substructures more in detail. In the following, we performed hydrolysis polycondensation reactions with precursor 1 in the presence of different anionic surfactants, in particular by a variation of the chain length of the alkyl group of the long chain substituted sulphates. Hydrolysis-polycondensation reactions were performed in acidic reaction media from the imidazolium precursor 1 in the presence of the anionic surfactants sodium dodecylsulphate, sodium hexadecylsulphate, sodium cetyl stearyl sulphate (a mixture of 60% sodium hexadecylsulphate/40% sodium octadecylsulphate) and sodium octadecylsulphate, giving rise to the formation of the materials 1-A12, 1-A16, 1-A16A18 and 1-A18, respectively. The diffractograms of these materials are shown in Fig. 3 and indicate that all the nanocomposites are structured. The diffractograms of the materials indicate two trends. Firstly, the d100 reections shift to smaller reection angles, indicating an

increase of the d-spacing in this series of materials. Whereas the d100 reection in the diffractogram of material 1-A12 is located at 2h = 2.64 corresponding to a d-spacing of 3.3 nm, this reection in material 1-A18 can be found at 2h = 2.13 (d-spacing: 4.1 nm). This trend can be explained by the lager size of the micelles formed from sodium hexadecylsulphate or sodium octadecylsulphate containing C16 or C18 alkyl chains compared to sodium dodecylsulphate with relatively short C12 chains. Similar trends have been observed in template directed syntheses of mesoporous silica in the presence of various long-chain substituted cationic surfactants [49]. The second trend concerns the architecture of the materials. The diffractogram of material 1-A12 displays the (1 0 0) and (2 0 0) reections, characteristic for materials with lamellar structure. In the diffractograms of the materials 1-A16 or 1-A16A18, the (1 1 0) reection is also visible. The diffractogram of material 1A18 shows nearly identical intensity of the (1 1 0) and (2 0 0) reections, indicating that this material displays hexagonal architecture. These results suggest that lamellar materials are formed in the presence of short chain surfactants such as sodium dodecylsulphate, whereas 2d-hexagonal nanocomposites are formed using C18-substituted sodium octadecylsulphate as structure directing agent. All the presented diffractograms were obtained from as-synthesized materials and contain the anionic surfactant. In the following we addressed the question if the materials retain the regular architectures after surfactant elimination. Fig. 4 shows the X-ray diffractograms of material 1-A16A18-W and 1-A18-W, obtained from the nanostructured materials 1-A16A18 and 1-A18 by washing in acidic ethanol solution. The diffractograms do not show any diffraction ray and indicate that the materials are amorphous. Similar results were obtained for the whole series of structured composites

296

T.P. Nguyen et al. / Microporous and Mesoporous Materials 142 (2011) 292300

Fig. 4. Small angle X-ray diffractograms of the materials 1-A18-W and 1-A16A18W.

synthesized from the pure precursor 1 in the presence of various anionic surfactants. These results show that structured and stable PMO-type materials cannot be obtained from the pure precursor 1. Although structured nanocomposites were formed from this precursor in the presence of anionic surfactants, the pure hybrid materials are non-structured due to the presence of propyl chains in the precursor, conferring a high degree of exibility to the organic substructures within the hybrid material. For this reason, the textural stability of the materials is low and does not allow the formation of porous PMO type materials as reported for more rigid disilylated precursors containing methylene, ethylene or pphenylene fragments (vide supra). We showed that nanostructured surfactant/hybrid silica nanocomposites can be obtained from the dialkylimidazolium precursor 1 and various anionic surfactants. However, the morphological stability of the formed materials is not sufcient for the synthesis of nanostructured pure silica hybrid materials. The regular architectures of the structured nanocomposites collapse during the elimination of the surfactant by washing. For this reason, we studied in the following co-condensation reactions of the ionic precursor 1 and the silica network former tetraethyl-ortho-silicate (TEOS). Co-condensation reactions with TEOS allow to confer increased rigidity to the silica network and therefore yield mechanically more robust materials. Additionally, we were interested to investigate the repercussions of the presence of silica network formers such as TEOS in the co-condensation mixture on the architecture of the formed materials. 3.2. Co-condensation reactions with precursor 1 and TEOS Co-condensation reactions of organic precursors with silica network formers such as TEOS are commonly applied to obtain struc-

tured and porous phases [50,51]. We already used this strategy for the synthesis of nanostructured PMO type materials bearing the SiIMes structure [34]. Here we were interested to know if this strategy is also applicable on the precursor 1 and allows to synthesize nanostructured hybrid materials containing the di-alkyl-imidazolium substructure. For this reason, we carried out co-condensation reactions of the imidazolium precursor 1 with various amounts of TEOS (see Scheme 2, x = 1, y = 4, 6, 9, 12, 18). A series of materials were obtained by co-condensation reactions in the presence of anionic, cationic and neutral surfactants. We rstly performed hydrolysis-polycondensation reactions of the precursor 1 and TEOS in the presence of the anionic surfactant sodium cetyl stearyl sulphate (containing 60% sodium hexadecylsulphate/40% sodium octadecylsulphate) in molar ratios of 1/4, 1/ 6, 1/9 and 1/18. These reactions gave rise to the formation of the materials 1-A16A18-1/4, 1-A16A18-1/6, 1-A16A18-1/9 and 1A16A18-1/18. We conrmed the incorporation of the organo-ionic imidazolium substructures via solid state NMR measurements. 29Si OP MAS NMR allows to quantify the molar ratio between T and Q signals in the materials and therefore can serve to determine the proportion of the immobilized organo-ionic groups in the materials. The spectra of selected samples are given in Fig. 5. It appears that the T/Q ratios found in the hybrid gels are in very good agreement with the expected values calculated from the precursor 1/ TEOS ratio in the hydrolysis-polycondensation mixtures (Table 1). These results suggest that the ionic precursor is quantitatively incorporated in the silica network during the hydrolysis co-condensation reactions. The X-ray diffractograms of the obtained materials are shown in Fig. 6. It appears that the material 1-A16A18-1/4 is a nanostructured material. The diffractogram shows the three characteristic (1 0 0), (1 1 0) and (2 0 0) diffraction rays indicating hexagonal architecture on a mesoscopic length scale. The position of the (1 0 0) ray at 2h = 2.20 corresponding to a d-spacing of 4.0 nm is typical for the materials obtained with sodium cetyl stearyl sulphate and reects the templating effect of the anionic structure directing agent [32]. The materials obtained by co-condensation reactions with increasing amounts of TEOS show progressively

Fig. 5. Solid state 29Si OP MAS NMR spectra of the materials 1-A16A18-1/6, 1A16A18-1/9 and 1-A16A18-1/18.

Table 1 Calculated and found T/Q ratios in the 29Si OP MAS NMR spectra of the materials 1A16A18-1/6, 1-A16A18-1/9 and 1-A16A18-1/18. Material T/Q ratio in the materials Calcd. 1-A16A18-1/6 1-A16A18-1/9 1-A16A18-1/18 33/67 22/78 11/89 Found 33/67 21/79 10/90

T.P. Nguyen et al. / Microporous and Mesoporous Materials 142 (2011) 292300

297

Fig. 6. Small angle X-ray diffractograms of the nanocomposites 1-A16A18-1/4, 1A16A18-1/6, 1-A16A18-1/9 and 1-A16A18-1/18.

Fig. 7. Small angle X-ray diffractograms of the materials 1-CTAB-1/9, 1-P123-1/9, 1-P123-1/12 and 1-P123-1/18.

decreasing order. Materials 1-A16A18-1/6 and 1-A16A18-1/9 show intermediate regularity with reduced long range order or wormlike architectures. Finally, material 1-A16A18-1/18 is a completely amorphous solid. This result is in nice agreement with the fact that the hydrolysis-polycondensation of pure TEOS in the presence of anionic surfactants gives rise to the formation of amorphous silica [52]. It appears that structured phases are formed only from hydrolysis-polycondensation mixtures containing high concentrations of the ionic imidazolium precursor. Thus, the formation of structured solids can be related to the inuence of the ionic imidazolium precursor, whereas the presence of TEOS has an unfavorable effect for the formation of nanostructured phases. The diffractograms of the materials obtained by hydrolysispolycondensation reactions in the presence of cationic or neutral surfactants reveal a very different situation. Here, amorphous phases were obtained from reaction mixtures containing 1/TEOS ratios lower than 1/10. Materials displaying worm like architectures were obtained from hydrolysis polycondensation reactions mixtures containing molar precursor 1/TEOS ratios of 1/12 (Fig. 7 and 1-P123-1/12) as already reported by Yin and co-workers [45] and Karimi et al. [30]. Finally, in order to obtain highly structured phases with dened 2d-hexagonal architectures, the 1/TEOS ratio has to be increased to at least 1/18. In this case, a typical SBA-15 type silica with a d-spacing of 11.2 nm of was obtained (Fig. 7 and 1-P123-1/18). These results reect a completely different mechanism for the formation of structured silica hybrid phases. In the presence of cationic or neutral structure directing agents, the formation of nanostructured silica phases is governed by interactions involving the surfactant and siloxy groups, either (i) ionic interactions between cationic surfactants and silanolate groups in

the case of materials of the M41S family [50], or (ii) hydrogen bonding involving SiOH groups in the case of templating reactions in the presence of non-ionic surfactants and yielding for example materials of the SBA-15 type [53,54]. In both cases, the presence of organic precursors in general and the organo-ionic imidazolium precursor in particular in the hydrolysis polycondensation mixtures disturbs the supramolecular arrangement of silica precursors and templates, which is necessary for the formation of nanostructured phases. The addition of the imidazolium precursor 1 therefore has an unfavorable effect towards the formation of nanostructured silica phases. For this reason, a high excess of the silica precursor (TEOS) in the reaction mixture is necessary in order to generate nanostructured phases. After having investigated the formation of structured nanocomposites by co-condensation reactions involving the ionic precursor 1 and TEOS, we were now interested to address the morphological stability of the obtained materials after surfactant elimination by washing. Fig. 8 shows the X-ray diffractograms of materials 1A16A18-1/4-W and 1-A16A18-1/6-W, obtained from the parent materials after washing with acidic ethanol. It appears that the order in the hybrid materials is generally lower than in the nanocomposites. Although the materials contain domains with highly regular aligned channels as shown by transmission electron microscopy (Photo 1, left), the characterization of the materials by X-ray diffraction strongly suggest essentially worm-like architectures (Fig. 8). In contrast, the TEM micrograph of material 1-P123-1/12-W, synthesized in the presence of the non-ionic surfactant P123 in acidic reaction media, shows irregular arrangement of the pore structure and, in agreement to its X-ray diffractogram, principally worm-like architecture (Photo 1, right).

298

T.P. Nguyen et al. / Microporous and Mesoporous Materials 142 (2011) 292300

Fig. 9. Nitrogen sorption isotherms of materials 1-A16A18-1/4-W and 1-A16A18-1/ 6-W.

Table 2 Textural properties of selected materials containing dialkylimidazolium segments after washing. Material Specic surface area SBET (m2 g1) 384 607 973 655 Pore volume (cm3 g1) d100-spacing (nm) Average pore diametera (nm)

1-A1618-1/4-W 1-A1618-1/6-W 1-A1618-1/9-W 1-P123-1/12-W

0.24 0.38 0.67 1.04

5.2 4.5 4.4 11.4

3.0 2.3 2.5 5.6

Fig. 8. Small angle X-ray diffractogram of the materials 1-A16A18-1/4-W and 1A16A18-1/6-W.

a Pore diameters were calculated by the BJH method from the adsorption branch of the isotherms.

The characterization of the materials by nitrogen sorption experiments reveal sufcient robustness of the silica hybrid framework in order to maintain the porous phases during the washing process. The adsorptiondesorption isotherms of the materials 1A16A18-1/4-W and 1-A16A18-1/6-W are given as examples in Fig. 9. The textural properties of the materials are summarized in Table 2. The results of the nitrogen sorption experiments indicate several trends in the materials obtained via anionic templating. Increasing proportions of TEOS in the hydrolysis-polycondensation mixture leads to increased specic surface areas and pore volumes in the resulting materials. Low inuence can be observed concern-

ing the lattice parameters (d100 spacings) and pore diameters in the resulting materials. All materials obtained by templating with the anionic surfactant sodium cetyl stearyl sulphate show d100 spacings in the range of 4.45.2 nm and average pore diameters in the supermicroscopic area (approximately 2 nm). However, the limited applicability of BET equation on the supermicroscopic length scale has to be taken into account concerning the determination of the specic surface area. In contrast to these results, the material 1-P123-1/12-W, obtained by templating in the presence of the non-ionic surfactant P123, shows a typical SBA-15 type architecture with larger pore diameter and d100 spacing and, as a consequence, increased pore volume compared to the materials obtained by anionic templating.

Photo 1. TEM micrographs of the materials 1-A1618-1/6-W and 1-P123-1/12-W.

T.P. Nguyen et al. / Microporous and Mesoporous Materials 142 (2011) 292300

299

4. Conclusion The results of this study are twofold. The rst part concerns the syntheses and the textural properties of silica-based materials containing di-alkyl imidazolium fragments. In this area, our results can be summarized as follows: (1) The hydrolysis polycondensation of the disilylated di-propyl imidazolium precursor 1 in the presence of anionic surfactants allows to synthesize nanostructured surfactant-silica hybrid nanocomposites. These materials show limited textural properties due to the high exibility of the propyl chains of the precursor. After elimination of the surfactant by washing, the resulting silica hybrid materials are nonstructured and non-porous. (2) Porous materials can be obtained by co-condensation approaches with TEOS. Templating with anionic surfactants allows to obtain structured nanocomposites both from the pure precursor 1 and from hydrolysis polycondensation mixtures containing high proportions of the ionic precursor. In contrast, templating under conventional conditions using cationic or neutral surfactants, allows the formation of nanostructured materials only from hydrolysis-polycondensation mixtures containing a large proportion of the silica network former TEOS. All materials obtained from co-condensation reactions show sufcient mechanical stability to avoid textural collapse during the washing step and allow to maintain porosity in the resulting materials. These approaches two are complementary: templating with anionic surfactants allows to synthesize porous materials containing high amount of the ionic substructure whereas templating reactions with non-ionic surfactants yield structured materials containing low quantities of the organo-ionic groups. The second part of our study concerns the development of new templating strategies for the synthesis of nanostructured silica hybrid materials and the use of anionic surfactant as structure directing agents. Anionic surfactants have already been used for the synthesis of nanostructured silica by the aid of silylated ammonium species [47]. In the eld of silica hybrid materials, we recently observed that nanostructured silica hybrid materials containing cationic ammonium substructures can efciently be synthesized by micelle templating with anionic surfactants [32]. Here, we report that micelle templating with anionic surfactants is also an efcient tool for the synthesis of silica-based materials containing di-alkyl imidazolium substructures. This approach is of particular interest as for the rst time, the formation of nanostructured phases with regular architectures is governed by interactions involving the organic fragment of the precursor. For this reason, this approach is of high interest for the synthesis of a large series of original silica hybrid materials bearing ionic substructures which have high potential in the elds of catalysis, separation and detection.

Appendix A. Supplementary data Supplementary data associated with this article can be found, in the online version, at doi:10.1016/j.micromeso.2010.12.014. References
[1] S. Inagaki, S. Guan, Y. Fukushima, T. Ohsuna, O. Terasaki, J. Am. Chem. Soc. 121 (1999) 96119614. [2] T. Asefa, M.J. MacLachan, N. Coombs, G.A. Ozin, Nature 402 (1999) 867871. [3] B.J. Melde, B.T. Holland, C.F. Blanford, A. Stein, Chem. Mater. 11 (1999) 3302 3308. [4] Y. Wan, D.Y. Zhao, Chem. Rev. 107 (2007) 28212860. [5] F. Hoffmann, M. Cornelius, J. Morell, M. Froba, Angewandte ChemieInternational Edition 45 (2006) 32163251. [6] O. Muth, C. Schellbach, M. Froba, Chem. Commun. (2001) 20322033. [7] C. Vercaemst, M. Ide, P.V. Wiper, J.T.A. Jones, Y.Z. Khimyak, F. Verpoort, P. Van Der Voort, Chem. Mater. 21 (2009) 57925800. [8] S. Bracco, A. Comotti, P. Valsesia, B.F. Chmelka, P. Sozzani, Chem. Commun. (2008) 47984800. [9] O. Olkhovyk, M. Jaroniec, J. Am. Chem. Soc. 127 (2005) 6061. [10] K. Landskron, G.A. Ozin, Science 306 (2004) 15291532. [11] W. Whitnall, L. Cademartiri, G.A. Ozin, J. Am. Chem. Soc. 129 (2007) 15644 15649. [12] A. Ide, R. Voss, G. Scholz, G.A. Ozin, M. Antonietti, A. Thomas, Chem. Mater. 19 (2007) 26492657. [13] A. Kuschel, S. Polarz, J. Am. Chem. Soc. 132 (2010) 65586565. [14] S. MacQuarrie, M.P. Thompson, A. Blanc, N.J. Mosey, R.P. Lemieux, C.M. Crudden, J. Am. Chem. Soc. 130 (2008) 1409914101. [15] Q.H. Yang, J. Liu, L. Zhang, C. Li, J. Mater. Chem. 19 (2009) 19451955. [16] V. Rebbin, R. Schmidt, M. Froba, Angewandte Chemie-International Edition 45 (2006) 52105214. [17] P.N. Minoofar, R. Hernandez, S. Chia, B. Dunn, J.I. Zink, A.C. Franville, J. Am. Chem. Soc. 124 (2002) 1438814396. [18] N. Mizoshita, Y. Goto, M.P. Kapoor, T. Shimada, T. Tani, S. Inagaki, Chemistry-A European Journal 15 (2009) 219226. [19] M.H. Valkenberg, C. deCastro, W.F. Holderich, Green Chem. 4 (2002) 8893. [20] M.H. Valkenberg, C. deCastro, W.F. Holderich, Applied Catalysis A-General 215 (2001) 185190. [21] C. DeCastro, E. Sauvage, M.H. Valkenberg, W.F. Holderich, J. Catal. 196 (2000) 8694. [22] C.P. Mehnert, R.A. Cook, N.C. Dispenziere, M. Afeworki, J. Am. Chem. Soc. 124 (2002) 1293212933. [23] C.P. Mehnert, E.J. Mozeleski, R.A. Cook, Chem. Commun. (2002) 30103011. [24] M. Alvaro, B. Ferrer, H. Garcia, F. Rey, Chem. Commun. (2002) 20122013. [25] M. Alvaro, B. Ferrer, V. Fornes, H. Garcia, Chem. Commun. (2001) 25462547. [26] N. Mizoshita, K. Yamanaka, T. Shimada, T. Tani, S. Inagaki, Chem. Commun. (2010) 92359237. [27] C.S.J. Cazin, M. Veith, P. Braunstein, R.B. Bedford, Synthesis-Stuttgart (2005) 622626. [28] B. Lee, H.J. Im, H.M. Luo, E.W. Hagaman, S. Dai, Langmuir 21 (2005) 53725376. [29] M. Litschauer, M.A. Neouze, J. Mater. Chem. 18 (2008) 640646. [30] B. Karimi, D. Elhamifar, J.H. Clark, A.J. Hunt, Chemistry-A European Journal 16 (2010) 80478053. [31] B. Karimi, A. Maleki, D. Elhamifar, J.H. Clark, A.J. Hunt, Chem. Commun. 46 (2010) 69476949. [32] T.P. Nguyen, P. Hesemann, T.M.L. Tran, J.J.E. Moreau, J. Mater. Chem. 20 (2010) 39103917. [33] A. El Kadib, P. Hesemann, K. Molvinger, J. Brandner, C. Biolley, P. Gaveau, J.J.E. Moreau, D. Brunel, J. Am. Chem. Soc. 131 (2009) 28822892. [34] T.P. Nguyen, P. Hesemann, P. Gaveau, J.J.E. Moreau, J. Mater. Chem. 19 (2009) 41644171. [35] V. Polshettiwar, P. Hesemann, J.J.E. Moreau, Tetrahedron 63 (2007) 6784 6790. [36] B. Gadenne, P. Hesemann, V. Polshettiwar, J.J.E. Moreau, Eur. J. Inorg. Chem. (2006) 36973702. [37] B. Gadenne, P. Hesemann, J.J.E. Moreau, Chem. Commun. (2004) 17681769. [38] R. Ciriminna, P. Hesemann, J.J.E. Moreau, M. Carraro, S. Campestrini, M. Pagliaro, Chemistry-A European Journal 12 (2006) 52205224. [39] G.E. Fryxell, J. Liu, T.A. Hauser, Z.M. Nie, K.F. Ferris, S. Mattigod, M.L. Gong, R.T. Hallen, Chem. Mater. 11 (1999) 21482154. [40] B. Lee, L.L. Bao, H.J. Im, S. Dai, E.W. Hagaman, J.S. Lin, Langmuir 19 (2003) 42464252. [41] Y.H. Ju, O.F. Webb, S. Dai, J.S. Lin, C.E. Barnes, Ind. Eng. Chem. Res. 39 (2000) 550553. [42] J. Yu, J.L. Shi, H.R. Chen, J.N. Yan, D.S. Yan, Microporous Mesoporous Mater. 46 (2001) 153162. [43] W.P. Guo, J.Y. Park, M.O. Oh, H.W. Jeong, W.J. Cho, I. Kim, C.S. Ha, Chem. Mater. 15 (2003) 22952298. [44] V. Polshettiwar, P. Hesemann, J.J.E. Moreau, Tetrahedron Lett. 48 (2007) 5363 5366. [45] H.H. Zhao, N.Y. Yu, J.Q. Wang, D.Y. Zhuang, Y. Ding, R. Tan, D.H. Yin, Microporous Mesoporous Mater. 122 (2009) 240246.

Acknowledgements Thy Phuong Nguyen thanks the Agence Universitaire de la Francophonie for a doctoral scholarship. Peter Hesemann thanks the Groupement de Recherche PARIS and the Rseau de Recherche 3, Chimie pour le Dveloppement Durable of the CNRS for nancial support. The authors are indebted to Philippe Gaveau (Institut Charles Gerhardt de Montpellier) for solid state NMR measurements.

300

T.P. Nguyen et al. / Microporous and Mesoporous Materials 142 (2011) 292300 [50] [51] [52] [53] S.L. Burkett, S.D. Sims, S. Mann, Chem. Commun. (1996) 13671368. C.E. Fowler, S.L. Burkett, S. Mann, Chem. Commun. (1997) 17691770. T. Yokoi, H. Yoshitake, T. Tatsumi, Chem. Mater. 15 (2003) 45364538. D.Y. Zhao, J.L. Feng, Q.S. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D. Stucky, Science 279 (1998) 548552. [54] D.Y. Zhao, Q.S. Huo, J.L. Feng, B.F. Chmelka, G.D. Stucky, J. Am. Chem. Soc. 120 (1998) 60246036.

[46] A.E. Garcia-Bennett, N. Kupferschmidt, Y. Sakamoto, S. Che, O. Terasaki, Angewandte Chemie-International Edition 44 (2005) 53175322. [47] S. Che, A.E. Garcia-Bennett, T. Yokoi, K. Sakamoto, H. Kunieda, O. Terasaki, T. Tatsumi, Nat. Mater. 2 (2003) 801805. [48] S. Che, Z. Liu, T. Ohsuna, K. Sakamoto, O. Terasaki, T. Tatsumi, Nature 429 (2004) 281284. [49] J.S. Beck, J.C. Vartuli, W.J. Roth, M.E. Leonowicz, C.T. Kresge, K.D. Schmitt, C.T.W. Chu, D.H. Olson, E.W. Sheppard, S.B. McCullen, J.B. Higgins, J.L. Schlenker, J. Am. Chem. Soc. 114 (1992) 1083410843.

You might also like