Unusual Catalytic Effect of the Two-Dimensional Molecular Space with

A novel organic–inorganic hybrid 2D molecular space with regular triphenylphosphine groups (triphenylphosphineamidephenylsilica, PPh3APhS) was ...
0 downloads 0 Views 2MB Size
ARTICLE pubs.acs.org/Langmuir

Unusual Catalytic Effect of the Two-Dimensional Molecular Space with Regular Triphenylphosphine Groups Chen Chen,† Xue Shao,† Ken Yao,*,† Jian Yuan,*,‡ Wenfeng Shangguan,‡ Toshikazu Kawaguchi,§,|| and Katsuaki Shimazu§,|| †

Department of Chemistry, Shanghai University, Shanghai 200444, PR China Research Center for Combustion & Environmental Technology, Shanghai Jiao Tong University, Shanghai 200240, PR China § Division of Environmental Materials Science, Graduate School of Environmental Science and Section of Materials Science, Faculty of Environmental Earth Science, Hokkaido University, Sapporo 060-0810, Japan )



bS Supporting Information ABSTRACT:

A novel organicinorganic hybrid 2D molecular space with regular triphenylphosphine groups (triphenylphosphineamidephenylsilica, PPh3APhS) was successfully synthesized through grafting triphenylphosphine groups in the 2D structure of layered aminophenylsilica dodecyl sulfate (APhTMS-DS), which was developed in our previous research, with regular ammonium groups. The 2D structures were kept after the grafting reaction of triphenylphosphine groups in PPh3APhS. The catalytic potentials of 2D molecular space with regular triphenylphosphine groups were investigated. An unusual catalytic effect was found in a carbon phosphorus ylide reaction. The PPh3-catalyzed reaction of modified allylic compounds, including bromides and chlorides with tropone yielded a [3 + 6] annulation product. However, an unusual [8 + 3] cycloadduct was obtained in the reaction of modified allylic compounds, including bromides and chlorides with tropone catalyzed by PPh3APhS. Otherwise, the stable catalytic intermediate was successfully separated, and the reaction activity of the catalytic intermediate was confirmed in the reaction of modified allylic compounds with tropone catalyzed by PPh3APhS. This research is the first successful example of directly influencing catalytic reaction processes and product structures by utilizing the chemical and geometrical limits of 2D molecular spaces with regular catalyst molecules and affords a novel method for controlling catalytic reaction processes and catalyst design.

’ INTRODUCTION Molecular reaction processes are usually influenced by the molecular environment because of the interaction between molecules. Thus, it is important to afford an appropriate reaction environment for molecular reaction processes. However, it is difficult to control the molecular environment in an open space because of the huge number of molecules. A nanospace material with special chemical and geometrical structures can afford a reaction space with chemical and geometrical limits for molecules and influence the molecular reaction processes15 Twodimensional layer spaces of layered materials are expandable with the size of the intercalated molecules. The intercalated molecules usually form monolayers or bilayers in the interlayer space of layered materials independent of the molecular size as a result of the interaction between the layer structure and guest molecules. r 2011 American Chemical Society

Thus, 2D molecular spaces with regular molecular structure may be promising reaction spaces for influencing and controlling molecular reaction processes by the utilization of the geometrical and chemical limits of the interlayer space with special structure.629 Previously, we developed novel organicinorganic hybrid 2D layered spaces with regular ammonium groups in the structure (layered aminophenylsilica APhTMS-DS) using organosilanes with an amino group template with an anionic surfactant (sodium dodecyl sulfate, SDS) under acidic conditions as illustrated in Scheme 1.3032 The 2D lamellar organicsilica structures were also Received: June 20, 2011 Revised: August 24, 2011 Published: September 09, 2011 11958

dx.doi.org/10.1021/la2023083 | Langmuir 2011, 27, 11958–11965

Langmuir

ARTICLE

Scheme 1. Schematic Illustrations of the APhTMS-DS (a) 2D Structure and (b) Molecule Structure

reported to be obtained by the direct concentration of n-alkyltrichlorosilanes,33 organotrialkoxysilanes,34 3-aminopropyltrimethoxysilane,35 and cyanoalkyltrialkoxysilane.36 The inorganic part of the layered APhTMS-DS was crystalline hexagonal SiO, and the ammonium groups were considered to be alternately arranged on both sides of the sheet (Scheme 1b). The layered APhTMS-DS with regular ammonium groups exhibited a stable 2D structure and better grafting reaction properties. The regular ammonium groups afforded regular reactable points in the 2D structure and made it possible to develop a series 2D molecular spaces with various chemical and geometrical structures. We have also developed an organicinorganic hybrid 2D molecular space material with regular pyridine groups (pyridine-4-amidephenylsilica, PAPhS) through grafting pyridine groups in the 2D structure of layered APhTMS-DS, and the catalytic capability of PAPhS was confirmed by utilizing Knoevenagel condensation reactions.37 It is known that the chemical and geometrical structures of catalysts directly influence the catalytic reaction processes and product structures. Otherwise, the regular array of catalyst molecules is important in influencing catalytic reaction processes. The organicinorganic hybrid 2D molecular space with regular catalyst molecules can be considered to be an accumulation of 2D functional surfaces with regular catalyst molecules. The 2D molecular spaces with regular catalyst molecules can offer regular

reaction surfaces in a confined space for catalytic reaction processes in which the unusual catalytic effect may be expected because of the chemical and geometrical limits. Here we report the first successful example of directly influencing the catalytic reaction process and the structures of the catalytic reaction product by utilizing the chemical and geometrical limits of the 2D molecular spaces with regular catalyst molecules. The catalytic behavior of the 2D molecular space with regular triphenylphosphine groups (layered triphenylphosphine-amidephenylsilica, PPh3APhS), which was synthesized through grafting triphenylphosphine groups in the structure of layered APhTMS-DS with regular ammonium groups using 4-(diphenylphosphino)-benzoyl chloride, was found to be different from that of free triphenylphosphine (PPh3) in allylic-phosphorus ylide reactions. The PPh3catalyzed reaction of modified allylic compounds, including bromides and chlorides with tropone yielded a [3 + 6] annulation product.38 However, an unusual [8 + 3] cycloadduct was obtained in the reaction of modified allylic compounds, including bromides and chlorides with tropone catalyzed by PPh3APhS. The cycloaddition reactions of tropone have been reported to yield [4 + 6] and [4 + 2] cycloadducts with dienes,3941 [8 + 2] cycloadducts with tetracarbethoxyallene, phenylsulfonylallene,42 and allenic ketone/ester,41,43 and [3 + 6] cycloadducts with 2-[(trimethylsilyl)methyl]allyl carboxylates44 and modified allylic 11959

dx.doi.org/10.1021/la2023083 |Langmuir 2011, 27, 11958–11965

Langmuir

ARTICLE

Scheme 2. Schematic Illustrations of PPh3APhS: (a) 2D Structure and (b) Molecule Structure

compounds.38 The formation of an unusual PPh3APhS-catalyzed [8 + 3] annulation product of tropone exhibits the potential of influencing catalytic reaction processes and product structures by utilizing the chemical and geometrical limits of 2D molecular spaces with regular catalyst molecules and affords a novel method for controlling catalytic reaction processes and catalyst design.

’ EXPERIMENTAL SECTION All reagents were purchased from Sigma-Aldrich and ABCR GmbH & Co. KG (Germany). Layered APhTMS-DS was synthesized by the very slow titration (3 mL/min) of hydrochloric acid (0.5 mol 3 dm3) into 40 mL of an aqueous solution of p-aminophenyltrimethoxysilane (APhTMS, 2.78 mmol) and sodium dodecyl sulfate (SDS, 2.92 mmol). Then, the suspension was stirred at room temperature for 3 weeks and the pH value was controlled to between 2 and 3 as described previously.3032 The precipitates were filtered and washed with deionized water and ethanol and then dried in vacuum. PPh3APhS was synthesized via the following procedure: First, 0.5 g of 4-(diphenylphosphino)benzoic acid was treated with 15 mL of SOCl2 at 40 °C for 3 h to form 4-(diphenylphosphino)benzoyl chloride, and then the 4-(diphenylphosphino)benzoyl chloride was washed with benzene three times after residual SOCl2 was removed under vacuum. Then, 0.2 g of APhTMS-DS was added to a 20 mL toluene solution of 4(diphenylphosphino)benzoyl chloride, and the suspension was stirred at 50 °C for 48 h. The precipitate was filtered and washed with THF and then stirred with 200 mL of 0.3 M HCl for 48 h in order to remove the residual DS anion. The final precipitate was filtered and washed with deionized water and EtOH and then dried in vacuum. The 13C CP/MAS and 31P CP/MAS NMR spectra were recorded on a Bruker DMX 300 MHz spectrometer using TMS and orthophosphoric acid as references, respectively. FTIR spectra were obtained using an Avatar 370 spectrophotometer from Nicolet. Sample preparation involved dispersing and gently grinding the powder in KBr. X-ray diffraction (XRD) data on the powder samples were recorded with an X-ray diffractometer (Rigaku, D/max 2550) using Cu Kα radiation (0.1541 nm) under the conditions of 40 kV and 30 mA. The amount of P in PPh3APhS was measured as follows: first the sample was dissolved in a 0.2 M NaOH aqueous solution at room temperature. Then, the amount

of P was determined by inductively coupled plasma (ICP) emission spectroscopy (Thermo ICAP 6300). Chemical analysis was performed by elementary organic microanalysis for C, N, and S in a VerioEL element analyzer. The reaction conditions of the catalysis reaction of ethyl 2-bromomethyl-2-propenoate or ethyl 2-chloromethyl-2-propenoate with tropone are as follows: Under N2, a solution of ethyl 2-bromomethyl2-propenoate (0.30 mmol) or ethyl 2-chloromethyl-2-propenoate (0.30 mmol) and tropone (0.25 mmol) in toluene (2 mL) was added with a syringe pump to a mixture of PPh3APhS (0.06 mmol) and K2CO3 (0.50 mmol) in toluene (1.0 mL), and the reaction mixture was stirred at 110 °C for 2 h. After the reaction was completed as monitored by TLC, the mixture was filtered and eluted with ethyl acetate. Then, the filtrate was removed under reduced pressure, and the residue was purified by flash chromatography on silica gel (1:20 ethyl acetate/petroleum ether) to give the [8 + 3] cycloadduct as a colorless oil. PPh3APhS was recycled through washing with deionized water to remove potassium carbonate. The catalytic activity of PPh3APhS had not significantly decreased after five cycles of the catalytic reaction. 1 H and 13C NMR spectra were recorded on a Bruker MSL-500WB spectrometer in CDCl3. Mass spectra were recorded on an HP 5973 network mass-selective detector. The spectral data of the catalytic products are as follows. Compound [8 + 3] cycloadduct: oil. 1H NMR (500 MHz, CDCl3): δ 7.18 (t, J = 7.5 Hz, 1H), 6.71 (t, J = 7.5 Hz, 1H), 6.59 (d, J = 8.0 Hz, 1H), 6.28 (s, 1H), 5.58 (s, 1H), 4.25 (q, 2H), 4.19 (s, 2H), 1.33 (t, J = 7.5 Hz, 3H), 1.25 (s, 2H). 13C NMR (125 MHz, CDCl3): δ 166.58, 147.76, 135.27, 129.31, 125.02, 125.02, 117.00, 111.98, 96.26, 61.02, 51.35, 29.85, 14.37. IR (neat): ν 2960, 2872, 1714, 1634, 1600, 1506, 1388, 1370, 1293, 1261 cm1. MS (m/z): 218 (M+), 204, 149 (100), 104, 77, 55. HRMS-EI calcd for C13H14O3 218.0943, found 218.0940 (Supporting Information).

’ RESULTS AND DISCUSSION The 2D molecular space with regular triphenylphosphine groups (PPh3APhS) was synthesized through grafting triphenylphosphine groups in the structure of layered APhTMS-DS with regular ammonium groups using 4-(diphenylphosphino)benzoyl chloride as illustrated in Scheme 2. First, the 13C CP/MAS NMR 11960

dx.doi.org/10.1021/la2023083 |Langmuir 2011, 27, 11958–11965

Langmuir

ARTICLE

Figure 2.

Figure 1. 13C CP/MAS NMR spectra of (a) APhTMS-DS and (b) PPh3APhS before treatment with HCl and (c) PPh3APhS after treatment with HCl.

spectra of PPh3APhS were used to affirm the formation of PPh3APhS. The 13C CP/MAS NMR spectra of APhTMS-DS and PPh3APhS before and after treatment with HCl are shown in Figure 1. The resonance peaks from 5 to 70 ppm observed in the 13 C CP/MAS NMR spectrum of APhTMS-DS (Figure 1a) were

31

P CP/MAS NMR spectrum of PPh3APhS.

attributed to the methyl and methylenes of dodecyl sulfate (DS) existing in the layer space of APhTMS-DS. The resonance peaks at 125 and 135 ppm observed in the 13C CP/MAS NMR spectrum of APhTMS-DS (Figure 1a) were attributed to the superposition of resonances of carbon species in the aromatic rings of aminophenylsilica because of the aromatic rings fixed in the framework of silica as a layer plate.3032 The 13C CP/MAS NMR spectra of PPh3APhS before and after treatment with HCl shows resonance peaks at 167 ppm (Figure 1b,c) that were assigned to carbonyl groups in amide C(dO)NH formed between ammonium groups in APhTMS-DS and acylchloride groups in 4-(diphenylphosphino)benzoyl chloride as shown in Scheme 2b. Two smaller resonance peaks at 68 and 25 ppm were observed in the 13C CP/MAS NMR spectrum of PPh3APhS before treatment with HCl instead of the sharp resonance peaks from 5 to 70 ppm for APhTMS-DS (Figure 1a). It was considered that the ammonium groups in APhTMS-DS were not completely reacted with 4-(diphenylphosphino)benzoyl chloride and that some DS anions remained in PPh3APhS before treatment with HCl. The resonance peaks assigned to the methyl and methylenes of DS completely disappeared in the 13C CP/MAS NMR spectrum of PPh3APhS after treatment with HCl (Figure 1c). The residual DS anion in PPh3APhS was replaced with Cl after treatement with HCl. The resonance peaks at around 132 ppm observed in the 13C CP/MASNMR spectrum of PPh3APhS before and after treatment with HCl (Figure 1b,c) were attributed to the superposition of resonances of carbon species in the aromatic rings of PPh3APhS (Scheme 2b). The results of the 13C CP/MAS NMR spectrum of PPh3APhS indicated that the triphenylphosphine groups were successfully grafted into the 2D space of layered aminophenylsilica to form PPh3APhS. The existence of PPh3 also could be confirmed through the 31P CP/MAS NMR spectrum of PPh3APhS as shown Figure 2. The 31P CP/MAS NMR spectrum of PPh3APhS shows only one broad resonance at 29 ppm. This resonance can be assigned to the P atom of PPh3 in PPh3APhS. The small symmetric peaks on both sides are attributed to spinning side bands. Only one peak observed in the 31P CP/MAS NMR spectrum of PPh3APhS indicates that PPh3 was regularly grafted into the 2D space of layered aminophenylsilica. The formation of PPh3APhS was further confirmed using IR spectra. The IR spectrum of PPh3APhS after treatment with HCl 11961

dx.doi.org/10.1021/la2023083 |Langmuir 2011, 27, 11958–11965

Langmuir

ARTICLE

exhibited a new vibrational band at 1660 cm1 compared with that of layered APhTMS-DS (Figure 3), which was assigned to carbonyl groups in amide formed between the ammonium groups in layered APhTMS-DS and acylchloride groups in 4-(diphenylphosphino)benzoyl chloride as shown in Scheme 2b. The stronger bands at 2851, 2920, and 2953 cm1, which were assigned to vibrational bands of methyl and methylenes of dodecyl sulfate (DS) in APhTMS-DS (Figure 3a), were not observed in PPh3APhS after treatment with HCl (Figure 3b). It is considered that the DS anions in layered APhTMS-DS disappeared and that the triphenylphosphine groups were successfully grafted into layered aminophenylsilica.

Figure 3. IR spectra of (a) APhTMS-DS, (b) PPh3APhS, and (c) PPh3APhS-ethyl 2-methyl-2-propenoate (intermediate A0 ).

Figure 4. Powder XRD patterns of (a) APhTMS-DS and (b) PPH3APhS.

The X-ray diffraction patterns were used to examine the 2D structure of PPh3APhS. The XRD patterns of APhTMS-DS and PPh3APhS are shown in Figure 4. The X-ray diffraction peak of PPh3APhS assigned to the 001 reflection was shifted from 2θ = 2.2° in APhTMS-DS to 2θ = 2.6°, corresponding to an interlayer distance shift from 4 nm in APhTMS-DS to 3.4 nm (Figure 4). The change of the interlayer distance was consistent with the change in the molecular length from APhTMS-DS to PPh3APhS as shown in Schemes 1b and 2b. The diffraction peaks observed at 2θ = 5.3° (1.7 nm) and 2θ = 8.1° (1.1 nm) in PPh3APhS were assigned to the 002 and 003 reflections. The better XRD response indicated that the 2D layered structure had been retained in PPh3APhS after the grafting reaction and that the triphenylphosphine groups were regularly arranged in the 2D structure. Otherwise, the rigid PPh3 molecules were considered to increase in the regularity of the 2D layered structures of PPh3APhS. The elemental analytical result of layered PPh3APhS is shown in Table 1. No sulfur atom was observed in PPh3APhS after treatment with HCl. The result of no sulfur element in PPh3APhS further indicated that DS in the 2D molecular space of PPh3APhS was completely replaced by 4-(diphenylphosphino)benzoyl chloride and Cl. The 6.6 wt % of phosphorus in PPh3APhS was determined by ICP. The formula of PPh3APhS was proposed to be SiO1.5C6H4(NHCOC6H4P(C6H5)2)0.826(NH3Cl)0.174. It is known that the chemical and geometrical structures of the catalyst directly influence the catalytic reaction processes and product structures. Two-dimensional molecular spaces with regular catalyst molecules form a catalytic reaction space with geometrical and chemical limits and may influence the catalytic reaction processes and production structures. Thus, the unusual catalytic effect may be obtained in the 2D molecular spaces with regular catalyst molecules. An annulation reaction of ethyl 2-bromomethyl-2-propenoate and ethyl 2-chloromethyl-2-propenoate with tropone was used to investigate the catalytic effect of the 2D molecular spaces with regular triphenylphosphine groups. The annulation reaction of 2-bromomethyl-2-propenoate with tropone under the catalysis of PPh3 yielded the [3 + 6] cycloadduct (Scheme 3).38 However, an unusual [8 + 3] cycloadduct was obtained in the annulation reaction of 2-bromomethyl-2-propenoate with tropone catalyzed by PPh3APhS (Scheme 3). Otherwise, no reaction occurred when 2-bromomethyl-2-propenoate and tropone were mixed with APhTMSDS and potassium carbonate. Apparently, the 2D molecular space with regular triphenylphosphine groups exhibited an unusual catalytic effect in the annulation reaction of ethyl 2-bromomethyl-2-propenoate with tropone. The unusual catalytic behavior in the annulation reaction of ethyl 2-bromomethyl-2-propenoate with tropone catalyzed by PPh3APhS was considered because of the catalytic mechanism of PPh3APhS, which is different from that of free PPh3. Tropone is a molecule with interesting electronic properties. It was suggested that a nucleophilic attack on tropone should occur preferentially at C(2) and C(7) via frontier molecular orbital theory.45 The mechanism of the PPh3-catalyzed [3 + 6] annulation reaction of ethyl 2-bromomethyl-2-propenoate with

Table 1. Elemental Analytical Results of PPh3APhS after Treatment with HCl sample

C (wt %)

N (wt %)

S (wt %)

P (wt %)

proposed formula

PPh3APhS

67.20

3.63

0

6.6

SiO1.5C6H4(NHCOC6H4P(C6H5)2)0.826 (NH3Cl)0.174

11962

dx.doi.org/10.1021/la2023083 |Langmuir 2011, 27, 11958–11965

Langmuir tropone was considered as shown in Scheme 4. The formation of [3 + 6] cycloadducts can be rationalized in terms of the formation of the phosphonium salt of ethyl 2-bromomethyl-2-propenoate through an SN2 reaction, which was deprotonated by potassium carbonate to afford A. Then, subsequent nucleophilic C2 addition of A to tropone yielded zwitterionic intermediate B, followed by cyclization from intramolecular conjugate addition leading to C. Finally, the elimination of PPh3 completed the [3 + 6] catalytic cycle (Scheme 4).38 A different [8 + 3] cycloadduct was yielded in the annulation reaction of ethyl 2-bromomethyl-2-propenoate with tropone catalyzed by PPh3APhS. It was considered to be due to the effect of the chemical and geometrical limits of 2D molecular space with regular triphenylphosphine groups. The reaction might be initiated by the formation of the phosphonium salt of ethyl 2-bromomethyl-2-propenoate in the 2D molecular space of PPh3APh, which was deprotonated by potassium carbonate to afford A0 (Scheme 5). Then, A0 attacked at C2 of tropone, leading to B0 , which was followed by the nucleophilic attack of the Scheme 3

Scheme 4

ARTICLE

negatively charged oxygen at the central carbon of B0 , leading to C0 . Finally, the elimination of PPh3APhS formed the [8 + 3] catalytic cycle (Scheme 5).41,43 The formation of an unusual PPh3APhS-catalyzed [8 + 3] annulation product exhibited the potential to influence catalytic reaction processes and production structures utilizing the chemical and geometrical limits of 2D molecular spaces with regular catalyst molecules The existence of catalytic intermediate A0 could be confirmed used a designed experiment. First, PPh3APhS was mixed with ethyl 2-bromomethyl-2-propenoate and potassium carbonate in toluene and stirred at 110 °C for 1 h under N2. Then the mixture was filtered and washed with ethyl acetate and deionized water to remove redundant ethyl 2-bromomethyl-2-propenoate and potassium carbonate and dried in vacuum to obtain separated PPh3APhS. The IR spectrum of separated PPh3APhS after reaction with ethyl 2-bromomethyl-2-propenoate and potassium carbonate showed a new vibrational peak at 1707 cm1 (Figure 3c), which was assigned to carbonyl groups of the ester group in ethyl 2-bromomethyl-2-propenoate. When separated PPh3APhS was reacted with tropone under N2 again, the [8 + 3] cycloadduct was obtained and the IR spectrum changed to initial PPh3APhS as shown in Figure 1b. Apparently, catalytic intermediate A0 (PPh3APhS-ethyl 2-methyl-2-propenoate) was formed through the reaction of PPh3APhS with ethyl 2-bromomethyl2-propenoate and potassium carbonate. Otherwise, no change was observed in the IR spectrum of separate PPh3APhS, when PPh3APhS was reacted with ethyl 2-bromomethyl-2-propenoate without potassium carbonate. Apparently, intermediate A0 could not be formed without potassium carbonate. The deprotonation process of base was found to be a crucial reaction step in the formation of active intermediate A0 . The separation of the stable catalytic intermediate and the confirmation of reaction activity may be very helpful in investigating the catalytic reaction mechanism and designing the catalytic reaction processes. The same catalytic reaction result was obtained in the annulation reaction of ethyl 2-chloromethyl-2-propenoate with tropone as shown in Table 2. An [3 + 6] and [8 + 3] cycloadduct was obtained in the annulation reaction of ethyl 2-chloromethyl-2-propenoate with tropone catalyzed by PPh3 and PPh3APhS, respectively. The existence and reaction activity of intermediate A0 were also confirmed in the annulation reaction of ethyl 2-chloromethyl2-propenoate with tropone catalyzed by PPh3APhS. The different catalytic behavior between free PPh3 and PPh3APhS in the annulation reaction of ethyl 2-bromomethyl-2propenoate and ethyl 2-chloromethyl-2-propenoate with tropone

Scheme 5

11963

dx.doi.org/10.1021/la2023083 |Langmuir 2011, 27, 11958–11965

Langmuir

ARTICLE

Table 2. Annulation Reaction Results of Tropone with Ethyl 2-Bromomethyl-2-propenoate and Ethyl 2-Chromomethyl-2propenoatea catalyst

a

product

yields (%)

ethyl 2-bromomethyl-2-propenoate

PPh3

[3 + 6] cycloadduct

72

ethyl 2-bromomethyl-2-propenoate ethyl 2-chromomethyl-2-propenoate

PPh3APhS PPh3

[8 + 3] cycloadduct [3 + 6] cycloadduct

60 64

ethyl 2-chromomethyl-2-propenoate

PPh3APhS

[8 + 3] cycloadduct

52

The [3 + 6] and [8 + 3] cycloadducts were not observed in the reactions catalyzed by PPh3APhS and PPh3, respectively.

was considered because of the effect of chemical and geometrical limits of regular triphenylphosphine groups in 2D molecular space. This research is the first successful example of directly influencing catalytic reaction processes and product structures utilizing the chemical and geometrical limits of 2D molecular spaces with regular catalyst molecules. It exhibits the potential of influencing catalytic reaction processes and production structures utilizing the chemical and geometrical limits of 2D molecular spaces with regular catalyst molecules and affords a novel method for controlling catalytic reaction processes and catalyst design.

’ CONCLUSIONS A novel organicinorganic hybrid 2D molecular space with regular triphenylphosphine groups (triphenylphosphine-amidephenylsilica, PPh3APhS) was synthesized through grafting triphenylphosphine groups in the 2D structure of layered APhTMS-DS with regular ammonium groups, which was developed in our previous research, using 4-(diphenylphosphino)benzoyl chloride. The catalytic behavior of PPh3APhS was found to be different from that of free triphenylphosphine (PPh3) in a carbonphosphorus ylide reaction. The PPh3-catalyzed reactions of ethyl 2-bromomethyl-2-propenoate and ethyl 2-chloromethyl-2-propenoate with tropone yielded a [3 + 6] annulation product. However, an unusual [8 + 3] cycloadduct was obtained in the same reaction catalyzed by PPh3APhS. The stable catalytic intermediate was successfully separated, and the reaction activity of the catalytic intermediate was confirmed in the reactions of ethyl 2-bromomethyl-2-propenoate and ethyl 2-chloromethy-2propenoate. ’ ASSOCIATED CONTENT Supporting Information. 1H and 13C NMR and COSY spectra of the [8 + 3] cycloadduct catalyzed by PPh3APhS. This material is available free of charge via the Internet at http:// pubs.acs.org.

bS

’ AUTHOR INFORMATION Corresponding Author

*Tel: +86-21-6613-4857. Fax: +86-21-5638-8125. E-mail: [email protected] (K.Y.); [email protected] (J.Y.).

’ ACKNOWLEDGMENT This work was supported by the National High Technology Research and Development Program of China (863 program, 2007AA05Z155) and the National Natural Science Foundation of China (20873082).

’ REFERENCES (1) Otani, W.; Kinbara, K.; Zhang, Q.; Ariga, K.; Aida, T. Chem.— Eur. J. 2007, 13, 1731. (2) Yang, M.; Sheehan, P. E.; King, W. P.; Whitman, L. J. J. Am. Chem. Soc. 2006, 128, 6774. (3) Baker, S. E.; Sawvel, A. M.; Zheng, N.; Stucky, G. D. Chem. Mater. 2007, 19, 4390. (4) Galimberti, M.; Lostritto, A.; Spatola, A.; Guerra, G. Chem. Mater. 2007, 19, 2495. (5) Georgakilas, V.; Bourlinos, A.; Gournis, D.; Tsoufis, T.; Trapalis, C.; Mateo-Alonso, A.; Prato, M. J. Am. Chem. Soc. 2008, 130, 8733. (6) Sato, H.; Yamagishi, A.; Kato, S. J. Am. Chem. Soc. 1992, 114, 10933. (7) Roland-Swanson, C.; Besse, J. P.; Leroux, F. Chem. Mater. 2004, 16, 5512. (8) Sasai, R.; Iyi, N.; Fujita, T.; Arbeloa, F. L.; Martínez, V. M.; Takagi, K.; Itoh, H. Langmuir 2004, 20, 4715. (9) Beaudot, P.; De Roy, M. E.; Besse, J. P. Chem. Mater. 2004, 16, 935. (10) Iyi, N.; Matsumoto, T.; Kaneko, Y.; Kitamura, K. Chem. Mater. 2004, 16, 2926. (11) M€uller, R.; Hrobarikova, J.; Calberg, C.; Jer^ome, R.; Grandjean, J. Langmuir 2004, 20, 2982. (12) Kooli, F.; Khimyak, Y. Z.; Alshahateet, S. F.; Chen, F. X. Langmuir 2005, 21, 8717. (13) Williams, G. R.; O’Hare, D. Chem. Mater. 2005, 17, 2632. (14) Hu, G.; O’Hare, D. J. Am. Chem. Soc. 2005, 127, 17808. (15) Reinholdt, M. X.; Kirkpatrick, R. J.; Pinnavaia, T. J. J. Phys. Chem. B 2005, 109, 16296. (16) Allada, R. K.; Pless, J. D.; Nenoff, T. M.; Navrotsky, A. Chem. Mater. 2005, 17, 2455. (17) Moujahid, E. M.; Dubois, M.; Besse, J. P.; Leroux, F. Chem. Mater. 2005, 17, 373. (18) Darder, M.; L opez-Blanco, M.; Aranda, P.; Leroux, F.; RuizHitzky, E. Chem. Mater. 2005, 17, 1969. (19) Fujita, S; Sato, H.; Kakegawa, N.; Yamagishi, A. J. Phys. Chem. B 2006, 110, 2533. (20) Leroux, F.; Raymundo-Pinero, E.; Nedelec, J. M.; Beguin, F. J. Mater. Chem. 2006, 16, 2074. (21) Feng, Y. J.; Williams, G. R.; Leroux, F.; Taviot-Gueho, C.; O’Hare, D. Chem. Mater. 2006, 18, 4312. (22) Liu, Z. P.; Ma, R. Z.; Osada, M.; Iyi, N.; Ebina, Y.; Takada, K.; Sasaki, T. J. Am. Chem. Soc. 2006, 128, 4872. (23) Xue, S. Q.; Reinholdt, M.; Pinnavaia, T. J. Polymer 2006, 47, 3344. (24) Baker, S. E.; Sawvel, A. M.; Zheng, N.; Stucky, G. D. Chem. Mater. 2007, 19, 4390. (25) Galimberti, M.; Lostritto, A.; Spatola, A.; Guerra, G. Chem. Mater. 2007, 19, 2495. (26) Xiong, Z. G.; Xu, Y. M. Chem. Mater. 2007, 19, 1452. (27) Ide, Y.; Ogawa, M. Angew. Chem., Int. Ed. 2007, 46, 8449. (28) Mouawia, R.; Mehdi, A.; Reye, C.; R. Corriu, J. P. J. Mater. Chem. 2008, 18, 2028. (29) Joshi, P. C.; Aldersley, M. F.; Delano, J. W.; Ferris, J. P. J. Am. Chem. Soc. 2010, 131, 13369. 11964

dx.doi.org/10.1021/la2023083 |Langmuir 2011, 27, 11958–11965

Langmuir

ARTICLE

(30) Yao, K.; Imai, Y.; Shi, L. Y.; Abe, E.; Adachi, Y.; Nishikubo, K.; Tateyama, H. Chem. Lett. 2004, 33, 1112. (31) Yu, F. T.; Yao, K.; Shi, L. Y.; Wang, H. Z.; Fu, Y.; You, X. Q. Chem. Mater. 2007, 19, 335. (32) Yao, K.; You, X. Q.; Shi, L. Y.; Wan, W.; Yu, F. T.; Chen, J. M. Langmuir 2008, 24, 302. (33) Parikh, A. N.; Schivley, M. A.; Koo, E.; Seshadri, K.; Aurentz, D.; Mueller, K.; Allara, D. L. J. Am. Chem. Soc. 1997, 119, 3135. (34) Loy, D. A.; Baugher, B. M.; Baugher, C. R.; Schneider, D. A.; Rahimian, K. Chem. Mater. 2000, 12, 3624. (35) Kaneko, Y.; Iyi, N.; Matsumoto, T.; Fujii, K.; Kurashima, K.; Fujita, T. J. Mater. Chem. 2003, 13, 2058. (36) Mouawia, R.; Mehdi, A.; Reye, C.; R. Corriu, J. P. J. Mater. Chem. 2008, 18, 2028. (37) Chen, J. M.; Yao, K.; Shangguan, W. F.; Yuan, J. Langmuir 2009, 25, 5993. (38) Du, Y. S.; Feng, J. Q. Lu, X. Y. Org. Lett . 2005, 7, 1987 and references therein. (39) Funk, R. L.; Bolton, G. L. J. Am. Chem. Soc. 1986, 108, 4655 and references therein. (40) Isakovic, L.; Ashenhurst, J. A.; Gleason, J. L. Org. Lett. 2001, 3, 4189 and references therein. (41) Ishar, M. P. S.; Gandhl, R. P. Tetrahedron 1993, 49, 6729. (42) Hayakawa, K.; Nishiyama, H.; Kanematsu, K. J. Org. Chem. 1985, 50, 512. (43) Kumar, K.; Kapur, A.; Ishar, M. P. S. Org. Lett. 2000, 2, 787 and references therein. (44) Trost, B. M.; Seoane, P. R. J. Am. Chem. Soc. 1987, 109, 615. (45) Biggi, G.; Deloma, F.; Pietra, F. J. Am. Chem. Soc. 1973, 95, 7101.

11965

dx.doi.org/10.1021/la2023083 |Langmuir 2011, 27, 11958–11965