You are on page 1of 7

Microporous and Mesoporous Materials 48 (2001) 285291 www.elsevier.

com/locate/micromeso

Hydroamination of 6-aminohex-1-yne over zinc based homogeneous and zeolite catalysts


Jochen Penzien, Thomas E. Mller *, Johannes A. Lercher u
Lehrstuhl f r Technische Chemie II, Technische Universitt Mnchen, Lichtenbergstrae 4, D-85747 Garching, Germany u a u Received 6 August 2000; accepted 12 March 2001

Abstract Ion-exchanged zeolites are ecient heterogeneous catalysts for the intramolecular addition of an amine HNR2 to a CC triple bond. This was shown for the cyclisation of 6-aminohex-1-yne to 2-methyl-1,2-dehydropiperidine. For this reaction zinc(II)-exchanged zeolite BEA was one of the most active catalysts. The heterogeneous catalyst was more active than the corresponding homogeneous catalyst Zn(CF3 SO3 )2 , which is concluded to be related to co-catalysis between Lewis and Brnsted acid sites present in the zeolite. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Hydroamination; Beta zeolite; Addition; Amine; Alkyne

1. Introduction The addition of amines to carboncarbon double and triple bonds is of special interest as a new and atom ecient transformation [1,2]. It allows the formation of CN bonds in simple as well as complex nitrogen containing molecules. Despite the economic advantages relative to traditional routes to amines [3], only one process, the synthesis of t-butylamine from ammonia and isobutene in the presence of a zeolite catalyst, has been commercialised by BASF [4]. Ecient catalysts based on late transition metals compounds are only known for the addition of amines to alkynes or activated CC double bonds. Examples are the

* Corresponding author. Tel.: +49-89-289-13538; fax: +4989-289-13544. E-mail address: thomas.mueller@ch.tum.de (T.E. Mller). u

addition of anilines to terminal alkynes and norbornene with Ru(0) [5] and Ir(I) [6] complexes, the amination of vinylpyridines with a Rh(I) catalyst [7] and the cyclisation of aminoalkynes with Zn(II) salts [8]. Zinc cations also catalyse the reaction between ammonia and acetylene. However, only low selectivities have been observed [9]. In general, the increasing interest in new catalysts for the direct addition of an amine NH bond across carboncarbon multiple bonds is reected in the rising number of patents on heterogeneous hydroamination catalysts [10 15]. Two dierent reaction sequences are proposed for the hydroamination reactions catalysed by late transition metal compounds: (i) Activation of the amine can be achieved by oxidative addition of the amine to the metal centre. This step requires two electrons of the metal and is established for electron rich metal centres

1387-1811/01/$ - see front matter 2001 Elsevier Science B.V. All rights reserved. PII: S 1 3 8 7 - 1 8 1 1 ( 0 1 ) 0 0 3 4 3 - 2

286

J. Penzien et al. / Microporous and Mesoporous Materials 48 (2001) 285291

such as Rh(I). If coupled with the insertion of a CC unsaturated group into the metalnitrogen bond and the reductive elimination of the product, catalytic alkene/alkyne hydroamination can be realised. Evidence for this route comes from the Rh catalysed reaction of styrene with secondary amines, where the hydroamination product baminoethylbenzene and the oxidative amination product b-aminoethenylbenzene are simultaneously formed [16]. This suggests a hydrido-b-aminoethyl-rhodium complex to be the common intermediate (Eq. (1)). However, dierent rate laws for the two products might also indicate that two distinct reaction sequences exist. (ii) Activation of the alkene or alkyne can be achieved by coordination to a late transition metal which renders the p-system susceptible to attack by amine nucleophiles. This leads to an intermediate 2-ammonioalkyl/alkenyl complex from which protolytic cleavage of the metalcarbon bond releases the hydroamination product (Eq. (2)).

This reaction sequence does not require changes in the oxidation state of the metal centre and is proposed for catalysts based on Zn2 for which oxidative addition of the amine seems unlikely. The intermediate 2-ammonioalkenyl complex was observed during the cyclisation of 6-aminohex1-yne with [Pd(Triphos)](CF3 SO3 )2 in signicant concentrations, indicating that the subsequent protolytic cleavage of the palladiumcarbon bond is the rate limiting step [17]. It should be emphasised that, at present, experimental evidence is insucient to distinguish unambiguously between the two reaction pathways. None of the above-described reactions have been explored over solid catalysts. Here we report on the rst successful utilisation of zinc ion-exchanged zeolites for an intramolecular hydroamination, i.e., the cyclisation of 6-aminohex-1-yne to 2-methyl-1,2-dehydropiperidine (Eq. (3)). Zinc was chosen as the catalytically active metal cation, as it has only one oxidation state available, allowing

J. Penzien et al. / Microporous and Mesoporous Materials 48 (2001) 285291

287

for a more straightforward mechanistic analysis. The role of Lewis (Zn2 cations) and Brnsted acid sites as catalytically active sites are discussed in the light of the analogy between homogeneous and heterogeneous catalysis. 2. Experimental Zinc ion-exchanged zeolites were prepared by slurrying the H-BEA (Sdchemie AG, T-4546, u MA039 H/99, 10 g) with 0.5 M zinc acetate in water (70 cm3 ) at 353 K and separated subsequently by centrifugation. The procedure was repeated four times to ensure complete ion-exchange. Finally the sample was calcined (5 K/min to 773 K, 4 h at 773 K). To achieve lower degrees of ion-exchange a less concentrated zinc acetate solution was used and the procedure was repeated less often. The zinc loadings (determined by atomic absorption spectroscopy) and the BET surface area of the materials are given in Table 1. Temperature programmed desorption (TPD) of ammonia and acetonitrile was performed in a home built apparatus and a modied Setaram microbalance. Before adsorption, the zeolites were activated by heating the sample in vacuum to 723 K (at 10 K/min) and kept at this temperature for 1 h. One mbar of CH3 CN and NH3 were adsorbed at the temperatures given in the caption of Fig. 2. Then the sample was outgassed in vacuum for 1.5
Table 1 Zinc loadings and BET surface area of the materials used in this study Material H-BEA Zn-BEA ZnO/SiO2 ZnO Zinc concentration (mmol Zn2 /g) 0.030.66 0.73 Sp. surface area (m2 /g) 536 460510 181 6.50

and 3 h, respectively, to remove weakly sorbed molecules and heated at 10 K/min to 923 and 973 K, respectively, monitoring the rate of desorption by mass spectroscopy. In a typical catalytic experiment, a mixture of 6-aminohex-1-yne (0.06 cm3 , 0.5 mmol), zinc exchanged BEA zeolite and toluene (15 cm3 ) was heated at reux (384 K) and samples taken for GC-analysis at regular intervals. The choice of solvent prevented leaching of Zn2 ions. For the experiments shown in Fig. 3, 11 mg Zn-BEA (0.66 mmol Zn2 /g), 1.9 mg ZnCF3 SO3 2 , 6.9 mg ZnO/ SiO2 , 8.3 mg ZnO and, for the experiments given in Fig. 4, 11 mg of the respective Zn-BEA catalyst were used. For the experiments involving an acid or base, CF3 SO3 H (9.2 mm3 , 0.1 mmol) or tetrabutylammonium acetate (6.3 mg, 0.02 mmol) was added to the reaction mixture of 6-aminohex-1-yne (0.06 cm3 , 0.5 mmol), Zn(CF3 SO3 )2 (1.9 mg, 5.2 lmol) and toluene (15 cm3 ), respectively. 3. Results and discussion 3.1. Catalyst characterisation Zn-BEA samples with loadings of between 0.03 and 0.66 mmol Zn2 per gram zeolite were prepared. For the parent H-BEA a Brnsted acid site concentration of 0.46 mmol H /g was determined gravimetrically by decomposition of the NH ex4 changed zeolite. Assuming the exchange of two protons by one Zn2 ion, the exchange degree in the Zn-BEA samples thus varied from 15% to 293%. Adsorption of NH3 on Zn-BEA with up to 0.31 mmol Zn2 /g showed, that for each Zn2 cation, slightly less than two NH3 molecules were adsorbed (Fig. 1). This suggests ion-exchange of the protons by ZnOH rather than Zn2 or ZnO Zn bridges as reported previously [18,19]. For samples containing more than 0.46 mmol Zn2 /g

288

J. Penzien et al. / Microporous and Mesoporous Materials 48 (2001) 285291

Fig. 1. Concentration of NH3 adsorbed on dierent zinc exchanged BEA zeolites.

additional zinc cannot be associated with cationexchange positions and is speculated to be present as ZnO. The sample with the highest zinc concentration was further characterised by TPD using CH3 CN

and NH3 as probe molecules. A comparison of the TPD proles obtained for the Zn-BEA zeolite with those of the parent H-BEA are compiled in Fig. 2. Adsorption was also performed at higher temperatures and the desorption curves are included in the gure. The contributions of the dierent desorption states were approximated by deconvolution using a linear combination of Gauss functions [20]. Acetonitrile was chosen as the probe molecule as it adsorbs preferentially on Lewis acid sites while it interacts only weakly with Brnsted acid sites [21,22]. On the parent H-BEA the rate of desorption of CH3 CN showed only one major peak at 430 K. The desorption maximum is attributed to physisorption or weak adsorption of CH3 CN molecules on Brnsted acid sites. On ZnBEA the rate of desorption of CH3 CN showed a maximum at 450 K and a shoulder at 650 K. Thus, the maximum on Zn-BEA was observed at approximately the same temperature than with HBEA. In addition, roughly the same number of molecules desorb from that state. This suggests interaction of CH3 CN with similar sites in both materials. The shoulder at 650 K can be attributed to new sorption sites created by zinc ions. The

Fig. 2. TPD experiments on H-BEA (left) and Zn-BEA (0.66 mmol Zn2 /g, right) using CH3 CN (top) and NH3 (bottom) as the probe molecules. The adsorption temperatures were 323, 423 K (CH3 CN on H-BEA), 313, 423, 493 K (CH3 CN on H-BEA), 313, 373, 493 K (NH3 on H-BEA) and 313, 373, 433, 613 K (NH3 on Zn-BEA).

J. Penzien et al. / Microporous and Mesoporous Materials 48 (2001) 285291

289

higher desorption temperature suggests a stronger interaction of CH3 CN, which is tentatively attributed to desorption from Zn2 sites. In contrast to acetonitrile, ammonia sorbs unspecically on Brnsted and Lewis acid sites [23]. The TPD of NH3 adsorbed on H-BEA showed a major peak at 430 K with a shoulder at 560 K suggesting two dierent adsorption sites for ammonia. On Zn-BEA a higher concentration of NH3 molecules was adsorbed than on H-BEA. Desorption peaks were observed at 420 K with shoulders at 520 and 780 K. The peak at 420 K had approximately the same area on both Zn-BEA and H-BEA. The increase in the intensity of the peaks at 520 and 780 K compared to H-BEA is attributed to sites associated with the presence of Zn2 cations. 3.2. Catalytic results The role of the cation-concentration in zeolite BEA for catalysis was explored for concentrations of up to 0.66 mmol/g of Zn2 . The rate of cyclisation increased from 1:7 103 to 6:0 102 mol (gcat h)1 . Up to a concentration of 0.31 mmol/g, the reaction rate increased linearly with the concentration of Zn2 but, at higher concentrations, the further increase had no eect (see Figs. 3 and 4). Additional zinc is, therefore, speculated to be present, in the zeolite pores, in the form of ZnO, which does not contribute to the catalytic activity.

Fig. 4. Initial rate for the cyclisation of 6-aminohex-1-yne with zinc exchanged BEA zeolites containing dierent amounts of Zn2 cations.

As the micropores of the zeolite catalysts may restrict transport of reactants and/or products, the intrinsic catalytic activity was also tested with homogeneous Zn-based catalysts. Surprisingly, the rate with the best homogeneous catalyst known for this reaction, i.e., Zn(CF3 SO3 )2 was lower than for Zn-BEA (5:5 102 mol (gcat h)1 ). Note that, under conditions employing the same analytic concentrations of Zn2 cations in the reaction mixture, the catalytic activity decreased in the order Zn-BEA > ZnCF3 SO3 2 > ZnO=SiO2 (Fig.5).

Fig. 3. Cyclisation of 6-aminohex-1-yne with Zn-BEA catalysts loaded with dierent amounts of Zn2 cations.

290

J. Penzien et al. / Microporous and Mesoporous Materials 48 (2001) 285291

Fig. 5. Cyclisation of 6-aminohex-1-yne to 2-methyl-1,2-dehydropiperidine with Zn-BEA, Zn(CF3 SO3 )2 , ZnO/SiO2 and ZnO.

At this point we would like to discuss the potential role of Zn2 as the catalytically active site. For the reaction sequence as depicted in Fig. 6 it is necessary that the reactant coordinates to the Lewis acid site. Despite the preference of the coordination at the amine end of the molecule, the relatively weak interactions of ammonia and amines with Zn2 ion-exchanged sites in zeolites makes it likely that the 6-aminohex-1-yne is also bound for a fraction of time at the CC triple bond. Such an interaction allows the attack of the amine nitrogen atom at the secondary carbon atom of the CC triple bond. The surface intermediate formed in this way requires a proton in two of the subse-

quent elementary steps, suggesting that Brnsted and Lewis acid sites may be necessary in an ideal catalyst. Note that the zeolites oer the possibility that both sites (Lewis acidic Zn2 and Brnsted acid sites) are present in close vicinity, while in the homogeneous catalyst only the Lewis acid sites exist. In order to test this hypothesis, CF3 SO3 H was added in 20-fold excess to the homogeneous catalyst. The presence of such a high amount of acid increased the activity of Zn(CF3 SO3 )2 to such an extent that the specic activities of zinc ions in both catalyst systems were approximately identical (1:9 102 mol (molZn2 h)1 ). It should be noted that, in contrast, the activity of Zn2 based catalysts was lowered by the presence of bases such as tetrabutylammonium acetate to 5:0 103 mol (gcat h)1 . This strongly suggests that either there is not only one rate-determining step involving Zn2 , or the presence of protons in the reaction mixture shifts an equilibrium in the reaction sequence. In the absence of denite experimental evidence, we would like only to point out that two elementary reaction steps, i.e., the cleavage of the metalcarbon bond in the intermediate 2-ammonioalkenyl zinc complex, or the isomerisation of 2-methylenepiperidine to the nal product 2-methyl-1,2-dehydropiperidine, could be involved. In addition to an eect on these two steps, the presence of pro-

Fig. 6. Mechanism proposed for the cyclisation of 6-aminohex-1-yne.

J. Penzien et al. / Microporous and Mesoporous Materials 48 (2001) 285291

291

tons in the reaction mixture reduces the possibility that the protonated amine group interacts with the Zn2 cations. Therefore, it increases the probability that the acetylenic group can interact with the catalyst.

the catalytic action and this appears to be the reason for the excellent catalytic properties of Zn-BEA. Acknowledgements Support from DSM, the ``Stiftung StipendienFonds des Verbandes der Chemischen Industrie e.V.'' and ``Dr.-Ing. Leonhard-Lorenz-Stiftung'' is gratefully acknowledged. Xaver Hecht and Martin Neukamm are thanked for the experimental support. References
[1] D.M. Roundhill, Catal. Today 92 (1997) 1. [2] T.E. Mller, M. Beller, Chem. Rev. 98 (1998) 675. u [3] K. Tanabe, W.F. Hoelderich, Appl. Catal. A Gen. 181 (1999) 399. [4] W.F. Hoelderich, G. Heitmann, Catal. Today 38 (1997) 227. [5] M. Tokunaga, M. Eckert, Y. Wakatsuki, Angew. Chem. Int. Ed. 21 (1999) 3222. [6] R. Dorta, P. Egli, F. Zrcher, A. Togni, J. Am. Chem. Soc. u 119 (1997) 10857. [7] M. Beller, H. Trauthwein, M. Eichberger, C. Breindl, T.E. Mller, Eur. J. Inorg. Chem. 7 (1999) 1121. u [8] T.E. M ller, M. Grosche, E. Herdtweck, A.-K. Pleier, u E. Walter, Y.-K. Yan, Organometallics 19 (2000) 170. [9] W. Reppe, Liebigs Ann. Chem. 601 (1956) 81. [10] K. Eller, R. Kummer, P. Stops, BASF AG, WO 97/21661A1, 1997. [11] J.F. Knifton, P.E. Dai, B.L. Benac, Texaco Chem. Co, CA 2092964-A, 1994. [12] H.J.R. De Boer, P.D. Savage, Shell Int. Res. Maat. B.V., EP 578323-A2, 1994. [13] Mitsui Toatsu Chem. Inc., JP 4139156-A, 1992. [14] D.D. Dixon, W.F. Burgoyne, Air Prod. Chem. Inc., EP 334332-A1, 1989. [15] M. Deeba, Air Prod. Chem. Inc., US 4536602, 1985. [16] M. Beller, H. Trauthwein, M. Eichberger, C. Breindl, J. Herwig, T.E. M ller, O.R. Thiel, Chem. Eur. J. 5 (1999) 1306. u [17] T.E. Mller, M. Berger, M. Grosche, E. Herdtweck, F.P. u Schmidtchen, Organometallics, submitted for publication. [18] H. Berndt, G. Lietz, J. V lter, Appl. Catal. A Gen. 146 o (1996) 365. [19] V.B. Kazansky, V. Yu, A.I. Serikh, R.A. van Santen, B.G. Anderson, Catal. Lett. 66 (2000) 39. [20] C. Costa, J.M. Lopes, F. Lemos, F. Ramoa Ribeiro, J. Mol. Catal. A 144 (1999) 221. [21] L.A.M.M. Barbosa, R.A. van Santen, Catal. Lett. 63 (1999) 97. [22] J.F. Haw, M.B. Hall, A.E. Alvarado-Swaisgood, E.J. Munson, Z. Lin, L.W. Beck, T. Howard, J. Am. Chem. Soc. 116 (1994) 7308. [23] R.J. Gorte, Catal. Lett. 62 (1999) 1.

4. Conclusions The intramolecular hydroamination, i.e., the cyclisation of 6-amino-1-hexyne to 2-methyl-1,2dehydropiperidine is eectively catalysed by Zn2 ion-exchanged zeolites BEA. The intrinsic activity of these heterogeneous catalysts is superior to catalysts based on homogeneous zinc complexes such as Zn(CF3 SO3 )2 . This is concluded to be caused by the simultaneous presence of Brnsted and Lewis acid sites in Zn-BEA. Brnsted acid sites accelerate the reaction and addition of a 20fold molar excess of a Brnsted acid (CF3 SO3 H) to Zn(CF3 SO3 )2 led to comparable rates for both catalysts. ZnO/SiO2 and pure ZnO show signicantly lower catalytic activity. This is attributed, on the one hand, to the lower Lewis acid strength decreasing in the sequence Zn-BEA > ZnO= SiO2 > ZnO and, on the other hand, to the decreasing Brnsted acid strength and acid site concentration in this sequence. Mechanistically, the reacting 6-aminohex-1-yne is thought to coordinate via its acetylenic bond onto the Zn2 at ion-exchange sites of Zn-BEA. The coordinated acetylenic bond is then attacked by the amine. The resulting formal zwitter-ion rearranges and forms an intermediate 2-methylene-piperidine which isomerises to the nal product 2-methyl-1,2-dehydropiperidine. The specic role of the Brnsted acid sites in the catalytic cycle is not fully understood at present. Three possibilities exist (see Fig. 6), i.e., (A) the promotion of the formal 1,3-proton shift which leads to cleavage of the metalcarbon bond in the intermediate 2-ammonioalkenyl zinc complex, (B) the acceleration of the isomerisation from 2-methylene-piperidine to 2-methyl-1,2-dehydropiperidine and (C) an enhancement of the probability of bonding the acetylenic group by reversible protonation of the amine group. In all three cases the direct vicinity of Brnsted and Lewis acid sites seems benecial for

You might also like