Open Access ArticleBeth L. Moorea, Annhelen Lua, Deborah A. Longbottomb and Rachel K. O'Reilly*a
aDepartment of Chemistry, University of Warwick, Gibbet Hill Road, Coventry, CV4 7AL, UK. E-mail: Rachel.OReilly@Warwick.ac.uk
bDepartment of Chemistry, University of Cambridge, Lensfield Road, Cambridge, CB2 1EW, UK
First published on 6th February 2013
The MacMillan catalyst is an established organocatalyst capable of catalyzing a variety of organic reactions. Through the synthesis of a novel monomer containing the MacMillan catalytic functionality, a variety of copolymers have been synthesized with the comonomer, diethylene glycol methyl ether methacrylate (DEGMA). Reversible addition–fragmentation chain transfer (RAFT) polymerization was used for the synthesis of these functional polymers with good control over molecular weight, catalyst incorporation and polydispersity. These polymers showed lower critical solution temperature (LCST) behaviour where the cloud point was found to be dependent upon the degree of catalyst incorporation and catalyst loading was also found to have an effect on the Tg of the copolymers. The catalytic activity of the functional copolymers is demonstrated by the Diels–Alder reaction between cyclopentadiene and trans-hexen-1-al and shows enantioselectivity close to those previously reported by MacMillan. The polymers can be reused in multiple Diels–Alder reactions via a pseudo continuous process, maintaining high conversion and enantioselectivity throughout the cycles.
Recent advances in polymerization techniques now allow the polymerization of functional monomers containing catalytically active groups with high levels of control, producing a myriad of polymer bound catalysts, which can often be recovered and recycled to some degree. For example, our group has reported the successful anchoring of L-proline and DMAP catalysts, incorporating them into polymeric frameworks which self-assemble into recyclable and reusable nanostructures.16–18 Huerta et al. have immobilized L-proline into a polymer able to fold into specific conformations which places the catalytic units in a hydrophobic domain.19 Ge et al. used a similar polymeric micellar system to support imidazole, which was demonstrated to efficiently catalyze ester hydrolysis reactions.20 Use of the MacMillan catalyst was first reported in 2000: it is a novel imine–enamine catalyst capable of accelerating the rate of the Diels–Alder reaction with good control over enantioselectivity and in the majority of cases, diastereoselectivity.21 The catalyst efficiently lowers the LUMO of the dienophile providing a substantial increase in rate of reaction, whilst its steric bulk controls the enantioselectivity. This work was then further extended in 2002 to a range of α,β-unsaturated ketone Diels–Alder substrates22 and beyond this, it has found use in a broad range of processes, including the Mukaiyama–Michael23 Friedel–Crafts alkylation,24 cascade reactions,25 transfer hydrogenation and even the first enantioselective organocatalytic hydride reduction.26 The range of reactions catalyzed by the MacMillan catalyst make it a very powerful, practical and truly versatile organocatalyst.
There have been several attempts to support the MacMillan catalyst, aiming to maintain its high selectivity but still allow for recovery and reuse (outlined in the review by Hansen and Kristensen).7 The first of these was reported by Cozzi et al., where the catalyst was immobilized on PEG supports. In most cases, comparable selectivities to those first reported by MacMillan were achieved, although a marginal loss of selectivity on recycling was reported.27 Other supports include a JandaJel™ system,28 mesoceullar foams29 and sulfonated polystyrene.30 One system, developed by Pericas and co-workers in 2012, utilized superparamagnetic Fe3O4 nanoparticles as the support where the catalytic functionality was introduced through a click reaction.31 The supported catalyst successfully catalyzed the Friedel–Crafts alkylation reaction and was recycled using its magnetic properties. Unchanged enantioselectivities were reported (∼90%) over six cycles. Hansen and co-workers immobilized the MacMillan catalyst in 2010 using a different approach: a polymerizable monomer was synthesized and copolymerized with poly(ethylene glycol) in a suspension polymerization.32 However, a decrease in enantioselectivity with recycling was reported (81% to 51% ee after four cycles). Their work bears the closest resemblance to the work discussed in this paper, as it moves away from the anchoring strategy using pre-made polymers (akin to solid-phase peptide synthesis) to instead use catalyst functionalized monomers to form a functional polymer scaffold. The advantages of this bottom-up strategy include greater control over catalyst loading compared to post-modification schemes. In addition, the catalyst position in the polymer chain can be controlled by employing sequential addition techniques, which may also be used to synthesize block copolymers. The polarity and hence solubility of the polymer is also easily tuned by simply changing the ratio of functionalized monomer to comonomer.
Recent advances in controlled radical polymerization (CRP) techniques have allowed for more complex and functional polymer architectures to be readily synthesized. This is achieved through the polymerization of monomers containing functional groups which were previously difficult to polymerize using conventional and sensitive living polymerization techniques. Reversible addition–fragmentation chain transfer (RAFT) polymerization has been used in this work, as this technique has been shown to be appropriate for a range of functional monomers and conditions.33–37 We herein report the synthesis of a novel MacMillan functionalized monomer and its subsequent copolymerization yielding well-defined copolymers. Efficient catalysis of a model Diels–Alder reaction between cyclopentadiene and trans-hexen-1-al is then demonstrated, followed by reuse of the copolymers in several catalytic cycles via a pseudo continuous process.
:
propan-2-ol, 90
:
10 using a Chiracel OD-H column 250 mm × 4.6 mm × 5 μm, with guard column (5 μm)) showed the following retention times: S-enantiomer, tR = 7.9 min and R-enantiomer, tR = 7.2 min.
CH2), 2.85 (3H, s, N(CH3)), 2.94 (1H, dd, J = 10.5 Hz, 15 Hz, Ar–CHH–CH–), 3.48 (1H, dd, J = 3.6 Hz, 15 Hz, Ar–CHH–CH–), 4.55 (1H, dd, J = 3.6 Hz, 10.5 Hz, Ar–CHH–CH–), 5.76 (1H, s, C(CH3)
CHH–), 6.25 (1H, s, C(CH3)
CHH–), 7.09 (2H, d, J = 8.4 Hz, Ar), 7.39 (2H, d, J = 8.4 Hz, Ar). 13C NMR (75 MHz, CD3OD): δ 18.5 (CH(CH3)
CH2), 22.2 (C(CH3)2), 24.5 (C(CH3)2), 25.7 (N(CH3)), 34.7 (Ar–CH2), 59.7 (Ar–CH2–CH), 78.9 (C(CH3)2), 123.5 (Ar), 128.1 (CH(CH3)
CH2), 131.4 (Ar), 134.3 (Ar), 137.2 (CH(CH3)
CH2), 151.9 (C(O)–CH(CH3)
CH2), 168.1 (N(CH3)–C(O)), (ESI Fig. S1 and S2†). HR ESI-MS: found 303.1698 m/z [M + H]+ expected 303.1709.
:
propan-2-ol, 90
:
10) and showed only the presence of the S-enantiomer, tR = 7.9 min (R-enantiomer, tR = 7.2 min, ESI Fig. S5†).
:
H2O (95
:
5 v/v%) was used a solvent (to remove acetal side-products). The aliquot (∼0.1 mL) was stirred in H2O
:
TFA
:
CHCl3 (1
:
1
:
2) (∼4 mL) before being neutralized with NaHCO3 (∼2 mL) and then extracted into Et2O (2 × 5 mL). Conversion was determined by 1H NMR spectroscopy and ee% measured by GC, injection temperature 250 °C, column temperature 80 °C, ramp to 160 °C at 4.5 °C min−1, exo isomers tR = 12.8 and 13.2 min, endo isomers tR = 12.9 and 13.4 min. 1H NMR (300 MHz, CDCl3): δ 9.33 (1H, d, J = 3.0 Hz, C(O)H exo), 9.45 (1H, d, J = 8.1 Hz, C(O)H starting material), 9.75 (1H, d, J = 3.0 Hz, C(O)H endo). Data from M1 in H2O: conversion 94% (1H NMR spectroscopy), exo
:
endo ratio 1.00
:
1.05 (1H NMR spectroscopy), exo 73% ee, endo 88% ee (GC analysis), (ESI Fig. S7†).
:
H2O (95
:
5 v/v%, 1 mL, 0.09 M). TFA (0.130 mL, 1 eq.) and trans-hexen-1-al (0.200 mL, 1 eq.) were added to the polymer solution and left to stir for 5 min before cyclopentadiene (0.300 mL, 2 eq.) was added. After 4 hours an aliquot was taken for analysis and the reaction mixture was washed with Et2O (2 × 10 mL) and CHCl3 (2 × 10 mL) removing the organic starting materials and products. The remaining aqueous layer was then diluted with DMSO. The water–DMSO solution containing the polymer was then dialyzed extensively against deionized water (MWCO = 3500 Da) and freeze-dried to give a white solid. The recovered polymer was weighed and reused in a second cycle (ESI Table S5†).
:
H2O (95
:
5 v/v%, 1 mL, 0.09 M). TFA (0.13 mL, 1 eq.) and trans-hexen-1-al (0.2 mL, 1 eq.) were added to the polymer solution and left to stir for 5 min before cyclopentadiene (0.3 mL, 2 eq.) was added. The reaction being neutralized with NaHCO3 (2 mL) and extracted into Et2O (2 × 5 mL). The remaining polymer solution was washed with hexane, extracting the starting materials and products leaving the polymer in the acidic CH3OH
:
H2O solution. To this solution, more reagents (trans-hexen-1-al and cyclopentadiene) were added and the catalysis/reuse process repeated (ESI Table S6†).![]() | ||
| Scheme 1 Synthesis of S-MacMillan functionalized monomer (M1) from S-tyrosine methyl ester. | ||
![]() | ||
| Scheme 2 A representative RAFT polymerization scheme of M1 and DEGMA. | ||
| Polymer | Feed ratio (DEGMA : M1) | Mn,th (kDa) | Mna (kDa) | Mw/Mna | Catalyst incorporationb (%) |
|---|---|---|---|---|---|
| a Measured by THF SEC against PMMA standards.b Measured by 1H NMR spectroscopy. | |||||
| P1 | 92 : 8 | 17.0 | 11.8 | 1.35 | 6 |
| P2 | 76 : 24 | 21.7 | 11.6 | 1.39 | 26 |
| P3 | 69 : 31 | 20.7 | 11.8 | 1.33 | 33 |
| P4 | 58 : 42 | 23.3 | 12.0 | 1.39 | 38 |
| P5 | 37 : 63 | 22.7 | 11.5 | 1.47 | 57 |
| P6 | 13 : 87 | 26.8 | 11.4 | 1.46 | 73 |
| P7 | 0 : 100 | 27.8 | 5.8 | 1.31 | 100 |
To confirm that no racemization had taken place during the polymerization process, the standard non-polymerizable S-MacMillan catalyst was synthesized and subjected to the same polymerization conditions as M1. Following recovery, analysis by chiral-HPLC revealed only the S-enantiomer, suggesting no racemization occurred during polymerization and hence, it has been inferred that only the S-enantiomer of M1 is incorporated into the polymer (ESI Fig. S5†, R-enantiomer tR = 7.2 min and S-enantiomer tR = 7.9 min).
To determine the distribution of catalytic functionality along the polymer chain, monomer reactivity ratios were investigated. A number of polymerizations were carried out using a variety of monomer ratios, i.e. M1
:
DEGMA, 90
:
10, 70
:
30, 50
:
50, 30
:
70, 10
:
90. Conversions of both monomers were determined by 1H NMR spectroscopy with conversions kept low (between 5 and 15%). The mol fractions of the two monomers in the initial feed and in the final polymer were used in the Contour program, developed by van Herk;42 and the reactivity ratios were determined to be M1 = 0.892 and DEGMA = 0.575 (ESI Fig. S6†). These values suggest that whilst it is not a completely random system, the two monomers are well distributed throughout the resultant polymer.
The comonomer DEGMA is known to exhibit an LCST and thus the LCST of some of copolymers were examined (ESI Table S1†). The cloud points were found to be dependent on polymer concentration and were therefore examined at multiple concentrations (0.5, 1.0, 2.0 and 5.0 mg mL−1). The cloud points for the PDEGMA homopolymer (Mw = 16.7 kDa) were found to vary from 27.4 °C for 0.5 mg mL−1 to 24.4 °C for 5.0 mg mL−1. For P1 (6% loading), higher LCST cloud points were uniformly observed compared with DEGMA: from 37.7 °C for 0.5 mg mL−1 through to 26.5 °C for 5.0 mg mL−1, further increasing for P2 (26% loading): 58.8 °C for 0.5 mg mL−1 through to 42.4 °C for 5.0 mg mL−1. At lower concentrations of P2, the increase in absorption was found to be significantly smaller compared to the other polymers suggesting that less polymer is precipitated, potentially indicating that the polymer is losing its LCST behaviour. Analysis on polymers with higher incorporation proved inconclusive as the temperature range became too high. A possible explanation for the increase in LCST cloud point temperature is the ability of M1 to hydrogen-bond to itself as well as to water, and therefore increase the temperature at which entropic loss (due to formation of hydrogen-bonded structures) outweighs the enthalpic gain (from the formed bonds). Hence, as the incorporation of M1 increases, the LCST cloud point increases.
In 1998, Bergbreiter and co-workers reported the use of poly(N-isopropyl acrylamide) (PNIPAM), another polymer known to exhibit an LCST behaviour, as a recoverable polymeric system.43 The polymer was recovered and recycled by heating above its LCST, resulting in precipitation. The utility of LCST polymers for catalyst recovery has also been reported by our group in 2012, also using PNIPAM to efficiently recover a DMAP functionalized polymer.16 However, whilst some of the polymers reported here exhibit this behaviour, the temperature at which the polymer precipitates is strongly dependent on the degree of catalyst incorporation and polymer concentration. As the LCST gets higher (>40 °C) recovery via this method is less efficient and it is therefore not a general method for recycling in this case. Whilst P1 has an LCST cloud point that could potentially be utilized, due to small scale of the reactions this was difficult to investigate.
:
endo ratio of 1.00
:
1.00; an exo ee% 84 and an endo ee% of 93.![]() | ||
| Scheme 3 Model Diels–Alder reaction, where R = C3H7 and where there are four possible products: the endo and exo products of both enantiomers. The favoured enantiomer is likely to be the R-endo-enantiomer if the reaction follows the same pattern as suggested by MacMillan et al.,21 arising from attack from the opposite side to the phenyl group in the catalyst. | ||
:
H2O 95
:
5 v/v% vs. H2O)
| Catalyst | Solvent | Conversiona (%) | exo : endoa | exo eeb% | endo eeb% |
|---|---|---|---|---|---|
| a Determined by 1H NMR spectroscopy.b Determined by chiral GC analysis. | |||||
| M1 | CH3OH : H2O | 84 | 1.00 : 1.04 | 79 | 88 |
| H2O | 94 | 1.00 : 1.05 | 73 | 88 | |
| P1 | CH3OH : H2O | 87 | 1.00 : 1.00 | 79 | 83 |
| H2O | 75 | 1.00 : 1.08 | 69 | 84 | |
| P2 | CH3OH : H2O | 96 | 1.00 : 1.01 | 80 | 88 |
| H2O | 70 | 1.00 : 1.08 | 73 | 81 | |
| P3 | CH3OH : H2O | 100 | 1.00 : 1.12 | 81 | 86 |
| H2O | 86 | 1.00 : 1.11 | 74 | 85 | |
| P4 | CH3OH : H2O | 90 | 1.00 : 0.62 | 75 | 85 |
| H2O | 85 | 1.00 : 1.14 | 74 | 85 | |
| P5 | CH3OH : H2O | 94 | 1.00 : 0.77 | 78 | 74 |
| H2O | 95 | 1.00 : 1.07 | 74 | 88 | |
| P6 | CH3OH : H2O | 84 | 1.00 : 1.08 | 76 | 83 |
| H2O | 91 | 1.00 : 1.07 | 74 | 82 | |
| P7 | CH3OH : H2O | 92 | 1.00 : 1.08 | 77 | 84 |
| H2O | 86 | 1.00 : 1.13 | 76 | 85 | |
The results were similar for both monomers and polymers examined: conversions and the exo
:
endo ratios were similar across all the reactions (1.00
:
0.98–1.14) and the ees comparable but uniformly slightly lower than those reported in the literature for both endo and exo products. Moreover, changing the solvent from CH3OH
:
H2O (95
:
5 v/v%) to H2O appears to have little effect on the reaction. These results were interesting as the importance of catalyst site isolation has been previously demonstrated and it was anticipated that varying catalyst loading would have an effect on the catalytic activity.44 In the absence of any catalyst the reaction is significantly slower reaching 9.9% (H2O) and 0.99% (MeOH
:
H2O 95
:
5 v/v%) conversions after 4 hours.
Therefore, the reaction kinetics of two polymers were investigated to determine whether catalytic loading affected the rate of reaction. This was carried out using P2 (26% loading) and P6 (73% loading) (ESI Table S2†). Interestingly, this still did not lead to a tangible difference: the kinetics for both polymers were found to be very similar (25 min P6 = 66%, P2 = 53% and 240 min P6 = 96% and P2 = 93%). The exo
:
endo ratio (1.00
:
1.06–1.19) and enantioselectivities (74–80% ee for the major product) are also comparable for both polymers for the duration of the reaction. The comparable results for the two polymers suggest that in this case, catalyst site isolation does not affect polymer catalytic activity and that polymers with low catalyst loading (hence less expensive catalyst monomer is required) are still extremely efficient.
These initial catalysis results, carried out at room temperature, indicate that activity and selectivity are also unaffected by the Tg of the polymers. In order to confirm this, additional experiments were carried out at various temperatures (4–60 °C, ESI Table S3†). M1, P2 (Tg 36.1 °C) and P7 (Tg 140.0 °C) were used to catalyze the model Diels–Alder reaction (Scheme 3) and showed no significant differences: each displayed an increase in conversion and decrease in selectivity with increasing reaction temperature, confirming that the glass transition temperature does not have an effect.
Despite being a good handle for post-polymerization modifications,45 the RAFT end group present on the polymers after polymerization could potentially interfere with the catalysis reaction. In order to investigate the potential role of the end group in the Diels–Alder reaction it was removed from P6 using a radical induced end-group removal chemistry with 1-EPHP as the proton donor (ESI Fig. S8†). The activity of the polymer pre- and post-end group removal is comparable (91% vs. 91%), as is the enantioselectivity (endo 82 vs. 75% ee, ESI Table S4†). Therefore the presence of the end group, as well as the chemistries employed to remove it, have limited effect on the selectivity of the reaction.
The versatility of our polymer-supported MacMillan system was demonstrated by the catalytic efficiency of P2 in the Diels–Alder reaction of a range of substrates. Cyclopentadiene was reacted with a range of aldehydes (Table 5) structurally similar to the one presented in Scheme 3 but with a different R group. P2 catalyzed the reactions with great efficiency achieving conversions between 70 and 100% and good enantioselectivities (endo 72–89%), further demonstrating the catalytic ability of the polymers in a range of Diels–Alder reactions.
As previously discussed, the use of PDEGMA as an LCST polymer has proven to be quite difficult as the LCST was found to be dependent on the degree of incorporation of M1, as well as the polymer concentration in solution (ESI Table S1†), thus, rendering this recovery route only applicable to certain polymers at certain concentrations.
Therefore, initially attempts were made to recover these MacMillan supporting polymers by other established polymer recovery techniques (i.e. via dialysis and freeze-drying). However, these resulted in low recovery yields (from mechanical loss), as well as a notable loss of enantioselectivity in subsequent reactions (ESI Table S5†). In light of this, a new pseudo continuous process was developed for polymer recovery and reuse: due to significant differences in solubility between the polymer catalyst and the reagents/products, starting materials and products may simply be extracted into hexane. The polymer, which is insoluble in hexane, remains in the reaction solvent, and can then be reused multiple times, achieving good results for each cycle (Table 6).
:
H2O (95
:
5 v/v%) in multiple cycles via a pseudo continuous process
Conversion for all cycles has remained high, staying above 70% and pleasingly, enantioselectivities were also maintained (endo 79–88%). Crucially, it was found that if the acid (TFA) was not added in each new cycle, high conversions were maintained. However, if acid was added together with the new reagents, a significant drop in enantioselectivity was observed (ESI Table S6†). Therefore, for efficient recycling, it can be assumed that both acid and polymer remain in the CH3OH
:
H2O layer and extra acid should not be added during subsequent reuse reactions.
Footnote |
| † Electronic supplementary information (ESI) available: Further characterization of the synthesized monomer and polymers. HPLC chromatogram of the catalyst after polymerization. A typical GC spectrum for the chiral analysis of the Diels–Alder products. LCST cloud point data for polymers at different concentrations. Catalysis data for P2 and P7 at different temperatures; P2 and P6 kinetic data; P6 before and after end group removal and recycling by freeze-drying. Analysis of the recovered polymer after pseudo continuous experiments. See DOI: 10.1039/c3py21125h |
| This journal is © The Royal Society of Chemistry 2013 |