Fangyu Hu,
Shoulei Xie,
Liming Jiang* and
Zhiquan Shen
MOE Key Laboratory of Macromolecular Synthesis and Functionalization, Department of Polymer Science and Engineering, Zhejiang University, Hangzhou 310027, China. E-mail: cejlm@zju.edu.cn
First published on 29th October 2014
The cationic ring-opening polymerization (CROP) of substituted 2-oxazolines using rare-earth metal triflates (RE(OTf)3) as initiator was investigated for the first time. In this work, we examined the polymerization characteristics of 2-ethyl-2-oxazoline (EtOx) initiated by Sc(OTf)3 under conventional thermal heating and microwave irradiation, and compared the respective outcomes with those obtained with the most frequently used initiator methyl tosylate (MeOTs). The results indicated that Sc(OTf)3 exhibits a higher catalytic efficiency to the EtOx polymerization than MeOTs under identical conditions. The controlled/living nature of the Sc(OTf)3-catalyzed CROP was confirmed by its linear first-order kinetics and the narrow molecular weight distribution of the resultant polymers as well as the block copolymerization of EtOx and 2-phenyl-2-oxazoline (PhOx). Based on in situ NMR spectroscopic studies and SEC analysis of PEtOx samples obtained from the control termination experiments, a possible initiating/propagating mechanism has been proposed for the living cationic ring-opening polymerization. Morever, this rare-earth catalytic system can also be applied to the ring-opening polymerization of some sterically hindered or aryl-substituted 2-oxazolines.
To date, various initiators including Lewis acids, such as boron trifluoride, and alkyl esters such as tosylates, triflates and halides have been reported for the cationic ring-opening polymerization of 2-oxazolines.5–7 Among them, tosylates and triflates could be particularly effective in combination with microwave irradiation, which enables the polymerization in minutes to a few hours.8,9 Although such a microwave-assisted polymerization was achieved in a laboratory equipment, the usefulness of this synthesis procedure and its manipulation still remain problematic for a scale up due to the intrinsic limitations of microwave techniques. As for tosylate and triflate, the yet most extensively used initiators, their high toxicity is a relevant key issue that must be taken in consideration in practical applications.10–13 Therefore, it is desirable to explore new catalytic systems to alleviate aforementioned problems and permit the CROP of 2-oxazolines to be performed on a normal apparatus. Also, we feel that more fundamental studies on the design and synthesis of polyoxazoline-based materials with new functional properties could be beneficial to the field.14–17
On the other hand, rare-earth metal triflates (RE(OTf)3) as a hard Lewis acid have been applied in the polymer synthesis. For example, Okamoto and his co-workers18–21 established a general method applicable to some polar monomers such as (meth)acrylamides and methacrylates wherein the added RE(OTf)3 catalytically affect the polymerization stereochemistry to isotactic-selective manner, producing corresponding highly stereoregular polymers. With the same protocol, we recently synthesized isotactic-rich polyacrylamide derivatives bearing a chiral oxazoline moiety in the side chain, which can serve as a chemosensor for the enantioselective recognition of 1,1′-bi-2-naphthol.22,23 More recently, RE(OTf)3 has proven to be an efficient catalyst for the ring-opening polymerization of lactones,24,25 tetrahydrofuran,26 and amino acid N-carboxyanhydrides.27 In view of this background, it was thought that this kind of rare-earth salts should have potential uses as initiator for the CROP of 2-oxazolines.
The present work aims at examining (i) the catalytic performance of rare-earth metal triflates for the cationic ring-opening polymerization of 2-oxazolines, and (ii) its monomer application scope. For this purpose, we investigated the Sc(OTf)3-catalyzed polymerizations of EtOx under the conventionally thermal heating and microwave irradiation conditions in comparison with those by using methyl tosylate as initiator. In addition, the influence of the temperature on polymerization kinetics was studied. Based on the observations from in situ NMR analyses and control termination experiments, a possible polymerization mechanism was proposed for the rare-earth catalysis. Finally, we briefly screened the polymerization behavior of other substituted 2-oxazolines in the presence of Sc(OTf)3. As depicted in Scheme 1, these monomers can be divided into two categories, namely, the sterically hindered 2-butyl-2-oxazolines bearing an alkyl group at the 4- or 5-position, and 2-phenyl-2-oxazoline derivatives where t-butyloxycarbonyl (Boc)-protected L-proline moiety was bound to the benzene ring via amide or ester linkage. Despite some Lewis acids had been utilized as an initiator for the polymerization of 2-oxazolines as early as 1960′s,28 we note that Lewis acidic rare-earth triflates have not yet been reported, to our knowledge.
:
1), giving ProPhOx-1 (1.55 g, 86%) as a white crystal. M.p. = 123.2–124.2 °C. 1H NMR (400 MHz; CDCl3): δ = 1.47 (s, OC(CH3)3, 9H), 1.91–2.11 (m, N(CH2)2CH2, 2H), 2.23–2.59 (m, NCH2CH2CH2, 2H), 3.41–3.65 (m, NCH2(CH2)2, 2H), 4.02 (t, NCH2CH2O, 2H), 4.35 (t, NCH2CH2O, 2H), 4.51–4.60 (m, OCCHN, 1H), 7.08–7.19 (m, Ph, 1H), 7.27–7.32 (m, Ph, 1H), 7.45–7.52 (m, Ph, 1H), 7.91–7.95 (m, Ph, 1H); 13C NMR (125 MHz; DMSO-d6): δ = 22.9, 23.9, 28.0, 29.3, 46.3, 54.5, 58.8, 66.7, 79.0, 120.9, 123.3, 126.2, 130.3, 132.5, 148.9, 153.7, 160.2, 170.4. MS (ESI+): m/z (%) = 383.0 (39.6) [M + Na]+, 743.1 (100) [2M + Na]+.
:
1) as an eluant, and the intermediate 2a was obtained as a white crystal (1.84 g, 56.7%). M.p. = 53.4–54.3 °C (lit.31 54.5–55.5 °C). The amidation of 2a was conducted by the same procedure as the esterification of 1b, and the desired product ProPhOx-2 was obtained as a white crystal (89.5%). M.p. = 130.3–131.5 °C. 1H NMR (400 MHz; CDCl3), δ = 1.32 (s, OC(CH3)3, 9H), 1.91–1.95 (m, N(CH2)2CH2, 2H), 2.13–2.40 (m, NCH2CH2CH2, 2H), 3.40–3.72 (m, NCH2(CH2)2, 2H), 4.02–4.49 (m, NCH2CH2O, NCH2CH2O and OCCHN, 5H), 7.09–7.13 (m, Ph, 1H), 7.46–7.48 (m, Ph, 1H), 7.84–7.87 (m, Ph, 1H), 8.80–8.83 (m, Ph, 1H), 12.79 (s, NHCO, 1H); 13C NMR (125 MHz; DMSO-d6): δ = 23.2, 23.9, 27.7, 28.0, 30.2, 31.0, 46.4, 54.3, 61.9, 66.3, 78.8, 112.8, 118.5, 122.4, 128.9, 132.5, 139.0, 152.9, 153.9, 163.4, 172.1. MS (ESI+): m/z (%) = 382.1 (93.8) [M + Na]+, 741.1 (100) [2M + Na]+.
:
1) to give 3a as yellowish solid (3.67 g, 45.4%). M.p. = 159.5–160.5 °C (lit.32 160–161 °C). ProPhOx-3 was prepared by a similar way to that of ProPhOx-2 except for the use of EDC as the dehydrant and THF as solvent (53.7%, white crystal). 1H NMR (400 MHz; CDCl3): δ = 1.44 (s, OC(CH3)3, 9H), 1.84–1.98 (m, N(CH2)2CH2 and NCH2CH2CH2, 4H), 3.22–3.61 (m, NCH2(CH2)2, 2H), 4.04 (t, NCH2CH2O, 2H), 4.40–4.44 (m, NCH2CH2O and OCCHN, 3H), 7.56–7.59 (m, Ph, 2H), 7.88–7.90 (m, Ph, 2H), 9.77 (br, NHCO, 1H); 13C NMR (125 MHz; DMSO-d6): δ = 171.6, 162.6, 153.2, 141.7, 128.5, 122.1, 118.5, 78.5, 67.2, 60.2, 54.3, 46.6, 33.3, 30.9, 30.1, 27.9, 23.6. MS (ESI+): m/z (%) = 360.1 (100) [M + H]+, 719.0 (62.6) [2M + H]+, 741.1 (23.8) [2M + Na]+.2-Butyl-2-oxazoline derivatives, including 4-ethyl-2-butyl-2-oxazoline (4-EtBuOx), 2-butyl-4-methyl-2-oxazoline (4-MeBuOx), and 2-butyl-5-methyl-2-oxazoline (5-MeBuOx) were synthesized according to a general procedure described in the literature.33
For comparison, the polymerization of EtOx with microwave assistance was carried out in a CEM Discover SP microwave synthesizer. In this case, a stock solution of the monomer and Sc(OTf)3 or MeOTs as initiator with various monomer-to-initiator ratios was prepared in acetonitrile. The microwave vial was charged with 1–2 mL of this stock solution and heated in the microwave to a certain temperature. After a given time, the reaction mixture was cooled with compressed air and quenched by adding deionized water. The subsequent treatment of the polymerization product is the same as those described above.
To evaluate the catalytic performance of rare-earth triflates for CROP of 2-oxazolines, we first investigated the solution polymerization of EtOx under different conditions. The polymerizations with RE(OTf)3 (RE = Sc, Y, La, Dy or Lu) were found to proceed smoothly at 80 °C in acetonitrile. As can be seen from Table 1 (runs 1–5), after a reaction period of 2 hours the monomer conversion of 47–84% was observed by 1H NMR analysis, yielding corresponding polymers with number-average molecular weight of 5270–8440 and low polydispersity indices (PDI = 1.09–1.12). Of five triflates tested, Sc(OTf)3 seems furnish best results in terms of the monomer conversion and PDIs of the resultant PEtOxs. In contrast, methyl tosylate (MeOTs) exhibited a poor initiating efficiency, which gave a monomer conversion of ∼16% and a low-molecular-weight polymer under the same conditions (run 6). Noteworthy is also the strong influence of the solvent property on the catalytic activity of Sc(OTf)3, with a decrease even deactivation observed when the polymerization was carried out in other solvents such as n-BuCl, DMSO or DMF instead of CH3CN (runs 7–9).
| Run | Catalyst | Solvent | [M] : [I] |
Temp. (°C) | Time (min) | PEtOx | ||
|---|---|---|---|---|---|---|---|---|
| Conv.c/% | Mnd | PDId | ||||||
| a The monomer concentration was 3 M for runs 1–9 and 4.5 M for runs 10–19.b The data given in parentheses are values obtained under microwave (μW) radiation.c Monomer conversion was determined by 1H NMR.d Determined by SEC, PMMA calibration, DMF containing 50 mM LiBr as the eluent.e Under μW conditions. | ||||||||
| 1 | Sc(OTf)3 | CH3CN | 100 | 80 | 120 | 84 | 5270 | 1.09 |
| 2 | Y(OTf)3 | CH3CN | 100 | 80 | 120 | 65 | 6220 | 1.12 |
| 3 | La(OTf)3 | CH3CN | 100 | 80 | 120 | 47 | 8440 | 1.12 |
| 4 | Dy(OTf)3 | CH3CN | 100 | 80 | 120 | 73 | 6270 | 1.10 |
| 5 | Lu(OTf)3 | CH3CN | 100 | 80 | 120 | 71 | 6290 | 1.11 |
| 6 | MeOTs | CH3CN | 100 | 80 | 120 | 16 | 3440 | 1.18 |
| 7 | Sc(OTf)3 | n-BuCl | 100 | 80 | 120 | 48 | 4230 | 1.21 |
| 8 | Sc(OTf)3 | DMSO | 100 | 80 | 120 | 0 | — | — |
| 9 | Sc(OTf)3 | DMF | 100 | 120 | 120 | 0 | — | — |
| 10b | Sc(OTf)3 | CH3CN | 100 | 80 | 60 | 53 | 4080 | 1.09 |
| (93) | (7880) | (1.10) | ||||||
| 11b | Sc(OTf)3 | CH3CN | 100 | 90 | 40 | 66 | 4490 | 1.10 |
| (91) | (8910) | (1.11) | ||||||
| 12 | Sc(OTf)3 | CH3CN | 100 | 140e | 5 | >99 | 5560 | 1.14 |
| 13 | MeOTs | 93 | 5320 | 1.27 | ||||
| 14 | Sc(OTf)3 | CH3CN | 400 | 140e | 20 | >99 | 12 210 |
1.31 |
| 15 | MeOTs | 69 | 15 320 |
1.52 | ||||
| 16 | Sc(OTf)3 | CH3CN | 800 | 140e | 20 | >99 | 19 060 |
1.56 |
| 17 | MeOTs | 67 | 14 850 |
1.97 | ||||
| 18 | Sc(OTf)3 | CH3CN | 1600 | 140e | 20 | 89 | 27 240 |
1.67 |
| 19 | MeOTs | 44 | 14 100 |
2.08 | ||||
In a set of comparable experiments, we found that the microwave (μW) irradiation makes Sc(OTf)3 more effective for the CROP of EtOx compared to the conventional heating, i.e., the rare-earth salt can catalyze the polymerization affording high yields (>90%) of higher-molecular-weight polymers with narrow PDIs under microwave conditions in acetonitrile at 80 or 90 °C (runs 10 and 11). Such a significant improvement in catalytic efficiency pointed to the existence of a non-heat microwave effect in the Sc(OTf)3-catalyzed polymerization. Nevertheless, for the reported μW-assisted CROPs of 2-oxazolines there seemed to be no distinct correlation between the accelerated polymerization rate and the microwave effect, as demonstrated by Hoogenboom and Schubert et al.34
It is also interesting to note that Sc(OTf)3 always exhibited higher catalytic efficiency than MeOTs in the μW-assisted polymerization process (runs 12–19). For example, the Sc(OTf)3-catalyzed reaction came to completion within 5–20 minutes at 140 °C in the cases of monomer-to-initiator ratio of 100, 400 and 800, even at a [M]/[I] of 1600 the monomer conversion reached ∼90% yielding Mn as high as 27
000. However, with decreasing of the catalyst loading a relatively broad molecular weight distribution was observed for the resultant PEtOxs (PDI value changing from 1.14 to 1.67), which may be attributed to chain transfers during the CROP.35
Table 2 summarizes the bulk polymerization results of EtOx using Sc(OTf)3 as the initiator under conventionally thermal heating. It can be seen that this bulk polymerization process allowed the synthesis of polymers with molar masses in the range 5400 g mol−1 < Mn < 18
800 g mol−1 (PMMA standard) through regulation of reaction parameters. In experimenting we found that with the proceeding of polymerization the reaction mixture became highly viscous and later solidified. It usually occurred at the point that the monomer conversion was 70–80%. Most likely, the presence of such a viscous or heterogeneous system at the late stage of polymerization is responsible for the broader molecular weight distributions (PDI = 1.27–1.37). This speculation is supported by the data from Table 2. More specifically, in the polymerization with a monomer-to-initiator ratio of 100 (runs 1–3), elongating the reaction time from 30 min to 60 min resulted in a distinct increase in both Mn and PDI; however, both the values almost remained constant when further extending periods of time. From the results shown in runs 6–8 of Table 2, it appears that the combination of a lower temperature and longer reaction time is beneficial to the improvement of molar masses while keeping narrower polydispersity for the case of high [M]/[I] ratio.
| Run | [M] : [I] |
Temp./°C | Time/min | PEtOx | ||
|---|---|---|---|---|---|---|
| Yielda/% | Mnb | PDIb | ||||
| a Isolated yield determined by gravimetry.b Determined by GPC, PMMA calibration, DMF containing 50 mM LiBr as the eluent. | ||||||
| 1 | 100 | 80 | 30 | 34.2 | 5450 | 1.09 |
| 2 | 100 | 80 | 60 | 89.0 | 12 040 |
1.29 |
| 3 | 100 | 80 | 90 | 89.4 | 12 920 |
1.29 |
| 4 | 100 | 100 | 30 | 86.4 | 8860 | 1.28 |
| 5 | 200 | 100 | 30 | 86.1 | 9030 | 1.37 |
| 6 | 500 | 100 | 30 | 70.1 | 13 800 |
1.26 |
| 7 | 500 | 100 | 180 | 97.8 | 12 410 |
1.64 |
| 8 | 500 | 80 | 180 | 85.0 | 18 730 |
1.32 |
![]() | ||
| Fig. 1 (a) Kinetic plots for 2-ethyl-2-oxazoline polymerizations initiated with Sc(OTf)3 ([M]0/[I]0 = 100) in acetonitrile ([M]0 = 4.5 M) at 60, 70, 80, 90, and 100 °C under thermal heating; the apparent rate constants (kp) are expressed in L mol−1 min−1. (b) Evolution of the molar mass (Mn) and the PDI value with monomer conversion (determined by SEC, RI detection, PMMA calibration, DMF containing 50 mM LiBr as the eluent. See: Fig. S2 in ESI†). | ||
The fact that ln[M]0/[M]t increased linearly with reaction time (Fig. 1a) indicates that the concentration of propagating species remained constant throughout the polymerization. Assuming that the concentration of the active species [P*] was equal to the initial initiator concentration [I]0, the formula (eqn (1)) for determining the apparent rate constant (kp) was integrated in eqn (2). Then, kp values can be deduced by the slope of the ln([M]0/[M]t versus time, as shown in Fig. 1a. From the graphical resolution of the Arrhenius equation (eqn (3) and Fig. 2), the activation energy (Ea) and the frequency factor (A) are calculated as 69.54 kJ mol−1 and 9.77 × 107 L mol−1 s−1, respectively. The Ea value is comparable to that obtained in the MeOTs initiating system.8
![]() | (1) |
![]() | (2) |
| kp = Ae−Ea/RT | (3) |
![]() | ||
| Fig. 2 Arrhenius plot for the polymerizations of 2-ethyl-2-oxazoline with Sc(OTf)3 in acetonitrile (see Fig. 1a for kp values). | ||
As evidenced by the kinetic results, Lewis acidic Sc(OTf)3 is more reactive towards EtOx polymerization than other existing initiators such as methyl tosylate8 and alkyl iodides.36 With the rare-earth catalyst the high yield synthesis of PEtOx was achieved in a rapid manner (within several hours) under mild conditions. Furthermore, this living system can be expected for the preparation of block copolymers with a well-defined structure. For instance, we have successfully synthesized a diblock polymer comprised of PEtOx and PPhOx segments via a sequential polymerization route, wherein the formation of copolymer was confirmed by the observed SEC analyses of a peak shift from 20.3 to 18.4 min elution time from the first (PEtOx) to second block growth profiles and a narrow molecular weight distribution (PDI = 1.18) (Fig. 3).
Single-crystal X-ray diffraction analysis of Sc(OTf)3–EtOx complex revealed extensive coordinative interactions between the imine units of the monomer and the scandium cation (Fig. S4 in the ESI†). It is conceivable that the coordination would promote the nucleophilic attack of free monomers at the C-5 position of the bound oxazoline unit to generate an oxazolinium propagating species, which has been confirmed by in situ NMR spectroscopic studies. As shown in Fig. 5, the EtOx polymerization system before being subjected to the termination (i.e. a living polymer) displays the weak proton signals at 4.26 and 4.90 ppm (denoted as a and b). The two peaks should come from the active oxazolinium units, because they are different from those due to methylene protons of monomer (δ: 3.71, 4.17 ppm) and those of the polymer main chain (δ: 3.6–3.3 ppm) as well; furthermore, they vanished totally upon addition of terminating agents such as piperidine (see: Fig. S1 in the ESI†).
![]() | ||
| Fig. 5 1H NMR spectrum of a living polymer sample formed in the EtOx polymerization in CD3CN at 80 °C for 30 min ([EtOx]/[Sc] = 30). | ||
Simultaneously, 13C NMR spectroscopy provides us with more information about the propagating species. In the 13C NMR spectra of Sc(OTf)3, the anionic ligand displays a characteristic quartet at 114.4–123.8 ppm, and the position of these signals does not change distinctly in the presence of EtOx at ambient temperature (Fig. 6a and b). However, a ca. 1.2 ppm downfield shift was observed upon heating the mixture at 80 °C for 30 min (Fig. 6c), indicating that the ligand moved to the propagating ends as a bound counterion with the polymerization went on. This situation is similar to that seen in the ligand exchange of rare-earth triflates reported by Ling et al.37
![]() | ||
| Fig. 6 13C NMR spectra in CD3CN of (a) Sc(OTf)3, (b) Sc(OTf)3 + EtOx (30 eq.), 25 °C, and (c) the mixture composed of Sc(OTf)3 and EtOx (30 eq.) after heating for 30 min at 80 °C. | ||
It is noteworthy that the 13C NMR signal of the ligand −OTf still remained its intrinsic quartet feature in the living polymerization system (Fig. 6), which means they are in the identical chemical environments as counterions. Hence we presume that every metallic active species in situ-generated from the Sc(OTf)3–EtOx complexation may possess three identical chain-growing sites, in other words, one Sc(OTf)3 initiated three equivalent PEtOx chains. To clarify this point, we devised a set of control experiments, in which the parallel EtOx polymerizations initiated with Sc(OTf)3 under the same conditions ([M] = 4.5 M, [M]/[I] = 100, 90 °C, 60 min) were terminated by two kinds of terminating agents, the monofunctional piperidine and the bifunctional piperazine. Rather, the termination reaction was conducted at a predetermined terminator-to-initiator molar ratio ([T]/[Sc]) to prepare the desired polymers (P1–P5) for SEC analysis. Among them, P1, as a reference, was obtained by using intentionally excess piperidine ([T]/[Sc] = 6) in the termination step, while P2–P5 resulted from the piperazine termination with [T]/[Sc] of 0.5, 1.0, 1.5, and 2.0, respectively.
As shown in Fig. 7, P1 gives a unimodal molecular weight distribution with Mn = 3310 (PDI = 1.09) in THF. In contrast, the SEC traces of P2–P5 exhibit a bimodal shape and the peak of low molecular weight components (LMW, arbitrarily defined as Mn ≈ 3 kD) was found to decrease in going from P2 to P4 in concert with an increase in the terminator amount used in termination reactions. The Mn values of high molecular weight (HMW) polymers are roughly twice Mn's of LMW parts. These observations may be explained by the idea proposed above that one Sc(OTf)3 leads to three living polymer chains combination with Scheme 3 illustrating the initiating/propagating mechanism. In the piperazine-termination with [T]/[Sc] = 0.5, the nucleophilic sites (secondary amino groups) were not in sufficient quantities to stop all the living chains and thereby the resultant polymer P2 is composed of the piperazine-coupled HMW PEtOx and non-coupled LMW PEtOx, as evidenced by its SEC curve (Fig. 7). It was also the case for the termination with [T]/[Sc] = 1.0, but the relative content of LMW components in P3 decreased remarkably compared with P2.
![]() | ||
| Scheme 2 Synthesis of 2-phenyl-2-oxazoline (PhOx) derivatives bearing Boc-protected L-proline moiety. | ||
![]() | ||
| Scheme 3 Schematic representation of the cationic ring-opening polymerization of 2-ethyl-2-oxazoline (EtOx) with Sc(OTf)3 as the initiator. | ||
Interestingly, the SEC elution trace of P4 is close to a unimodel shape, indicative of this sample consisted of the coupled polymers predominantly. This result seems to be understandable, because in the case of [T]/[Sc] = 1.5 the bifunctional terminator just provides equivalent nucleophilic cites relative to the cationic propagating ends, whereby every piperazine molecule can catch two active chains in average during the termination process yielding the coupled product. The very minor distribution occurred in the SEC profile reveals the presence of a few PEtOx polymers resulting from chain transfer to the moisture or/and nucleophilic impurity that enter reaction vessel. On the other hand, an excess of piperazine would retard such coupling reactions owing to partial terminators is bound to serve as the monofunctional capping agent in this case, which increasing the content of LMW polymers, as shown in SEC curve of P5 (Fig. 7).
| Run | Monomer | [M] (mol L−1) | [M] : [I] |
Solvent | T/°C | t min−1 | Polymera | ||
|---|---|---|---|---|---|---|---|---|---|
| Yieldb/% | Mnc | PDIc | |||||||
| a The structural characterization of polymers are given in Supporting Information.b Isolated yield.c Determined by SEC, PMMA calibration, DMF containing 0.05 M LiBr as the eluent.d Under μW conditions. | |||||||||
| 1 | 4-EtBuOx | Neat | 60 | Bulk | 120 | 180 | 48 | 4650 | 1.13 |
| 2 | 4-MeBuOx | Neat | 60 | Bulk | 120 | 180 | 38 | 4530 | 1.15 |
| 3 | 5-MeBuOx | Neat | 60 | Bulk | 120 | 180 | 26 | 4230 | 1.20 |
| 4 | PhOx | 4.5 | 100 | CH3CN | 90 | 180 | 31 | 5510 | 1.21 |
| 5 | ProPhOx-1 | 1.0 | 20 | CH3CN | 90 | 180 | 85 | 2460 | 1.14 |
| 6 | ProPhOx-2 | 1.0 | 20 | CH3CN | 90 | 180 | — | — | — |
| 7 | ProPhOx-3 | 1.0 | 20 | CH3CN | 90 | 180 | — | — | — |
| 8 | ProPhOx-2 | 1.0 | 20 | CH3CN | 140d | 20 | — | — | — |
| 9 | ProPhOx-3 | 1.0 | 20 | CH3CN | 140d | 20 | 46 | 5840 | 1.42 |
Footnote |
| † Electronic supplementary information (ESI) available. CCDC 1026078. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c4ra11404c |
| This journal is © The Royal Society of Chemistry 2014 |