Keita
Fuchise
*,
Masayasu
Igarashi
,
Kazuhiko
Sato
* and
Shigeru
Shimada
*
Interdisciplinary Research Center for Catalytic Chemistry, National Institute of Advanced Industrial Science and Technology (AIST), 1-1-1 Higashi, Tsukuba, Ibaraki 305-8565, Japan. E-mail: k-fuchise@aist.go.jp; k.sato@aist.go.jp; s-shimada@aist.go.jp
First published on 19th February 2018
Organocatalytic controlled/living ring-opening polymerization of cyclotrisiloxanes, such as hexamethylcyclotrisiloxane, 1,3,5-trimethyl-1,3,5-triphenylcyclotrisiloxane, 1,3,5-trimethyl-1,3,5-trivinylcyclotrisiloxane, and 1,3,5-trimethyl-1,3,5-tris(3,3,3-trifluoropropyl)cyclotrisiloxane, using water as an initiator and strong organic bases, such as amidines, guanidines, phosphazene bases, and proazaphosphatrane, as catalysts produced a variety of polysiloxanes with controlled number-average molecular weights (Mn = 2.64–102.3 kg mol−1), narrow polydispersity (Đ = 1.03–1.16), and well-defined symmetric structures. Controlled syntheses of statistical copolymers and triblock copolymers were achieved by copolymerizations of two cyclotrisiloxanes. Various terminal functionalities were successfully introduced by the end-capping reaction of propagating polysiloxanes using functional chlorosilanes. Kinetic investigations demonstrated that the polymerization proceeded through the initiator/chain-end activation mechanism, namely activations of water in the initiation reaction and of terminal silanols in propagating polysiloxanes in the propagation reaction. Catalytic activities of strong organic bases were revealed to depend on their Brønsted basicity and efficiency of the proton transfer in the initiation and propagation reactions. Guanidines possessing an R–NC(N)–NH–R′ unit, in particular 1,3-trimethylene-2-propylguanidine, showed excellent performance as a catalyst. In this system, even non-dehydrated solvents are usable for the polymerization.
In the field of polymer science, there has been continuous motivation to develop controlled polymerization that can be conducted with simple starting materials and procedures. Organocatalytic polymerization has been studied over the last decades to realize it.18–21 In particular, organocatalytic ROP of cyclic monomers, such as lactones, cyclic carbonates, epoxides, and cyclic phosphoesters, has been studied using organic acids, such as sulfonic acids, bis(sulfonyl)imides, and phosphoric acids, as well as organic bases, such as amines, amidines, guanidines, phosphazene bases, proazaphosphatranes, cyclopropenimines, and N-heterocyclic carbenes (NHC), as catalysts.18–21 Many of them proceed in a controlled/living fashion and have been extensively applied to controlled synthesis of various well-defined polymers, including end-functionalized polymers, block copolymers, and star-shaped polymers. The success of the organocatalytic ROP inspired us to apply it to cyclooligosiloxanes and develop a convenient method to synthesize various polysiloxanes with well-defined structures.
ROP of cyclooligosiloxanes12,22–32 and cyclic carbosiloxanes33,34 have been attempted with organic acids and bases as catalysts. Regarding acidic catalysts, HOSO2CF3,22,23 HN(SO2CF3)2,24 HB(C6F5)4,25 Ph3CB(C6F5)4,26 and B(C6F5)3 (ref. 27,28) have been used for ROP of hexamethylcyclotrisiloxane (D(Me2)3). However, none of them produced poly(dimethylsiloxane) (PDMS) with a narrow molecular weight distribution. For the basic catalysts, only those with quite high Brønsted basicity, such as phosphazene bases, NHCs, and bicyclic guanidines, have been employed. Möller and coworkers were the first to employ an electronically neutral strong organic base. They reported a polymerization of octamethylcyclotetrasiloxane (D(Me2)4) using 1-tert-butyl-4,4,4-tris(dimethylamino)-2,2-bis[tris(dimethylamino)phosphoranylidenamino]-2λ5,4λ5-catenadi(phosphazene) (tBu-P4) as a catalyst and methanol as an initiator.12,29 The polymerization was very rapid and the resulting PDMS had a very high number-average molecular weight (Mn) up to 440 kg mol−1, although polydispersity (Đ) was broad, 1.7–1.9. Hupfeld and Taylor also reported that a polymerization of D(Me2)4 using tBu-P4 as a catalyst almost immediately produced PDMS with Mn of up to 4 × 103 kg mol−1 with Đ of 1.5–1.9.30 Clarson's group reported that a polymerization of 1,3,5,7-tetramethyl-1,3,5,7-tetraphenylcyclotetrasiloxane (D(Me,Ph)4) using tBu-P4 as a catalyst and methanol as an initiator at ambient temperature produced poly[methyl(phenyl)siloxane] (PMPS) with Mn of 122–249 kg mol−1 and Đ of 1.5.31 Baceiredo and colleagues reported ROP of D(Me2)4 using NHCs, such as 1,3-dicyclohexylimidazol-2-ylidene and 1,3-di-tert-butyl-4,5-dimethylimidazol-2-ylidene, as catalysts and alcohols as initiators to give PDMS with high Đ of 1.5–1.7.32 Waymouth and Hedrick and coworkers reported the only example that succeeded in obtaining PDMS with narrow Đ (<1.2) by organocatalytic ROP of D(Me2)3. They used 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD) as a catalyst and 4-pyrenebutan-1-ol as an initiator, although spectral and chromatographic data of the products were not provided.33 Thus, controlled synthesis of polysiloxanes by organocatalytic polymerization has not been well established. We considered that an appropriate choice of a catalyst and an initiator is the key to develop a controlled/living ROP of cyclooligosiloxanes. The acidity/basicity of the catalyst should not be too high to avoid side reactions, such as main chain scission and condensation of propagating polysiloxanes, that affect Mn, Đ, and terminal structures of resulting polysiloxanes. Furthermore, an initiation reaction between a cyclooligosiloxane and an initiator should generate propagating polysiloxanes with a sufficiently stable terminal structure at a sufficiently fast rate in comparison with a propagation reaction.
We herein report a controlled/living ROP of cyclotrisiloxanes using water as an initiator, strong organic bases as catalysts, and organochlorosilanes as end-capping agents. As shown in Scheme 1, the developed system is capable of polymerizing various cyclotrisiloxane in a controlled/living fashion and producing a variety of (telechelic) polysiloxanes with controlled Mn, narrow molecular weight distributions, and well-defined symmetric structure, which are difficult to obtain by the conventional anionic ROP using lithium compounds as initiators.10,35–38 This new polymerization method has a further advantage that non-dehydrated solvents can be used to give well-defined polysiloxanes.
Scheme 1 Ring-opening polymerization (ROP) of cyclotrisiloxanes using water as an initiator and strong organic bases as catalysts. |
Entry | Catalyst | MeCNpKBH | (h−1) | Đ of the productc (Conv. (%)) |
---|---|---|---|---|
a The polymerizations were carried out under the conditions of [D(Me2)3]0 = 1.80 mol L−1 and [D(Me2)3]0/[H2O]0/[C]0 = 10/1/0.005 for Et-P2 and TiBP, 10/1/0.007 for TBD, and 10/1/0.10 for the other bases. b Calculated with . c Đ of the obtained PDMS at the indicated conversion of monomer. d Not reported. e Not determined. | ||||
1 | TMGa | 23.3 | 0.0026 | 1.08 (50.7) |
2 | DBN | 23.79 | 0.019 | 1.10 (98.5) |
3 | DBU | 24.34 | 0.039 | 1.11 (99.1) |
4 | MTBD | 25.43 | 0.29 | 1.11 (99.6) |
5 | TBO | 0.0036 | n.d.e | |
6 | TMGb | 0.093 | 1.13 (99.3) | |
7 | TBN | 24.55 | 0.16 | 1.11 (99.9) |
8 | TMiPG | 1.7 | 1.11 (99.5) | |
9 | TMnPG | 2.6 | 1.11 (98.0) | |
10 | TBD | 25.96 | 6.3 | 1.12 (99.1) |
11 | BEMP | 27.58 | 0.14 | 1.14 (99.2) |
12 | tBu-P1(pyrr) | 28.35 | 0.60 | 1.23 (99.5) |
13 | Et-P2 | 32.94 | 1.1 × 103 | 1.22 (99.8) |
14 | TiBP | 33.53 | 5.2 × 102 | 1.13 (99.9) |
The catalytic activity of the bases was determined from the apparent rate coefficients of propagation (kp,app, h−1) observed in the polymerizations in tetrahydrofuran (THF) at 30 °C under the conditions of [D(Me2)3]0 = 1.80 mol L−1 and [D(Me2)3]0/[H2O]0/[C (catalyst)]0 = 10/1/0.005 for Et-P2 and TiBP, 10/1/0.007 for TBD, and 10/1/0.10 for other bases. The polymerization was initiated by adding a stock solution of water in THF and that of the catalysts in dry THF to a solution of D(Me2)3 in dry THF in this order. A linear relationship between the polymerization time, t (h), and −ln(1 − c), where c is monomer conversion, was observed in the first-order kinetic plot of each polymerization. According to the eqn (1), the kp,app values were determined from the slope observed in the plots.
−ln(1 − c) = kp,appt | (1) |
The observed kp,app values were normalized to , as listed in Table 1 to directly compare kp,app values observed in the polymerizations using different amounts of the catalysts. This conversion should be rational, since we confirmed an almost linear relationship of kp,app with the initial molar ratio of the catalyst and water, [C]0/[H2O]0, as discussed in the next section (see Fig. 5).
Fig. 1 shows the dependence of values on MeCNpKBH of strong organic bases. MeCNpKBH instead of those in THF can be used for the discussion, since Leito's group has reported that pKBH values in MeCN and THF have a good linear correlation.41,42 The catalytic activity of an organic base per its Brønsted basicity, , increased in the following order: TiBP < phosphazene bases < guanidines/amidines A < guanidines B, which indicated that MeCNpKBH values of the catalysts were not the only single factor that affected the rates of polymerization. Interestingly, the amidines/guanidines A and the phosphazene bases independently showed linear relationships between and their MeCNpKBH.
Among the phosphazene bases and TiBP with MeCNpKBH values of 27.5–33.5 (Table 1, entries 11–14), Et-P2 (1.1 × 103 h−1) and TiBP (5.2 × 102 h−1) showed significantly higher catalytic activity than the other bases due to their very high Brønsted basicity. On the other hand, BEMP (0.14 h−1) and tBu-P1(pyrr) (0.60 h−1) showed catalytic activities that were only comparable to MTBD (0.29 h−1) and TBN (0.16 h−1), which have more than 100 times weaker Brønsted basicity. The very high Brønsted basicity of the phosphazene bases and TiBP caused the frequent occurrence of undesired side reactions (vide infra), in particular, in the polymerizations catalyzed by Et-P2 and tBu-P1(pyrr), which made these bases unfavorable for controlling the polymerization as can be seen from the Đ of the products in each polymerization.
Regarding the amidines/guanidines A with MeCNpKBH values of 23–25.5 (Table 1, entries 1–4), TMGa (0.0026 h−1), DBN (0.019 h−1), DBU (0.039 h−1), and MTBD (0.29 h−1), showed only low catalytic activity, although undesired side reactions almost never occurred within the observed time range. In contrast, the guanidines B with MeCNpKBH values of 24.5–26 (Table 1, entries 7 and 10), such as TBD (6.3 h−1) and TBN (0.16 h−1), showed much higher than the amidines/guanidines A as shown in Fig. 1, even though their structures and Brønsted basicity were not very different. For example, TBD showed 22 times higher catalytic activity than MTBD, although its Brønsted basicity was only around 3.4 times stronger than MTBD. This high catalytic activity of the guanidines B may have originated from the R–NC(N)–NH–R′ unit that enables high efficiency of the initiation and propagation reactions as described in the section regarding the mechanism (see Scheme 3).
We hence evaluated the catalytic activity of four more guanidines B of which MeCNpKBH values have not been reported, such as 1,4,6-triazabicyclo[3.3.0]oct-4-ene (TBO), 1,1,2,3-tetramethylguanidine (TMGb), 1,3-trimethylene-2-isopropylguanidine (TMiPG), and 1,3-trimethylene-2-propylguanidine (TMnPG), since undesired side reactions rather frequently occurred in the polymerizations catalyzed by TBD and TBN. TMGb and TMiPG were newly synthesized in this study. TBO (0.0036 h−1), a bicyclic guanidine with two five-membered rings, showed much lower catalytic activity than TBD and TBN, presumably due to its much lower Brønsted basicity than TBD and TBN.43 TMGb (0.093 h−1), an acyclic tetramethylguanidine with an R–NC(N)–NH–R′ unit, showed 36 times higher catalytic activity than TMGa. The replacement of an H–NC(N)–NMe2 unit of TMGa with an Me–NC(N)–NH–Me unit certainly contributed to the increase in the catalytic activity. TMiPG (1.7 h−1) and TMnPG (2.6 h−1), monocyclic guanidines with a six-membered ring and the same number of carbon atoms as TBD, showed high catalytic activity next to TBD. The Brønsted basicity of TMnPG has been reported to be at least higher than TBN in CD3OD/D2O = 8/2 (w/w).43 TMiPG showed a slightly lower catalytic activity than TMnPG presumably because of the higher steric hindrance of the isopropyl group than that of the propyl group. It is noteworthy that the undesired side reactions were much less frequent in the polymerization catalyzed by TMiPG and TMnPG in comparison with that catalyzed by other bases with comparable catalytic activity. Hence, we identified TMnPG as the most suitable catalyst for the polymerization of cyclotrisiloxanes initiated by water.
Entry | [D(Me2)3]0/[H2O]0/[C]0 | Solventb | End-capping agent | Time (h) | Conv (%)c | M n (kg mol−1) | Đ | k p,app (h−1) | ||
---|---|---|---|---|---|---|---|---|---|---|
Calcdd | SECe | NMRc | ||||||||
a [D(Me2)3]0 = 1.80 mol L−1. b Volume ratios of two solvents are shown for the mixed solvents. c Determined by 1H NMR. d Calculated from Mn,calcd = [D(Me2)3]0/[H2O]0 × Conv. × (MW. of D(Me2)3 = 222.46) + (MW. of terminal structures). e Determined by SEC measurements in THF using polystyrene standards. | ||||||||||
1 | 10/1/0.10 | THF | None | 1.5 | 98.0 | 2.20 | 1.66 | 2.64 | 1.14 | 2.6 |
Me2PhSiCl | 2.47 | 1.90 | 3.17 | 1.11 | ||||||
2 | 10/1/0.05 | THF | Me2PhSiCl | 3 | 98.1 | 2.47 | 1.49 | 2.85 | 1.13 | 1.3 |
3 | 10/1/0.01 | DMAc/THF = 57/43 | None | 0.50 | 98.6 | 2.21 | 0.83 | 2.81 | 1.33 | 8.2 |
4 | 10/1/0.10 | CH2Cl2/THF = 65/35 | None | 1 | 98.6 | 2.21 | 1.80 | 2.70 | 1.12 | 4.6 |
5 | 25/1/0.25 | CH2Cl2/THF = 84/16 | None | 2 | 99.5 | 5.54 | 4.09 | 5.70 | 1.06 | 2.4 |
6 | 50/1/0.50 | CH2Cl2/THF = 91/9 | Me2PhSiCl | 1.57 | 93.8 | 10.7 | 9.75 | 12.2 | 1.04 | 1.7 |
7 | 100/1/1.0 | CH2Cl2/THF = 94/6 | Me2PhSiCl | 3 | 97.3 | 21.9 | 15.6 | 18.9 | 1.04 | 1.2 |
8 | 200/1/2.0 | CH2Cl2/THF = 96/4 | Me2PhSiCl | 5.5 | 95.0 | 42.5 | 31.8 | 39.9 | 1.06 | 0.47 |
9 | 550/1/5.5 | CH2Cl2/THF = 97/3 | None | 18 | 93.0 | 113.8 | 80.0 | 102.3 | 1.03 | 0.15 |
10 | 10/1/0.10 | CH2Cl2/THF = 57/43 | Me2HSiCl | 1 | 99.5 | 2.35 | 1.60 | 2.95 | 1.15 | |
11 | 10/1/0.10 | CH2Cl2/THF = 57/43 | Me2ViSiCl | 1 | 99.1 | 2.39 | 1.71 | 3.00 | 1.14 | |
12 | 10/1/0.10 | CH2Cl2/THF = 57/43 | AllylMe2SiCl | 1 | 99.3 | 2.42 | 1.67 | 2.77 | 1.13 | |
13 | 10/1/0.10 | CH2Cl2/THF = 57/43 | (ClCH2)Me2SiCl | 1 | 99.5 | 2.44 | 1.84 | 3.14 | 1.10 | |
14 | 10/1/0.10 | CH2Cl2/THF = 57/43 | (BrCH2)Me2SiCl | 1 | 98.8 | 2.52 | 1.71 | 3.08 | 1.13 | |
15 | 10/1/0.10 | CH2Cl2/THF = 57/43 | Me2(C6F5)SiCl | 1 | 99.3 | 2.68 | 1.72 | 3.38 | 1.13 | |
16 | 10/1/0.10 | CH2Cl2/THF = 57/43 | (EtO)3SiCl | 1 | 99.5 | 2.57 | 1.84 | 3.25 | 1.11 |
Fig. 2 SEC chromatograms of the PDMS-(OH)2 (Mn,NMR = 2.64 kg mol−1, Đ = 1.14) and the PDMS-(OSiMe2Ph)2 (Mn,NMR = 3.17 kg mol−1, Đ = 1.11) synthesized in THF at 30 °C under the conditions of [D(Me2)3]0/[H2O]0/[TMnPG]0 = 10/1/0.10 and [D(Me2)3]0 = 1.80 mol L−1 (Table 2, entry 1). |
The well-defined structures of the obtained PDMS-(OH)2 and PDMS-(OSiMe2Ph)2 were proven by 1H and 29Si{1H} NMR analysis as shown in Fig. 3. In each spectrum, signals due to the terminal groups and several monomeric units from the termini were separately observed from that due to the inner repeating units. In both 1H and 29Si{1H} NMR spectra of the PDMS-(OSiMe2Ph)2, the signals due to the PDMS-(OH)2 were not observed, which evidenced the quantitative end-capping of the PDMS-(OH)2.
Fig. 3 1H and 29Si{1H} NMR spectra of the synthesized PDMSs (Table 2, entry 1) in CDCl3: (a) 1H NMR spectrum of PDMS-(OH)2, (b) 1H NMR spectrum of PDMS-(OSiMe2Ph)2, (c) 29Si{1H} NMR spectrum of PDMS-(OH)2, (d) 29Si{1H} NMR spectrum of PDMS-(OSiMe2Ph)2. |
Structures of the obtained PDMS-(OH)2 and PDMS-(OSiMe2Ph)2 were further analyzed by positive ion matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS) using trans-2-[3-(4-tert-butylphenyl)-2-methyl-2-propenylidene]malononitrile (DCTB) as a matrix and sodium trifluoroacetate as a cationization agent. Only one series of peaks was observed in both of the spectra shown in Fig. 4, which indicated that the obtained PDMS-(OH)2 and PDMS-(OSiMe2Ph)2 consisted of only linear PDMS with two terminal hydroxy groups and two terminal dimethyl(phenyl)siloxy groups, respectively. The m/z values of the observed molecular ion peaks corresponded to the calculated molar mass of PDMS-(OH)2 and PDMS-(OSiMe2Ph)2 cationized by a sodium cation. Interestingly, the obtained PDMSs contained PDMSs of which the degree of polymerization was not a multiple of 3, although those PDMSs should not be produced if only a simple ring-opening reaction of D(Me2)3 occurred in the polymerization. This point will be discussed in a later section regarding the mechanism of the polymerization.
Fig. 4 Positive ion matrix-assisted laser desorption/ionization time-of-flight mass (MALDI-TOF MS) spectra of the synthesized PDMSs (Table 2, entry 1) measured in the reflector mode using DCTB as a matrix and sodium trifluoroacetate as a cationization agent. (a) PDMS-(OH)2 (Mn,NMR = 2.64 kg mol−1, Đ = 1.14) and (b) PDMS-(OSiMe2Ph)2 (Mn,NMR = 3.17 kg mol−1, Đ = 1.11). |
As already mentioned in the previous section, a linear relationship was observed in the first-order kinetic plot of the polymerizations carried out with [D(Me2)3]0/[H2O]0/[TMnPG]0 = 10/1/0.10 and 10/1/0.05 (Table 2, entries 1 and 2) as shown in Fig. 5a. The kp,app for the former and the latter polymerizations were determined to be 2.6 and 1.3, which suggested that the kp,app varies linearly with the initial ratio of the catalyst and water, i.e., [C]0/[H2O]0.
Fig. 5 (a) First-order kinetic plot and (b) the dependence of number-average molecular weight (Mn,NMR) and polydispersity index (Đ) of the resulting PDMS on the monomer conversion, c, for the polymerization of D(Me2)3 catalyzed by TMnPG in THF at 30 °C under the conditions of [D(Me2)3]0 = 1.80 mol L−1 and [D(Me2)3]0/[H2O]0/[TMnPG]0 = 10/1/0.10 (Table 2, entry 1) and 10/1/0.05 (Table 2, entry 2). |
Fig. 5b shows the dependence of Mn,NMR and Đ on monomer conversion (c). In both of the polymerizations, Mn,NMR increased linearly as c increased, while Đ remained in a range of 1.08–1.14 even until the late stage of polymerization. It was hence found that the propagation reaction was the dominant process in the polymerizations and undesired side reactions were not frequent.
Other than THF, dichloromethane (CH2Cl2) was usable as a solvent for the polymerization in homogeneous solutions. N,N-Dimethylacetamide (DMAc), an aprotic polar solvent, could also be applied for the polymerization (Table 2, entry 3), although reactions in this medium proceeded as an inhomogeneous mixture due to the low solubility of PDMS in DMAc. The contribution of each solvent to the increase in the rate of polymerization varied in the following order: DMAc (entry 3) ≫ CH2Cl2 (entry 4) > THF (entry 1). The occurrence of undesired side reactions (vide infra) increased depending on the content of each solvent in the following order: DMAc ≫ THF > CH2Cl2. Hence, CH2Cl2 was found to be the most suitable solvent for the polymerization.
The Mn,NMR of PDMS was linearly controllable by changing the initial feed ratio of D(Me2)3 and water for the polymerization performed in a mixed solvent of CH2Cl2 and THF at 30 °C for 1–18 h under the conditions of [D(Me2)3]0 = 1.80 mol L−1, [D(Me2)3]0/[H2O]0 = 10–550, and [D(Me2)3]0/[TMnPG]0 = 100 (Table 2, entries 4–9). The Mn,NMR values of the products linearly increased from 2.70 kg mol−1 to 102.3 kg mol−1, which agreed well with the values of Mn,calcd for each polymerization. Furthermore, all the products had low Đ ranging from 1.03 to 1.12 as determined by the SEC analysis as shown in Fig. 6. A condensation of propagating polymers gradually occurred when the conversion of monomers increased above 80–90%, in particular in the polymerizations carried out with high [D(Me2)3]0/[H2O]0 ratios (vide infra). The occurrence of the condensation was indicated by the appearance of an additional peak with an almost double molecular weight of the main peak in the SEC chromatogram.
Fig. 6 SEC chromatograms of PDMS synthesized by the polymerizations of D(Me2)3 initiated by water and catalyzed by TMnPG in a mixed solvent of CH2Cl2 and THF at 30 °C under different initial feed ratio of D(Me2)3 and water (Table 2, entries 4–9). |
Controlled synthesis of various telechelic polysiloxanes was easily achieved by an end-capping reaction using functional chlorosilanes. This was so far achieved solely with conventional anionic ROP using specially synthesized lithium compounds as an initiator.10,35–38 PDMS end-functionalized with dimethylhydrosilyl, dimethyl(vinyl)silyl, allyldimethylsilyl, (chloromethyl)dimethylsilyl, (bromomethyl)dimethylsilyl, dimethyl(2,3,4,5,6-pentafluorophenyl)silyl, and triethoxysilyl groups were obtained by the end-capping with the corresponding chlorosilanes at 30 °C for 15 min after the polymerization (Table 2, entries 10–16). 1H NMR, 29Si{1H} NMR, and MALDI-TOF MS spectra of the products proved the quantitative introduction of the functional groups as shown in Fig. S9–S15.† The products exhibited narrow Đ ranging from 1.10 to 1.15. Notably, telechelic polysiloxanes with halogenomethyl, 2,3,4,5,6-pentafluorophenyl, and triethoxysilyl groups with narrow Đ values were also obtained by our method, although these functional groups are potentially sensitive to the reaction conditions of the conventional anionic ROP of cyclotrisiloxanes using lithium compounds.
It is worth noting that, in contrast to the conventional anionic ROP initiated by lithium compounds, even non-dehydrated solvents were usable for the polymerization. D(Me2)3 using non-dehydrated THF (purity: >99.5%, stabilizer-free) as a solvent produced PDMS-(OSiMe2Ph)2 with Mn,NMR of 2.89 kg mol−1 and Đ of 1.11 under the same conditions as the polymerization shown in entry 1 of Table 2. Non-dehydrated CH2Cl2 (purity: >99.5%, stabilized with 2-methylbut-2-ene) was also usable (see ESI†). The maximum Mns reachable with non-dehydrated reagents and solvents depend on their water contents.45
Entry | Monomer | [M]0/[H2O]0/[C]0 | CH2Cl2/THF (v/v) | End-capping agent | Time (min) | Conv.a,b (%) | M n (kg mol−1) | Đ | k p,app (h−1) | (h−1) | |||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|
To polym. | To D4 | Calcd c | SEC d | NMR a | |||||||||
a Determined by 1H NMR. b Mol% of the monomer converted to the corresponding polysiloxanes and cyclotetrasiloxanes. c Calculated from Mn,calcd = [M]0/[H2O]0 × (Conv. to polymer) × (MW. of monomer) + (MW. of terminal structures). d Determined by SEC measurements in THF using polystyrene standards. e Not determined. f Around 1 h of an induction period was observed. | |||||||||||||
1 | D(Me,Ph)3 | 10/1/0.01 | 57/43 | Et3SiCl | 219 | 92.4 | 5.8 | 4.02 | 4.39 | 5.60 | 1.16 | 1.7f | 17 |
2 | D(Me,Ph)3 | 30/1/0.03 | 79/21 | Et3SiCl | 240 | 83.3 | 10.6 | 10.5 | 5.04 | 9.92 | 1.15 | 1.0f | |
3 | D(Me,Vi)3 | 10/1/0.01 | 57/43 | Me2PhSiCl | 72 | 95.5 | 3.1 | 2.76 | 2.73 | 3.64 | 1.11 | 3.5 | 35 |
4 | D(Me,Vi)3 | 30/1/0.03 | 79/21 | Me2PhSiCl | 51 | 80.6 | 5.0 | 6.65 | 5.28 | 5.68 | 1.15 | 2.4 | |
5 | D(Me,TFPr)3 | 10/1/0.004 | 70/30 | Me2PhSiCl | 40 | 92.8 | n.d.e | 4.63 | 2.94 | 6.00 | 1.12 | 3.9 | 99 |
6 | D(Me,TFPr)3 | 30/1/0.012 | 88/12 | Me2PhSiCl | 100 | 94.2 | n.d.e | 13.5 | 6.38 | 13.3 | 1.13 | 1.7 | |
7 | D(Me2)4 | 7.5/1/0.10 | 65/35 | None | 1500 | 6.2 | n.d.e | 0.16 | n.d.e | n.d.e | n.d.e | ||
8 | D(Me,Vi)4 | 7.5/1/0.01 | 57/43 | None | 1500 | 10.6 | n.d.e | 0.29 | n.d.e | n.d.e | n.d.e | ||
9 | D(Me2)5 | 6/1/0.10 | 65/35 | None | 1500 | 1.5 | n.d.e | 0.051 | n.d.e | n.d.e | n.d.e | ||
10 | D(Me2)3 + D(Me,Vi)3 | 25 + 8/1/0.05 | 70/30 | Me2PhSiCl | 540 | 64.9, >99.9 | n.d.e | 6.02 | 4.68 | 6.34 | 1.13 | ||
11 | 1st: D(Me2)3 | 25/1/0.05 | 67/33 | Me2PhSiCl | 540 | 84.9 | n.d.e | 4.74 | 4.16 | 5.11 | 1.09 | ||
2nd: D(Me,Vi)3 | 25/1/0.05 | 28 | 80.8 | 2.9 | 10.2 | 9.56 | 11.2 | 1.07 | |||||
12 | 1st: D(Me2)3 | 25/1/0.25 | 84/16 | Et3SiCl | 60 | 92.7 | n.d.e | 5.17 | 4.48 | 5.89 | 1.08 | ||
2nd: D(Ph2)3 | 7.5/1/0.25 | 44 | 87.4 | n.d.e | 9.30 | 8.71 | 11.7 | 1.06 |
In contrast to cyclotrisiloxanes, cyclotetrasiloxanes and cyclopentasiloxanes, such as D(Me2)4, D(Me,Vi)4 and decamethylcyclopentasiloxane (D(Me2)5), were almost non-reactive under similar reaction conditions for D(Me2)3 and D(Me,Vi)3 (Table 3, entries 7–9), even though tBu-P4 (MeCNpKBH = 42.7)39 and NHCs, such as 1,3-dialkylimidazol-2-ylidene and 1,3-dialkyl-4,5-dimethylimidazol-2-ylidene (estimated MeCNpKBH = 32–34 and 34–36, respectively),49 with much higher Brønsted basicity than TMnPG are known to catalyze the polymerization of cyclotetrasiloxanes.29–32
Consecutive copolymerizations of D(Me2)3 and D(Me,Vi)3 (ref. 52,53) as well as D(Me2)3 and hexaphenylcyclotrisiloxane (D(Ph2)3)54 gave triblock copolymers of PDMS and PMVS (PMVS-b-PDMS-b-PMVS, Table 3, entry 11) as well as PDMS and poly(diphenylsiloxane) (PDPS) (PDPS-b-PDMS-b-PDPS, Table 3, entry 12). The syntheses were successful even though homopolymerization of D(Ph2)3 using water and TMnPG was hard to control because of very low solubility of PDPS in common organic solvents.54–56 The polymerizations of D(Me2)3 were first carried out in CH2Cl2/THF at 30 °C under the conditions of [D(Me2)]0/[H2O]0/[TMnPG]0 = 25/1/0.05 for the former and 25/1/0.25 for the latter. The polymerizations were further continued by adding 25 equiv. (with respect to the initial amount of water) of D(Me,Vi)3 or 7.5 equiv. Of D(Ph2)3 after 540 min or 75 min from the initiation of the first polymerization. SEC chromatograms of the obtained PDMS-(OH)2, PMVS-b-PDMS-b-PMVS, and PDPS-b-PDMS-b-PDPS indicated that PDMS-(OH)2 generated in the first polymerization quantitatively initiated the second polymerization as shown in Fig. 7. The products did not have a gradient structure between the segments of PDMS and PMVS as well as PDMS and PDPS as observed in the 29Si{1H} NMR spectra shown in Fig. S20 (ref. 57) and S22,†22,54,58 although the conversions of D(Me2)3 in the first polymerization were not quantitative in both of the syntheses. This result indicated that D(Me2)3 almost did not react in the second stage of the polymerization due to very low reactivity of the propagating end of PMVS and PDPS against D(Me2)3. The obtained PDPS-b-PDMS-b-PDPS had a much narrower molecular weight distribution (Đ = 1.06) than those synthesized by the conventional anionic ROP using dilithium diphenylsilanediolate as an initiator (Đ = 1.4–1.8).54 These successful block copolymerizations demonstrated the controlled/living nature of the ROP of cyclotrisiloxanes using water and strong organic bases.
Fig. 7 SEC chromatograms of the products obtained in the consecutive polymerizations of (a) D(Me2)3 and D(Me,Vi)3 (Table 3, entry 11) and (b) D(Me2)3 and D(Ph2)3 (Table 3, entry 12). |
Scheme 2 Possible elementary reactions in the polymerization of cyclotrisiloxane using water as an initiator and strong organic bases as catalysts. |
Scheme 3 Proposed mechanism for the ring-opening reaction of cyclotrisiloxanes catalyzed by strong organic bases. |
Although only the intramolecular condensation of a propagating polysiloxane can be considered as a possible termination reaction, we did not observe this process in any polymerizations that we conducted. Hence, the polymerization is characterized by the absence of termination reactions, which ensures the controlled/living nature of the polymerization. The total numbers of reactive hydroxy groups in water, which has ‘two’ hydroxyl groups, and propagating polysiloxanes do not change throughout the polymerization even when side reactions (c), (d), and (e) occur. The polymerization can be terminated only by neutralization of the catalyst or end-capping of propagating polymers.
Another possible polymerization mechanism is the ‘nucleophilic monomer activation mechanism’ that is based on the activation of the monomer by the nucleophilic attack of the catalyst.19 However, if so, (1) catalytic activity of the strong organic bases would not be linear to their Brønsted basicity as shown in Fig. 1, since the nucleophilicity (‘silicophilicity’) of the bases and their basicity are independent as it was demonstrated in a condensation of a slanol and a organosilanes with a leaving group catalyzed by an organic base.65–67 Besides, (2) the first-order kinetic plot would show a convex curve when the interaction of a catalyst and a monomer is high.68
(2) |
[P*] is expressed by the product of the concentration of hydroxy groups not being activated, [P] (mol L−1), the concentration of catalyst not interacting with hydroxy groups, [C] (mol L−1), and the equilibrium constant for the activation of silanol groups by the catalyst, Ks (L mol−1), i.e., [P*] = Ks[P][C], since Ks is defined as Ks = [P*]/[P][C]. [P*] is hence expressed by eqn (3) using the initial concentration of hydroxy groups, [P]0 (mol L−1), the initial concentration of the catalyst, [C]0 (mol L−1), since [P] = [P]0 − [P*] and [C] = [C]0 − [P*].
[P*] = Ks([P]0 − [P*])([C]0 − [P*]) | (3) |
From the quadratic formula, [P*] is expressed by eqn (4) as the function of three constants, i.e., [P]0, [C]0, and Ks that varies depending on the catalyst, the temperature, and the solvent employed for a polymerization:
(4) |
[P*] is hence constant, regardless of monomer conversion, c, and frequency of the side reactions (c), (d), and (e). The integration of eqn (2) gives eqn (5):
−ln(1 − c) = kp[P*]t | (5) |
kp,app = kp[P*] | (6) |
Hence, the proposed mechanism for the polymerization shown in Scheme 2 were proven to be reasonable. The polymerization is considered to proceed through the initiator/chain-end activation mechanism.
Effects of reaction conditions on the kinetics of polymerization can be expected with eqn (4)–(6). Fig. 8a shows the expected dependence of kp,app on Ks and [C]0/[P]0 under the conditions of [M]0 = 1.80 mol L−1 and [M]0/[P]0 = 5. The calculated kp,app was normalized by dividing it by the kp,app on [M]0/[P]0 = 1. It was found that kp,app is almost proportional to [C]0/[P]0 when [C]0/[P]0 ≤ 1 regardless of Ks, which corresponded to the results shown in Fig. 5. In contrast, the increase in kp,app with increasing [C]0/[P]0 depends on Ks when [C]0/[P]0 > 1 and its rate of change decreases with increasing Ks. Fig. 8b shows the expected dependence of kp,app on Ks and [M]0/[P]0 under the conditions of [M]0 = 1.80 mol L−1 and [M]0/[C]0 = 100. The calculated 1/kp,app was normalized by multiplying it by the kp,app on [M]0/[P]0 = 5. It was found that kp,app is almost inversely proportional to [M]0/[P]0 when Ks is small, while kp,app only gradually decreases with increasing [M]0/[P]0 when Ks is large. Considering the values of kp,app observed in the polymerizations of D(Me2)3, D(Me,Ph)3, D(Me,Vi)3, and D(Me,TFPr)3 with different [M]0/[H2O]0 (entries 4, 5 of Table 2 and entries 1–6 of Table 3), the magnitude of Ks would increase in the following order: D(Me,TFPr)3 ≈ D(Me2)3 < D(Me,Ph)3 < D(Me,Vi)3.
Footnote |
† Electronic supplementary information (ESI) available: Experimental details and 1H NMR, 29Si{1H} NMR, and MALDI-TOF MS spectra of the synthesized polysiloxanes. See DOI: 10.1039/c7sc04234e |
This journal is © The Royal Society of Chemistry 2018 |