Bis(amido) rare-earth complexes coordinated by tridentate amidinate ligand: synthesis, structure and catalytic activity in the polymerization of isoprene and rac-lactide

Aleksei O. Tolpygin, Olesya A. Linnikova, Tatyana A. Glukhova, Anton V. Cherkasov, Georgy K. Fukin and Alexander A. Trifonov*
Institute of Organometallic Chemistry of Russian Academy of Sciences, Tropinina 49, GSP-445, 630950, Nizhny Novgorod, Russia. E-mail: trif@iomc.ras.ru; Fax: +7 831 4627497; Tel: +7 831 4623532

Received 29th December 2015 , Accepted 2nd February 2016

First published on 3rd February 2016


Abstract

A series of bis(amido) complexes of rare earth metals {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Ln(N(SiMe3)2)2, (Ln = Y (2), Nd (3), La (4)) was synthesized using the amine elimination reaction of Ln[N(SiMe3)2]3 (Ln = Y, Nd, La) and parent amidine 1 in THF, and the products were isolated in 60 (2), 61 (3) and 72% (4) yields, respectively. The X-ray studies revealed that complexes 2–4 are solvent-free and feature intramolecular coordination of the Ph2P[double bond, length as m-dash]O group to the Ln ion. Complexes 2–4 were investigated as precatalysts for isoprene polymerization. The ternary systems 2–4/borate/AlR3 (AlR3 = AliBu3, AliBu2H; borate = [Ph3C][B(C6F5)4], [PhNHMe2][B(C6F5)4]) were found to be active in isoprene polymerization and enable complete conversion of 1000 equivalents of monomer into polymer at 25 °C within 1–24 h affording polyisoprenes with polydispersities Mw/Mn = 1.12–9.46. The effect of the organoaluminium component and [Ln]/[AlR3] ratio on the catalytic activity and selectivity of the ternary catalytic systems was investigated. Complexes 2–4 proved to be efficient catalysts for the ring-opening polymerization of rac-lactide, which allow conversion of up to 500 equivalents of monomer into a polymer at room temperature within 30 min and afford atactic polylactides with high molecular weights and moderate molecular-weight distributions (1.29–2.12). Complexes 2–4 appear to be well-suited for achieving immortal polymerization of lactide, through the introduction of large amounts of isopropanol as a chain-transfer agent.


Introduction

Amido derivatives of the rare-earth metals have attracted a considerable deal of attention due to their high reactivity.1 They have been exhaustively explored as precursors for the preparation of electronic and ceramic materials,2 as well as catalysts for the synthesis of various valuable chemical products.3 Moreover they turned out to be efficient initiators for the polymerization of a large range of monomers such as ethylene,4 styrene,5 methyl methacrylate6 or cyclic esters.7 Formerly, in the pioneering work of Marks it was reported that the strength of the Ln–N bond is comparable to that of the Ln–C bond. By means of calorimetric titration of compounds Cp*2SmX (X = NMe2, CH(SiMe3)2) it was evidenced that the absolute bond disruption enthalpies (D) of Sm–N and Sm–C bonds have rather close values: X = NMe2, D = 48.2 kcal mol−1; X = CH(SiMe3)2, D = 47.0 kcal mol−1.8 Surprisingly, unlike alkyl species amido derivatives were scarcely investigated as precursors for the cis-stereoselective polymerization of butadiene or isoprene.9

Recently we demonstrated that bis(alkyl) complexes {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Ln(CH2SiMe3)2(THF)n (Ln = Y, n = 1, Ln = Er, n = 1, Ln = Lu, n = 0) containing tridentate amidinate ligand featuring intramolecular coordination of a side chain pendant Ph2P[double bond, length as m-dash]O group to metal ion are efficient components of ternary catalytic systems for isoprene polymerization.10 The systems Ln/borate/AliBu3 (borate = [PhNHMe2][B(C6F5)4], [Ph3C][B(C6F5)4]) enable complete conversion of 1000–10[thin space (1/6-em)]000 equivalents of isoprene into polymer at 20 °C within 0.5–2.5 h affording polyisoprenes with very high content of 1,4-cis units (up to 96.6%) and rather narrow polydispersities. However, a common disadvantage of bis(alkyl) species is their thermal sensibility and difficulties to handle and to store. In order to find new efficient catalytic systems based on less sensitive and more available compounds we focused on the synthesis of rare-earth bis(amido) complexes coordinated by tridentate amidinate ligand {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}. A comparative study of catalytic systems based on LLnR2 complexes (R = CH2SiMe3, N(SiMe3)2) can provide a deeper insight into the effect of leaving group in rare-earth complex on catalytic activity and selectivity of isoprene polymerization. Moreover application of bulky silylamido N(SiMe3)2 ligand makes possible to use derivatives of rare-earth metals having larger ion size (La, Nd) whose alkyl species are hardly accessible. At the same time derivatives of these metals are of special interest since normally provide higher 1,4-cis selectivity in diene polymerization. Herein we report on the synthesis, structure and catalytic activity of bis(amido) complexes {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Ln(N(SiMe3)2)2 (Ln = Y, Nd, La) and their catalytic activity in isoprene polymerization as well as in ring-opening polymerization of rac-lactide.

Results and discussion

To synthesize a series of bis(amido) rare-earth complexes coordinated by tridentate amidinate ligand bearing a pendant Lewis base Ph2P[double bond, length as m-dash]O group linked to a NCN fragment via a conformationally rigid o-phenylene linker the amine elimination approach was used. The equimolar amounts of Ln[N(SiMe3)2]3 (Ln = Y, Nd, La)11 and amidine 2-[Ph2P(O)]PhNHC(tBu)[double bond, length as m-dash]N(2,6-Me2C6H3) (1)10 were reacted in THF solution at room temperature (Scheme 1). After removal of the volatiles in vacuum and recrystallization of the resulted solid from THF/hexane mixture complexes {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Ln[N(SiMe3)2]2 (Ln = Y (2), Nd (3), La (4)) were isolated as pale yellow (2 and 4) or green (3) crystals in 60, 61 and 72% yields respectively. Complexes 2–4 are air and moisture sensitive, well soluble in THF and toluene and sparingly soluble in hexane.
image file: c5ra27960g-s1.tif
Scheme 1

The 1H NMR (400 MHz, C6D6, 25 °C) spectra of diamagnetic complexes 2 and 4 exhibit singlets at 0.39 ppm (2) and 0.37 ppm (4) due to the protons of N(SiMe3)2 group. The amidinate ligand appears as the expected set of resonances. The 31P{1H} (161.9 MHz, C6D6, 25 °C) spectrum of 2 displays a doublet at 41.2 ppm (2JPY = 6.4 Hz) due to coupling with 89Y nuclei while in that of complex 4 a singlet at 39.7 ppm was observed.

The molecular structures of complexes 2–4 were established by X-ray diffraction studies. Transparent crystals of complexes 2–4 suitable for X-ray diffraction studies were obtained by slow condensation of hexane into concentrated THF solution of complexes. The molecular structure of complexes 2–4 is shown in Fig. 1. Crystallographic data and structure refinement details are given in Table S1 (see ESI). Complexes 2–4 are isomorphous and crystallize in the monoclinic (C2/c) space group with eight molecules per unit cell. The X-ray studies revealed that complexes 2–4 feature intramolecular coordination of Ph2P[double bond, length as m-dash]O group to metal ion. In 2–4 the lanthanide ions are coordinated by two nitrogen and one oxygen atoms of the tridentate amidinate ligand and by two amido N(SiMe3)2 groups, thus resulting in coordination number of five. Despite high oxophilicity of rare earth metals coordination sphere of complexes 2–4 does not contains coordinated THF molecules. The C–N bonds within the amidinate fragment are slightly different (2: 1.332(2) and 1.353(2) Å; 3: 1.331(2) and 1.353(2) Å; 4: 1.330(3) and 1.354(4) Å). The M–N(amidine) bonds are non-equivalent: 2.461(2) and 2.416(2) Å (2); 2.565(2) and 2.508(2) Å (3); 2.614(3) and 2.563(2) Å (4). The Ln–N(amidine) bonds in 2–4 are longer compared to the related five-coordinate amidinate and guanidinate complexes (for comparison see: [4-Me-PhC(N-2,6-iPr2C6H3)2]Y(NHSiMe2)2(THF) 2.412(3), 2.351(2) Å;12 [(SiMe3)2NC(NCy)2]Y[N(SiHMe2)2]2(THF) 2.356(4), 2.328(3) Å;13 [PhC(N-2,6-Me2C6H3)2]2NdN(SiMe3)2 2.462(3)–2.484(3) Å, [CyC(N-2,6-Me2C6H3)2]2NdN(SiMe3)2 2.448(4)–2.472(4) Å;14 [2-NC5H4(CH2)2NC(p-MeC6H4)NPh]2La[N(SiMe3)2] (2.491(4)–2.609(4) Å)15. Obviously this difference in bond lengths originates from the coordination of sterically demanding Ph2P[double bond, length as m-dash]O group to the metal ions. The M–N(N(SiMe3)2) bond lengths are also somewhat different: 2.251(2) and 2.301(2) Å (2); 2.335(2) and 2.385(2) Å (3); 2.389(3) and 2.447(2) Å (4). The Ln–N(N(SiMe3)2) bond lengths in complexes 2–4 are in a good agreement with the values previously published for related five-coordinate amido species: [4-Me-PhC(N-2,6-iPr2C6H3)2]Y(NHSiMe2)2(THF), 2.250(3), 2.258(3) Å;12 [(SiMe3)2NC(NCy)2]Y[N(SiHMe2)2]2(THF) 2.241(4), 2.278(4) Å;13 CyNC{[N,N-(2,5-Me2C4H2N)CH2CH2]2N}NCyLn[N(SiMe3)2]2 (Ln = La, Nd)16 Nd–N 2.327(2), 2.341(2) Å, La–N 2.386(2), 2.400(2); La[2-NC5H4-(CH2)2NC(p-MeC6H4)NPh]2[N(SiMe3)2]) 2.453(4) Å.15 The MNCN fragments in complexes 2–4 are not planar. It is noteworthy that the value of the dihedral angle between the MNM and NCN planes (171.8° (2), 172.7° (3) and 172.9° (4)) increases with increase of ionic radius of the central metal atom (Y: 0.900 Å, Nd: 0.983 Å, La: 1.032 Å).17 The length of coordination bonds between metallocenter and oxygen atom of Ph2P[double bond, length as m-dash]O group (2.253(1) Å (2), 2.367(1) Å (3), 2.418(2) Å (4)) are in a good agreement with the values previously published for related five-coordinate amido species: [[(Me3Si)2N]2(Ph3PO)Y}2(μ-η22-N2) (Y–O 2.2545(11)–2.2437(11)),18, [La[(Me3Si)2N]2(PPh2)(Ph2PO)2] (La–O 2.472(8) Å)19].


image file: c5ra27960g-f1.tif
Fig. 1 ORTEP diagram (30% probability thermal ellipsoids) of 2–4 (M = Y (2), Nd (3), La (4)) showing the atom numbering scheme. Hydrogen atoms and carbon atoms in Ph-fragments of Ph2P[double bond, length as m-dash]O group are omitted for clarity. Selected bond lengths [Å] and angles [°] in 2: Y(1)–O(1) 2.253(1), Y(1)–N(1) 2.461(2), Y(1)–N(2) 2.416(2), Y(1)–N(3) 2.251(2), Y(1)–N(4) 2.301(2), N(1)–C(1) 1.332(2), N(2)–C(1) 1.353(2), N(1)–C(1)–N(2) 109.5(2), N(2)–Y(1)–N(1) 53.43(5), N(3)–Y(1)–N(4) 112.89(6). Selected bond lengths [Å] and angles [°] in 3: Nd(1)–O(1) 2.367(1), Nd(1)–N(1) 2.565(2), Nd(1)–N(2) 2.508(2), Nd(1)–N(3) 2.335(2), Nd(1)–N(4) 2.385(2), N(1)–C(1) 1.331(2), N(2)–C(1) 1.353(2), N(1)–C(1)–N(2) 110.4(2), N(2)–Nd(1)–N(1) 51.48(5), N(3)–Nd(1)–N(4) 113.65(6). Selected bond lengths [Å] and angles [°] in 4: La(1)–O(1) 2.418(2), La(1)–N(1) 2.614(3), La(1)–N(2) 2.563(2), La(1)–N(3) 2.389(3), La(1)–N(4) 2.447(2), N(1)–C(1) 1.330(3), N(2)–C(1) 1.354(4), N(1)–C(1)–N(2) 110.6(3), N(2)–La(1)–N(1) 50.45(7), N(3)–La(1)–N(4) 113.86(9).

Isoprene polymerization initiated by complexes 2–4

Bis(alkyl) rare earth species established a reputation as suitable precursors for catalytic systems performing controlled diene polymerization with high activities and selectivities.9d,20 In contrast, the application of rare earth metal amide complexes as precatalysts for isoprene polymerization remains scarce.9

The catalytic activity of complexes 2–4 in isoprene polymerization was explored in toluene at room temperature. The polymerization results are summarized in Table 1. It was established that neither complexes 2–4 nor binary systems 2–4/AliBu3, 2–4/AliBu2H do not perform catalytic activity. Unlike the related bisamides coordinated by bidentate amidinate ligand [PhC(N-2,6-R2C6H3)2]Ln[N(SiHMe2)2]2 (Ln = Sc, Y; R = Me, iPr)9c,9e complexes 2–4 activated by equimolar amounts of [Ph3C][B(C6F5)4] or [PhNHMe2][B(C6F5)4] turned out to be inactive in isoprene polymerization. However this is in line with the observation of Luo and co-workers9d who reported that yttrium bis(amido) amidinates containing N(SiMe3)2 groups instead of N(SiHMe2)2 do not perform catalytic activity in isoprene polymerization in the presence of borates. Upon combination with borate ([Ph3C][B(C6F5)4], [PhNHMe2][B(C6F5)4]) and alkylaluminium compounds (AliBu3, AliBu2H) complexes 2–4 enable isoprene polymerization. However catalytic activity of the ternary systems was found to be dramatically influenced by the nature of the metal atom, borate and alkyl aluminium components. Thus complexes 2 and 3 enable isoprene polymerization at ambient temperature (molar ratio [IP]/[Ln] = 1000) with 45–100% conversion of monomer into polymer (Table 1) in 1 h. Surprisingly complex 4 containing the largest La3+ ion17 displayed the lowest activity. The systems based on 4 required 24 h to achieve 92–100% conversions. Complex 2 turned out to be indifferent to the nature of borate or alkyl aluminium components: for both AlR3 and both borates the polyisoprene (PIP) yields were 92–100%, though complex 3 is somewhat more prone to the influence of the nature of AlR3 and borate components. The complex 3 based systems containing AliBu2H performed higher activities than those with AliBu3 for both borates. Neodymium complex 3 in the presence of AliBu3 was noticeably less active than 2 with both borates (64 vs. 92 for [Ph3C][B(C6F5)4] and 45 vs. 97 for [PhNHMe2][B(C6F5)4]). However when AliBu2H was used catalytic activities of complexes 2 and 3 became similar.

Table 1 Catalytic tests in isoprene polymerization initiated by systems 2–4/(AliBu3, AliBu2H)/([Ph3C][B(C6F5)4], [PhNHMe2][B(C6F5)4])a
  Comp. Borate AlR3 [IP]/[Ln] t, h Yield, % Cis-1,4 Trans-1,4 3,4- Mnb (×10−3) Mncalcc (×10−3) Mn/Mw
a Conditions: complex (10 μmol in toluene, [AlR3][thin space (1/6-em)]:[thin space (1/6-em)][Ln][thin space (1/6-em)]:[thin space (1/6-em)][borate] = 10/1/1, 1/1/1, T: 25 °C); HNB = [PhNHMe2][B(C6F5)4], TB = [Ph3C][B(C6F5)4].b Determined by GPC against polystyrene standard.c Mcalc = ([IP]/[Ln]) × 68.12 × (conversion), Y* = complex 5.10
1 Y TB 10AliBu3 1000 1 92 76.2 18.2 5.6 30.66 62.67 4.18
2 Y HNB 10AliBu3 1000 1 97 78.5 17.2 4.3 42.58 66.07 2.25
3 Y TB 10AliBu2H 1000 1 100 72.3 21.4 6.3 19.01 68.12 1.58
4 Y HNB 10AliBu2H 1000 1 96 71.5 21.9 6.6 10.04 65.39 1.53
5 Y TB 1AliBu2H 1000 1 54 76.7 15.0 8.3 62.55 36.78 1.70
6 Y HNB 1AliBu2H 1000 1 78.8 79.4 15.8 4.8 55.50 53.67 1.47
7 Nd TB 10AliBu3 1000 1 64 82.6 9.6 7.8 19.71 43.59 9.06
8 Nd HNB 10AliBu3 1000 1 45 88.7 4.9 6.4 13.73 30.65 9.46
9 Nd TB 10AliBu2H 1000 1 100 55.9 37.4 6.7 11.20 68.12 1.23
10 Nd HNB 10AliBu2H 1000 1 95 55.9 34.7 9.4 14.65 64.71 1.21
11 Nd TB 1AliBu2H 1000 1 84.3 68.6 23.0 8.4 59.35 57.42 1.22
12 Nd HNB 1AliBu2H 1000 1 100 89.6 3.4 7.0 115.36 68.12 1.627
13 La TB 10AliBu3 1000 24 94 45.2 46.8 8.0 13.48 64.03 2.76
14 La HNB 10AliBu3 1000 24 100 61.3 32.0 6.7 57.68 68.12 1.69
15 La TB 10AliBu2H 1000 24 96 38.6 52.0 9.4 8.45 65.39 2.21
16 La HNB 10AliBu2H 1000 24 100 39.8 52.7 7.5 6.74 68.12 2.96
17 La TB 1AliBu2H 1000 24 100 70.6 22.4 7.0 134.19 68.12 1.202
18 La HNB 1AliBu2H 1000 24 92 78.9 12.0 9.1 123.91 62.67 1.122
19 Y* TB 10AliBu3 1000 1 100 91.6 2.7 5.7 34.3 67.7 2.49
20 Y* HNB 10AliBu3 1000 0.5 100 87.2 8.6 4.2 75.3 68.1 1.49


The polyisoprenes obtained by using 2–3 combined with AliBu3 are characterized by rather broad molecular weight distribution (2.25–9.46), whereas application of AliBu2H allowed to decrease polydispersities considerably and to reach values 1.21–1.58. This observation can originate from higher efficiency of AliBu2H as a chain transfer agent21 and different geometry of the real catalytic species. Interestingly the La based systems are less influenced by the nature of alkyl aluminium component, with both AliBu3 and AliBu2H the PDI values remain in the region 1.69–2.96. At the same time the catalytic systems containing AliBu2H perform polymerizations with lower cis-1,4 selectivity. Thus the PIPs obtained with AliBu3 contain 45.2–88.7% cis-1,4 units, while the catalytic tests performed with AliBu2H afforded the polymers with contents of cis-1,4 units 38.6–72.3%. Interestingly the drop of cis-1,4 selectivity was less pronounced for Y-containing systems, whereas in the case of Nd the loss of cis-1,4 selectivity was ∼30%. The effect of [Ln]/[AliBu2H] molar ratio on molecular weight and microstructure of the obtained PIPs was also explored. A series of polymerizations initiated by the systems 2–4/AliBu2H/borate ([Ln]/[AliBu2H]/[borate] = 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1; borate = [Ph3C][B(C6F5)4], [PhNHMe2][B(C6F5)4]) containing one equivalent of AliBu2H was carried out (Table 1, entries 5, 6, 11, 12, 17, 18). It should be noted that decrease of [Ln]/[AliBu2H] molar ratio only in the case of Y results in the noticeable drop of catalytic activity while for the systems based on Nd and La compounds their efficiencies remain comparable. Also it was found that the ternary systems, containing one equivalent of AliBu2H display higher cis-1,4 selectivity compared to those with 10 equivalents of organoaluminium component. The influence of the [Ln]/[AliBu2H] ratio was particularly pronounced for the systems based on the lanthanum complex 4 (compare; 38.6 vs. 70.6% and 39.7 vs. 78.9%).

Comparison of the results of isoprene polymerization tests initiated by the systems based on bis(amido) complex 2 and related bis(alkyl) complex coordinated by the same amidinate ligand {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Y(CH2SiMe3)2 (THF) (5)10 revealed dramatic influence of the “leaving group” in Ln compound on both activity and selectivity (Table 1, entries 19 and 20). Thus complex 5 in combination with [PhNHMe2][B(C6F5)4] provides total conversion of monomer into polymer in 0.5 h affording a polymer with contents of 87.2% of cis-1,4 units and rather narrow polydispersity (1.49). Application of [Ph3C][B(C6F5)4] somewhat slows down the reaction, the total conversion requires already 1 h, however higher selectivity was demonstrated. So one can conclude that catalytic systems containing bis(alkyl) yttrium complexes provide better efficiency and cis-1,4 selectivity in isoprene polymerization and allow for formation of PIPs with rather narrow polydispersities.

The nature of real active species that forms in the ternary catalyst systems for isoprene polymerization have attracted a great deal of attention and to date the reactions of their components are explored in details.9a,c,e,24a Thus Luo and co-workers have shown that the reactions of related mono(amidinate) rare-earth-metal bis(silylamide) complexes [PhC(N-2,6-R2C6H3)2]Ln[N(SiHMe2)2]2(THF)y (R = Me, iPr; Ln = Sc, Y; y = 0, 1) with 1 equiv. of [Ph3C][B(C6F5)4] in toluene afford cationic amidinate rare-earth-metal amide complexes [{PhC(N-2,6-R2C6H3)2}LnNSiHMe2}{SiMe2N(SiHMe2)2}(THF)n][B(C6F5)4]. The treatment of [PhC(N-2,6-R2C6H3)2]Ln[N(SiHMe2)2]2(THF)y with excess AlMe3 resulted the heterometallic Ln/Al methyl complexes [PhC(N-2,6-R2C6H3)2]Ln[(μ-Me)2AlMe2]2.9c,d The cationic complex [Nd{N(SiMe3)2}2(THF)2][B(C6F5)4] was prepared by reaction of Nd[N(SiMe3)2]3 and [HNMe2Ph][B(C6F5)4].9a Recently we reported that the reaction of LLn[N(SiHMe2)2]3(thf) (L = amidine-amidopyrpdinate ligand) with [Et3NH]+[BPh4] affords cationic monoamido complex.24a The reaction of the cyclopentadienyl rare-earth metal bis(silylamide) complexes Cp*Ln[N(SiHMe2)2]2 with AlMe3 produced the heterometallic Ln/Al methyl complexes Cp*Ln[(μ-Me)2AlMe2]2, which were formed via the amide–alkyl exchange.9i Taking into account that neither binary systems 2–4/AliBu3, 2–4/AliBu2H nor 2–4 activated by equimolar amounts of [Ph3C][B(C6F5)4] or [PhNHMe2][B(C6F5)4] do not perform catalytic activity one can conclude that the most likely active species in our ternary catalyst system should be the Ln/Al heterobimetallic cationic alkyl complexes.

We have demonstrated a dramatic effect of “leaving group” in Ln compound on activity, selectivity as well as on molecular weight distribution of the obtained PIPs. Alkyl yttrium complex {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Y(CH2SiMe3)2(THF)10 provides better reaction rate, cis-1,4 selectivity and narrower polydispersities compared to its amide congener.

Rac-Lactide polymerization initiated by complexes 2–4

There is considerable interest in the controlled ring-opening polymerization (ROP) of lactides (LAs) by well-defined metal initiators because of the biodegradable and biocompatible nature of polylactides (PLAs) and their potential commercial applications.22 We investigated the catalytic activity of new compounds 2–4 containing potentially initiating nucleophilic amido group in the ROP of racemic lactide (rac-LA) (Scheme 2). Representative results are summarized in Table 2. Complexes 2–4 proved to be efficient catalysts of ROP of rac-LA under mild conditions, allowing for the total conversion of 100–500 equiv. of monomer (toluene, 25 °C, [LA] = 1.0 mol L−1) into polymer in 30 min. PLAs obtained in all catalytic tests have atactic structures as evidenced by NMR analyses (Pr = 0.50–0.55). This indicates a poorly effective chain-end control and suggests that polymerization of lactide mediated by 2–4 takes place in a non sterically congested environment.
image file: c5ra27960g-s2.tif
Scheme 2 Catalytic ring opening polymerization of rac-lactide.
Table 2 Polymerization of rac-lactide with complexes 2–4a
No. Cat. [LA]/[Ln]/ROH Time (min) Conv., % Mn × 10−3 Mncalc × 10−3 Mw/Mn
a General conditions: toluene, [LA] = 1.0 mol L−1, T = 25 °C. Reaction time was not necessarily optimized. Conversion of monomer M, as determined by 1H NMR spectroscopy. Experimental (corrected; see Experimental section) Mn and Mw/Mn values determined by GPC in THF vs. polystyrene standards. Mn value calculated by assuming two polymer chain per metal center with the relationship: (144 × conversion × [M]/2[Ln]).
1 Y 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0 30 100 37.7 7.2 1.47
2 Nd 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0 30 100 28.2 7.2 1.48
3 La 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0 30 100 32.0 7.2 1.34
4 Y 500[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0 30 94 126.5 33.8 1.63
5 Nd 500[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0 30 100 127.7 (82%), 1.9 (8%) 36.0 1.67, 1.48
6 La 500[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]0 30 100 231.8 (74%), 1.7 (26%) 36.0 1.90, 2.33
7 Y 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 30 92 22.5 6.6 1.61
8 Nd 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 30 100 22.4 7.2 1.44
9 La 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 30 100 24.4 7.2 1.40
10 Y 500[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 30 85 97.2 (92%), 1.5 (8%) 30.6 1.66, 1.44
11 Nd 500[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 30 100 77.5 (84%), 1.2 (16%) 36.0 1.53, 1.73
12 La 500[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 30 100 101.2 (81%), 1.2 (19%) 36.0 1.57, 1.83
13 Y 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2 30 88 25.2 6.3 1.73
14 Nd 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2 30 100 12.2 7.2 1.53
15 La 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2 30 94 10.8 6.7 1.43
16 Y 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3 30 94 7.4 6.8 1.32
17 Nd 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3 30 100 9.7 7.2 1.85
18 La 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3 30 100 9.9 7.2 1.91
19 Y 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]4 30 90 4.3 6.5 1.29
20 Nd 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]4 30 87 4.5 6.3 1.49
21 La 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]4 30 86 5.0 6.2 1.74
22 Y 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]10 30 89 1.56 6.4 1.51
23 Nd 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]10 30 89 2.4 6.4 1.42
24 La 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]10 30 88 2.4 6.3 1.53
25 Y 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20 30 91 0.7 6.6 1.64
26 Nd 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20 30 81 1.3 5.8 1.35
27 La 100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]20 30 88 1.9 6.3 1.61


Catalytic tests did not reveal noticeable effect of the nature of the metal center on activities of the complexes: at the ratio [LA]/[Ln] = 100 all three initiators in 0.5 h allowed to reach total conversion. However when the [LA]/[Ln] ratio was increased to 500 the Y compound was slightly less active (0.5 h, 94% vs. 100%; entries 4–6). All the PLAs obtained with complexes 2–4 at the ratio [LA]/[Ln] = 100 showed monomodal GPC traces with rather narrow polydispersity values Mw/Mn = 1.34–1.48. At ratio [LA]/[Ln] = 500 the polymers synthesized with Nd and La complexes displayed bimodal molecular weights distribution, while for the polymer obtained with Y compound it remained monomodal (entries 4–6). Increase of [LA]/[Ln] ratio lead to slight broadening of PDI. It is noteworthy that corrected experimental number-average molecular weights of obtained PLAs as determined by GPC noticeably exceed the values calculated from the initial monomer-to-lanthanide ratio, conversion and assuming that both amido groups are active initiators (Table 2). It should be noted that this general trend has been often observed for the ROP of lactide initiated by rare-earth amido complexes due to slow initiation.22b,23

It was associated with lower nucleophilicity of the amido group N(SiMe3)2, acting on the initiation stage as compared to the nucleophilic alkoxide group –OR (R – the growing polymer chain) resulted from opening of the lactide and the insertion of the resulting fragments into M–(N(SiMe3)2) bond [For examples of ROP of LAs initiated by rare-earth amido complexes see ref. 12, 14 and 24]. In some cases the GPC curves display bimodal character. That most likely originates from slow initiation and trans esterification reaction.

Possibility to provide control over polymerization reaction as well as to achieve immortal polymerization with these systems, that is, to generate several PLA chains per metal center by introducing several equivalents of a chain transfer agent, was explored with complexes 2–4 in the presence of iso-propanol which acts as a chain transfer agent (Table 2, entries 10–27). Addition of 1 equivalent of iso-propanol to the reaction mixture at both [LA]/[Ln] ratios results in some decrease of Mn but does not influence strongly PDI. Bimodal character of the GPC curves at [LA]/[Ln] = 500 ratio was retained. When 2 equivalents of iso-propanol were added ([LA]/[Ln] = 100) the experimental Mn still noticeably surpass the calculated ones. In the presence of 3–20 equivalents of iso-propanol polymerization rates and conversions remain high. The Mn values expectedly decrease with increase of concentration of iso-propanol however no match of experimental and calculated values was observed. The polydispersities fall into the region 1.30–1.90.

Interestingly two related yttrium bis(amide) complexes {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Y(N(SiMe3)2)2 and [p-MeC6H4NPh2]Y(N(SiMe3)2)2(THF)12 coordinated by amidinate ligands despite of different steric bulk and denticity of the latter perform very similar catalytic activity in ROP of lactide. The fact that in both cases atactic PLAs were obtained indicates that introduction of 2-[Ph2P(O)] group into the amidinate scaffold is insufficient for creation of sterically congested environment and providing control of polymer microstructure by the means of chain-end control.

Conclusions

The amine elimination reactions of equimolar amounts of Ln[N(SiMe3)2]3 (Ln = Y, Nd, La) and amidine 2-[Ph2P(O)]PhNHC(tBu)[double bond, length as m-dash]N(2,6-Me2C6H3) allowed for the synthesis of a series of the bis(amido) amidinate complexes {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Ln[N(SiMe3)2]2 in good yields. These complexes proved to be solvent free and featuring intramolecular coordination of Ph2P(O) group to Ln3+ ions in the solid state.

We demonstrated that amido complexes 2–4 activated by borate ([Ph3C][B(C6F5)4], [PhNHMe2][B(C6F5)4]) and alkylaluminium compounds (AliBu3, AliBu2H) similarly to the related bis(alkyl) species {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Ln(CH2SiMe3)2(THF)n (Ln = Y, n = 1, Ln = Er, n = 1, Ln = Lu, n = 0)10 enable isoprene polymerization. Catalytic activity of the ternary systems was found to be dramatically influenced by the nature of the metal atom, borate and alkyl aluminium components. Interestingly lanthanum complex 4 containing the largest La3+ ion displayed the lowest activity. Moreover comparison of the results of polymerization tests obtained with the systems based on complex 2 and related bis(alkyl) complex coordinated by the same amidinate ligand {2-[Ph2P(O)]C6H4NC(tBu)N(2,6-Me2C6H3)}Y(CH2SiMe3)2(THF)10 revealed a dramatic effect of “leaving group” in Ln compound on activity, selectivity as well as on molecular weight distribution of the obtained PIPs. Alkyl yttrium complex provides better reaction rate, cis-1,4 selectivity and narrower polydispersities.

Complexes 2–4 initiate ROP of rac-LA under mild conditions providing high reaction rates and formation of PLAs having atactic structures and rather narrow polydispersities. These initiators appear to be well suited for achieving immortal polymerization of lactide through the introduction of large amounts of isopropanol as a chain transfer agent, thus enabling the conversion of large amounts of monomer and the production of numerous macromolecular chains per metal initiator.

Experimental section

General conditions

All experiments were performed in evacuated tubes by using standard Schlenk techniques, with rigorous exclusion of traces of moisture and air. After being dried over KOH, THF was purified by distillation from sodium/benzophenone ketyl; hexane and toluene were dried by distillation from sodium and benzophenone ketyl prior to use. C6D6 was dried with sodium and condensed in vacuum into NMR tubes prior to use. CDCl3 was used without additional purification. Ln[N(SiMe3)2]3, (Ln = Y, Nd, La)11 and [2-Ph2P(O)]C6H4NHC(tBu)(2,6-Me2C6H3)10 were prepared according to literature procedures. Rac-Lactide was recrystallized two times from dry THF. All other commercially available chemicals were used after the appropriate purifications. NMR spectra were recorded with Bruker Avance DRX-400 and Bruker DRX-200 spectrometers in CDCl3, C6D6 at 20 °C, unless otherwise stated. Chemical shifts for 1H and 13C NMR spectra were referenced internally to the residual solvent resonances and are reported relative to TMS. IR spectra were recorded as Nujol mulls with a “Bruker-Vertex 70” instrument. Lanthanide metal analyses were carried out by complexometric titration.25 The C, H, N elemental analyses were performed in the microanalytical laboratory of the G. A. Razuvaev Institute of Organometallic Chemistry. GPC was carried out using chromatograph “Knauer Smartline” with Phenogel Phenomenex Columns 5u (300 × 7.8 mm) 104, 105 and Security Guard Phenogel Column with RI and UV detectors (254 nm). Mobile phase was THF, flow rate was 2 mL min−1. Columns was calibrated by Phenomenex Medium and High Molecular Weight Polystyrene Standard Kits with peak Mw from 2700 to 2[thin space (1/6-em)]570[thin space (1/6-em)]000 Da. The number average molecular masses (Mn) and polydispersity index (Mw/Mn) of the polymers were calculated with reference to an universal calibration versus polystyrene standards. Mn values of PLAs were corrected with a Mark–Houwink factor of 0.58 to account for the difference in hydrodynamic volumes between polystyrene and polylactide. Microstructures of PLAs were measured by homodecoupling 1H NMR spectroscopy at 25 °C in CDCl3 on a Bruker Avance DRX-400 spectroscopy instrument.

General procedure for synthesis of {2-[Ph2P(O)]C6H4NC(tBu)(2,6-Me2C6H3)}Y(N(SiMe3)2)2 (2)

A solution of Y[N(SiMe3)2]3 (0.342 g, 0.60 mmol) in 20 mL of THF was added to solution of 2-[Ph2P(O)]PhNHC(tBu)(2,6-Me2C6H3) (1) (0.288 g, 0.60 mmol) in 20 mL of THF. The reaction mixture was stirred at room temperature for 24 h and the volatiles were removed in vacuum. Recrystallization of the solid residue from THF/hexane mixture afforded yellow transparent crystals in 60% yield (0.320 g). Anal. calcd for C43H68N4OPSi4Y (889.25): C, 58.08; H, 7.71; N, 6.30; Y, 9.99. Found: C, 57.81; H, 7.38; N, 6.07; Y, 10.10. 1H NMR (400 MHz, C6D6, 25 °C): δ 0.39 (s, 36H, N(Si(CH3)3)2), 0.65 (s, 9H, C(CH3)3), 2.48 (s, 6H, C6H3(CH3)2), 6.49–6.54 (m, 1H, Aryl-H), 6.85–7.11 (m, 12H, Aryl-H), 7.61, 7.77 (br. s, both 2H, Aryl-H). 13C{1H} NMR (100.62 MHz, C6D6, 25 °C): δ 5.6 (N(Si(CH3)3)2), 19.9, 22.8 (C6H3(CH3)2), 30.0 (C(CH3)3), 41.7 (d, 3JCY = 2.3 Hz, C(CH3)3), 118.1 (d, JCP = 108.1 Hz), 119.2 (d, JCP = 14.3 Hz), 123.7, 126.9 (d, JCP = 7.5 Hz), 128.5 (d, JCP = 13.0 Hz), 132.6, 133.3 (d, JCP = 1.8 Hz), 133.6 (d, JCP = 13.3 Hz), 147.1, 156.1 (d, JCP = 3.2 Hz), (Aryl-C), 176.2 (d, 2JCY = 1.8 Hz) (NCN). 31P NMR (161.9 MHz, C6D6, 25 °C): δ 41.2 (d, 2JPY = 6.4 Hz). IR (Nujol, KBr; ν (cm−1)): 1686 (m), 1556 (w), 1410 (s), 1243 (s), 1216 (w), 1179 (m), 1166 (w), 1586 (m), 1065 (m), 1029 (w), 958 (s), 882 (m), 863 (w), 833 (s), 778 (w), 758 (w), 746 (w), 707 (w), 694 (m), 664 (m), 607 (m).

General procedure for synthesis of {2-[Ph2P(O)]C6H4NC(tBu)(2,6-Me2C6H3)}Nd(N(SiMe3)2)2 (3)

Analogous synthetic procedure was used. 1 (0.207 g, 0.43 mmol), Nd(N(SiMe3)2)3 (0.240 g, 0.43 mmol). Green crystals of 3 were isolated in 61% yield (0.247 g). Anal. calcd for C43H68N4NdOPSi4 (944.58): C, 54.68; H, 7.25; N, 5.93; Nd, 15.27. Found: C, 54.39; H, 7.04; N, 5.77; Nd, 15.51. IR (Nujol, KBr; ν (cm−1)): 1588 (m), 1554 (w), 1407 (w), 1244 (s), 1213 (w), 1175 (m), 1162 (w), 1126 (s), 1088 (m), 1066 (m), 1030 (w), 976 (s), 936 (m), 878 (m), 858 (w), 832 (s), 769 (w), 754 (w), 705 (m), 693 (m), 661 (m), 609 (m).

General procedure for synthesis of {2-[Ph2P(O)]C6H4NC(tBu)(2,6-Me2C6H3)}La(N(SiMe3)2)2 (4)

Analogous synthetic procedure was used. 1 (0.257 g, 0.53 mmol), La(N(SiMe3)2)3 (0.332 g, 0.53 mmol). Yellow crystals of 4 were isolated in 72% yield (0.358 g). Anal. calcd for C43H68LaN4OPSi4 (939.25): C, 54.99; H, 7.29; N, 5.97; La, 14.79. Found: C, 54.64; H, 7.17; N, 6.11; La, 14.84. 1H NMR (400 MHz, C6D6, 25 °C): δ 0.37 (s, 36H, NSi((CH3)3)2), 0.74 (s, 9H, C(CH3)3), 2.43, 2.46 (s, both 3H, C6H3(CH3)2), 6.43 (td, JHH = 7.1, JHH = 3.2 Hz, 1H, Aryl-H), 6.76–7.11 (m, 12H, Aryl-H), 7.64–7.69 (m, 2H, Aryl-H), 7.73–7.82 (m, 2H, Aryl-H). 13C{1H} NMR (100.62 MHz, C6D6, 25 °C): δ 4.5, 4.8 (NSi((CH3)3)2), 20.1, 22.1 (C6H3(CH3)2), 30.0 (C(CH3)3), 42.8 (C(CH3)3), 117.3 (d, JCP = 109.8 Hz), 117.2 (d, JCP = 14.4 Hz), 123.4, 124.9 (d, JCP = 7.5 Hz), 128.2 (d, JCP = 15.7 Hz), 128.4, 128.5, 128.5, 128.6, 132.2, 132.3, 132.3, 132.4, 133.3 (d, JCP = 1.8 Hz), 133.9 (d, JCP = 13.2 Hz), 144.7, 155.6 (d, JCP = 3.5 Hz) (Aryl-C), 177.1 (NCN). 31P NMR (161.9 MHz, C6D6, 25 °C): δ 39.7. IR (Nujol, KBr; ν (cm−1)): 1675 (m), 1584 (s), 1573 (s), 1549 (m), 1419 (s), 1276 (w), 1251 (m), 1214 (m), 1169 (m), 1161 (m), 1139 (m), 1126 (w), 1120 (w), 1094 (s), 1069 (s), 1038 (s), 1023 (w), 998 (w), 978 (s), 933 (s), 871 (s), 836 (m), 822 (m), 764 (s), 747 (s), 735 (s), 692 (s), 665 (m), 625 (w), 601 (w).

General procedure for rac-lactide polymerization

In a typical experiment (Table 2, entry 1), in a glovebox, a Schlenk flask was charged with a solution of 2 (8.9 mg, 10 μmol) in toluene (1.0 mL). To this solution, rac-lactide (0.144 g, 1.0 mmol, 100 equiv.) was added rapidly. The mixture was immediately stirred with a magnetic stir bar at 20 °C for 30 min. After an aliquot of the crude material was sampled by pipette for determining monomer conversion by 1H NMR, the reaction was quenched by adding of 1.0 mL of a 10% H2O solution in THF, and the polymer was precipitated from CH2Cl2/pentane (ca. 2[thin space (1/6-em)]:[thin space (1/6-em)]100 mL) five times. The polymer was dried in vacuo to a constant weight.

General procedure for isoprene polymerization

A typical polymerization procedure was carried out as following. Under a nitrogen atmosphere and room temperature borate (0.01 mmol) was added to a solution of (2–4) (0.01 mmol) in toluene (1 mL) and 10 equivalents of AliBu3 (0.1 mmol). After which to the reaction mixture isoprene (1000 equiv., 1 mL) was added. Polymerization was initiated and carried out for 1–24 hours. The reaction mixture was quenched by the addition of ethanol, and then poured into a large amount of ethanol to precipitate the polymer, which was dried under vacuum at 60 °C to a constant weight. Then, 1,4- 3,4-regioselectivity was determined by 1H and 13C{1H} NMR spectroscopy. GPC of polyisoprenes was performed in THF at 20 °C. The number average molecular masses (Mn) and polydispersity indexes (Mw/Mn) of the polymers were calculated with reference to a universal calibration against polystyrene standards.

X-ray crystallography

The X-ray data for 2–4 were collected on Bruker Smart Apex (2) and Agilent Xcalibur (EOS) (3, 4) diffractometers (graphite-monochromated, MoKα radiation, ω-scan technique, λ = 0.71073 Å, T = 100(2) K). SADABS26 and ABSPACK27 were used to perform area-detector scaling and absorption corrections. The structures were solved by direct methods and were refined on F2 using SHELX.28 All non-hydrogen atoms were found from Fourier syntheses of electron density and were refined anisotropically. All hydrogen atoms were placed in geometrically idealized positions and treated as riding with Uiso(H) = 1.2Ueq (Uiso(H) = 1.5Ueq for the hydrogen atoms in CH3 groups) of their parent atoms. The CCDC files 1439443 (2), 1439444 (3), 1439445 (4) contain the supplementary crystallographic data for this paper.

Acknowledgements

This work was financially supported by the Russian Science Foundation (Project 14-13-00742).

Notes and references

  1. (a) M. F. Lappert, A. V. Protchenko, P. P. Power and A. L. Seeber, Amides of the Group 3 and Lanthanide Metals, in Metal Amide Chemistry, John Wiley & Sons Ltd, Chichester, UK, 2008 Search PubMed; (b) Lanthanide amides: R. Anwander, Top. Curr. Chem., 1996, 179, 33–112 CrossRef CAS.
  2. (a) D. C. Bradley, Chem. Rev., 1989, 89, 1317 CrossRef CAS; (b) L. G. Hubert-Pfalzgraf, New J. Chem., 1987, 11, 663 CAS; (c) L. G. Hubert-Pfalzgraf, Appl. Organomet. Chem., 1992, 6, 627 CrossRef CAS.
  3. (a) Y. Liang and R. Anwander, Dalton Trans., 2013, 42, 12521 RSC; (b) S. Hong and T. J. Marks, Adv. Chem. Res., 2004, 37, 673 CrossRef CAS PubMed; (c) T. E. Müller, K. C. Hultzsch, M. Yus, F. Foubelo and M. Tada, Chem. Rev., 2008, 108, 3795 CrossRef PubMed; (d) K. C. Hultzsch, Adv. Synth. Catal., 2005, 347, 367 CrossRef CAS; (e) M. R. Douglass, C. L. Stern and T. J. Marks, J. Am. Chem. Soc., 2001, 123, 10221 CrossRef CAS PubMed; (f) I. V. Basalov, S. C. Roşca, D. M. Lyubov, A. N. Selikhov, G. K. Fukin, Y. Sarazin, J. F. Carpentier and A. A. Trifonov, Inorg. Chem., 2014, 53, 1654 CrossRef CAS PubMed; (g) I. V. Basalov, A. N. Dorcet, G. K. Fukin, Y. Sarazin, J. F. Carpentier and A. A. Trifonov, Chem.–Eur. J., 2015, 21, 6033 CrossRef CAS PubMed; (h) V. Rodriguez-Ruiz, R. Carlino, S. Bezzenine-Lafolée, R. Gil, D. Prim, E. Schulz and J. Hannedouche, Dalton Trans., 2015, 44, 12029 RSC; (i) C. Zeng, D. Yuan, B. Zhao and Y. Yao, Org. Lett., 2015, 17, 2242 CrossRef CAS PubMed; (j) M. Rastätter, A. Zulys and P. W. Roesky, Chem.–Eur. J., 2007, 13, 3606 CrossRef PubMed; (k) C. Wang, L. Huang, M. Lu, B. Zhao, Y. Wang, Y. Zhang, Q. Shen and Y. Yao, RSC Adv., 2015, 5, 9476 RSC; (l) H. Cheng, Y. Xiao, C. Lu, B. Zhao, Y. Wang and Y. Yao, New J. Chem., 2015, 39, 7667 RSC; (m) Z. Li, M. Xue, H. Yao, H. Sun, Y. Zhang and Q. Shen, J. Organomet. Chem., 2012, 713, 27 CrossRef CAS.
  4. T. J. Woodman, Y. Sarazin, G. Fink, K. Hauschild and M. Bochmann, Macromolecules, 2005, 38, 3060 CrossRef CAS.
  5. Y. Luo, X. Feng, Y. Wang, S. Fan, J. Chen, Y. Lei and H. Liang, Organometallics, 2011, 30, 3270 CrossRef CAS.
  6. L. Zhou, Y. Yao and Q. Shen, J. Appl. Polym. Sci., 2009, 114, 2403 CrossRef CAS.
  7. (a) H. X. Li, Z. G. Ren, Y. Zhang, Z. H. Zhang, J. P. Lang and Q. Shen, J. Am. Chem. Soc., 2005, 127, 1122 CrossRef CAS PubMed; (b) H. Ma and J. Okuda, Macromolecules, 2005, 38, 2665 CrossRef CAS; (c) H. E. Dyer, S. Huijser, N. Susperregui, F. Bonnet, A. D. Schwarz, R. Duchateau, L. Maron and P. Mountford, Organometallics, 2010, 29, 3602 CrossRef CAS; (d) R. H. Platel, L. M. Hodgson and C. K. Williams, Polymer, 2008, 48, 11 CAS.
  8. S. P. Nolan, D. Stern and T. J. Marks, J. Am. Chem. Soc., 1989, 111, 7844 CrossRef CAS.
  9. (a) V. Monteil, R. Spitz and C. Boisson, Polym. Int., 2004, 53, 576 CrossRef CAS; (b) O. Tardif and S. Kaita, Dalton Trans., 2008, 19, 2531 RSC; (c) Y. Luo, Y. Lei, S. Fan, Y. Wang and J. Chen, Dalton Trans., 2013, 42, 4040 RSC; (d) Y. Luo, S. Fan, J. Yang, J. Fang and P. Xu, Dalton Trans., 2011, 40, 3053 RSC; (e) F. Chen, S. Fan, Y. Wang, J. Chen and Y. Luo, Organometallics, 2012, 31, 3730 CrossRef CAS; (f) V. Monteil, R. Spitz and C. Boisson, Polym. Int., 2004, 53, 576 CrossRef CAS; (g) C. Boisson, F. Barbotin and R. Spitz, Macromol. Chem. Phys., 1999, 200, 1163 CrossRef CAS; (h) N. Martins, F. Bonnet and M. Visseaux, Polymer, 2014, 55, 5013 CrossRef CAS; (i) R. Anwander, M. G. Klimpel, H. M. Dietrich, D. J. Shorokhov and W. Scherer, Chem. Commun., 2003, 1008 RSC.
  10. A. O. Tolpygin, T. A. Glukhova, A. V. Cherkasov, G. K. Fukin, D. V. Aleksanyan, D. Cui and A. A. Trifonov, Dalton Trans., 2015, 44, 16465 RSC.
  11. D. C. Bradley, J. S. Ghotra and F. A. Hart, J. Chem. Soc., Dalton Trans., 1973, 10, 1021 RSC.
  12. K. B. Aubrecht, K. Chang, M. A. Hillmyer and W. B. Tolman, J. Polym. Sci., Part A: Polym. Chem., 2001, 39, 284 CrossRef CAS.
  13. Y. Wang, Y. Luo, J. Chen, H. Xue and H. Liang, New J. Chem., 2012, 36, 933 RSC.
  14. Y. Luo, P. Xu, Y. Lei, Y. Zhang and Y. Wang, Inorg. Chim. Acta, 2010, 363, 3597 CrossRef CAS.
  15. K. Kincaid, C. P. Gerlach, G. R. Giesbrecht, J. R. Hagadorn, G. D. Whitener, A. Shafir and J. Arnold, Organometallics, 1999, 18, 5360 CrossRef CAS.
  16. F. Wang, Y. Wei, S. Wang, X. Zhu, S. Zhou, G. Yang, X. Gu, G. Zhang and X. Mu, Organometallics, 2015, 34, 86 CrossRef CAS.
  17. R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751 CrossRef.
  18. J. F. Corbey, J. H. Farnaby, J. E. Bates, J. W. Ziller, F. Furche and W. J. Evans, Inorg. Chem., 2012, 51, 7867 CrossRef CAS PubMed.
  19. H. C. Aspinall, S. R. Moore and A. K. Smith, J. Chem. Soc., Dalton Trans., 1992, 1, 153 RSC.
  20. (a) M. Zimmermann, K. W. Törnroos and R. Anwander, Angew. Chem., Int. Ed., 2008, 47, 775 CrossRef CAS PubMed; (b) D. Li, S. Li, D. Cui and X. Zhang, Organometallics, 2010, 29, 2186 CrossRef CAS; (c) K. Lv and D. Cui, Organometallics, 2010, 29, 2987 CrossRef CAS; (d) L. Wang, D. Cui, Z. Hou, W. Li and Y. Li, Organometallics, 2011, 30, 760 CrossRef CAS; (e) D. Robert, E. Abinet, T. P. Spaniol and J. Okuda, Chem.–Eur. J., 2009, 15, 11937 CrossRef CAS PubMed; (f) E. Le Roux, Y. Liang, K. W. Törnroos, F. Nief and R. Anwander, Organometallics, 2012, 31, 6526 CrossRef CAS; (g) S. Li, D. Cui, D. Li and Z. Hou, Organometallics, 2009, 28, 4814 CrossRef CAS; (h) G. Du, Y. Wei, L. Ai, Y. Chen, Q. Xu, X. Liu, S. Zhang, Z. Hou and X. Li, Organometallics, 2011, 30, 160 CrossRef CAS.
  21. L. Guo, X. Zhu, S. Zhou, X. Mu, Y. Wei, S. Wang, Z. Feng, G. Zhanga and B. Deng, Dalton Trans., 2014, 43, 6842 RSC.
  22. (a) C. K. Ha and J. A. Gardella Jr, Chem. Rev., 2005, 105, 4205 CrossRef CAS PubMed; (b) O. D. Cabaret, B.-M. Vaca and D. Bourissou, Chem. Rev., 2004, 104, 6147 CrossRef PubMed; (c) R. E. Drumright, P. R. Gruber and D. E. Henton, Adv. Mater., 2000, 12, 1841 CrossRef CAS; (d) M. H. Chisholm and Z. Zhou, J. Mater. Chem., 2004, 14, 3081 RSC.
  23. (a) M. J. Stanford and A. P. Dove, Chem. Soc. Rev., 2010, 39, 486 RSC; (b) C. M. Thomas, Chem. Soc. Rev., 2010, 39, 165 RSC; (c) M. V. Yakovenko, N. Y. Udilova, T. A. Glukhova, A. V. Cherkasov, G. K. Fukin and A. A. Trifonov, New J. Chem., 2015, 39, 1083 RSC.
  24. (a) V. Rad'kov, T. Roisnel, A. Trifonov, J.-F. Carpentier and E. Kirillov, Eur. J. Inorg. Chem., 2014, 25, 4168 CrossRef; (b) S. Sun, K. Nie, Y. Tan, B. Zhao, Y. Zhang, Q. Shen and Y. Yao, Dalton Trans., 2013, 42, 2870 RSC; (c) J. Wang, T. Cai, Y. Yao, Y. Zhang and Q. Shen, Dalton Trans., 2007, 45, 5275 RSC; (d) Q. Wang, F. Zhang, H. Song and G. Zi, J. Organomet. Chem., 2011, 696, 2186 CrossRef CAS; (e) Y. Wang, Y. Lei, S. Chi and Y. Luo, Dalton Trans., 2013, 42, 1862 RSC; (f) F. Zhang, J. Zhang, H. Song and G. Zi, Inorg. Chem. Commun., 2011, 14, 72 CrossRef CAS.
  25. S. J. Lyle and M. M. Rahman, Talanta, 1953, 10, 1177 CrossRef.
  26. Bruker, SADABS, Bruker AXS Inc, Madison, Wisconsin, USA, 2001 Search PubMed.
  27. Agilent, CrysAlis PRO, Agilent Technologies Ltd, Yarnton, Oxfordshire, England, 2014 Search PubMed.
  28. G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Adv., 2015, 71, 3–8 CrossRef PubMed.

Footnote

Electronic supplementary information (ESI) available. CCDC 1439443 (2), 1439444 (3) and 1439445 (4). For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5ra27960g

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.