Next Article in Journal
Development of a Hybrid Bioinorganic Nanobiocatalyst: Remarkable Impact of the Immobilization Conditions on Activity and Stability of β-Galactosidase
Next Article in Special Issue
Free-Standing, Flexible Nanofeatured Polymeric Films Prepared by Spin-Coating and Anodic Polymerization as Electrodes for Supercapacitors
Previous Article in Journal
Influence of Freezing and Different Drying Methods on Volatile Profiles of Strawberry and Analysis of Volatile Compounds of Strawberry Commercial Jams
Previous Article in Special Issue
Additive Effects of Lithium Salts with Various Anionic Species in Poly (Methyl Methacrylate)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Side-Arm Assisted Anilido-Imine Based Rare-Earth Metal Complexes for Isoprene Stereoselective Polymerization

1
State Key Laboratory of Polymer Physics and Chemistry, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, China
2
Department of Polymer Science and Engineering, University of Science and Technology of China, Hefei 230026, China
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(14), 4154; https://doi.org/10.3390/molecules26144154
Submission received: 8 June 2021 / Revised: 30 June 2021 / Accepted: 6 July 2021 / Published: 8 July 2021

Abstract

:
Anilido-imine ligands o-C6H4(NHAr1)(CH=NAr2), in which Ar1 is 2,6-diisopropylbenzyl group and Ar2 contains fluorine (HL1) or methoxyl (HL2) group on ortho-position of phenyl substituent, were synthesized for constructing rare-earth metals based complexes of 1a1c (HL1 based Sc, Lu, Y) and 2a2c (HL2 based Sc, Lu, Y). Based on their NMR spectra and X-ray single-crystal structures, the side-arm group of -F and -OMe is identified to chelate to the corresponding central metal. The twisted angles between two planes formed by chelated heteroatoms (N, N, F for HL1 and N, N, O for HL2) are observed, in which the largest dihedral angle (53.3°) for HL1-Y and the smallest dihedral angle (44.32°) for HL2-Sc are detected. After being activated by AliBu3 and [Ph3C][B(C6F5)4], these catalysts showed great activity for isoprene polymerization. Bearing the same ligand HL1, smaller scandium based complex 1a and middle size of lutetium based 1b provided lower cis-1,4-selectivity (57.3% and 64.2%), larger yttrium complex 1c displayed high cis-1,4-selectivity (84%). Chelating by crowded HL2, small size of scandium complex 2a provided impressive trans-1,4-selectivity (93.0%), middle lutetium based 2b displayed non-selectivity and larger yttrium complex 2c showed clear cis-1,4-selectivity (83.3%). Moreover, 2a/AliBu3 system showed the quasi-living chain transfer capability.

Graphical Abstract

1. Introduction

According to the different regio- and stereo-selective polymerization method, polyisoprene is commonly divided into cis-1,4-polyisoprene, trans-1,4-polyisoprene and 3,4-polyisoprene. Among them, cis-1,4-polyisoprene has similar chemical composition, stereoselectivity and mechanical properties to natural rubber. It is widely used in tires, conveyor belts, adhesives, sports equipment, etc. In recent years, due to the shortage of natural resources and the aggravation of environmental pollution, it has become a trend to produce high performance tires with low rolling resistance, high wet skid resistance and low fuel consumption [1]. It is reported that the high-performance rubber can be obtained by mixing a small amount of trans-1,4-polyisoprene with cis-1,4-polyisoprene or natural rubber [2]. In addition, trans-1,4-selective polyisoprene has unique applications in medical materials and shape memory materials, etc. For example, trans-1,4-polyisoprene is an ideal material for making medical splints, orthopedic components and prosthetics.
So far, a large number of rare-earth metal catalysts for the cis-1,4-selective and 3,4-selective polyisoprene have been reported [3,4,5,6,7,8,9,10,11,12], such as aryldiimine (NCN)-ligated rare-earth metal dichlorides (cis-1,4-selectivity up to 98.8%) [13]; bis(carbene)phenyl (CCC) rare-earth metal dibromides (cis-1,4-selectivity up to 99.6%) [14]; bis(phosphino)-carbazolide (PNP)-chelated rare-earth metal complexes (cis-1,4-selectivity up to 99%) [15,16], etc. For 3,4-selective polyisoprene, there are amidino N-heterocyclic carbene ligated lutetium complex (3,4-selectivity 98.7%) [17], iminophosphonamido (NPN) ligated rare-earth metal bis(alkyl)s (3,4-selectivity 99.4%) and [Me2Si(C5Me4)(μ-PCy)YCH2SiMe3]2 (Cy = cyclohexyl) (3,4-selectivity 100%), etc. [18,19].
Comparatively, there are only a few reports on the trans-1,4-selective catalysts [2,20,21,22,23]. For the trans-1,4 polymerization of isoprene, the half-sandwich rare-earth metal complexes (CpQ)Ln(AlMe4)2 (CpQ = 2,3,4,5-tetramethyl-1-(8-quinolyl)cyclopentadienyl; Ln = Y, La) (trans-1,4-stereoselectivity >93%) [24], (Flut-Bu)La(AlMe4)2 (trans-1,4-stereoselectivity >85%) reported by Anwander and the chiral mononuclear dialkyl pincer complexes [(S,S)-BOPA]Ln(CH2SiMe3)2 (BOPA = (S,S)-bis(oxazolinylphenyl)amido; Ln = Sc, Lu) discovered by Xiaofang Li (trans-1,4-stereoselectivity > 99%) are efficient catalysts [25,26]. Interestingly, efficient modulating the stereoselective polymerization of isoprene has also been achieved by changing the size of metal center and/or rational adjusting the steric/electronic effect of chelating ligand or co-catalyst. For example, Shojiro Kaita et.al reported high cis-1,4/trans-1,4 selective polymerization of butadiene controlled by synergetic influence of metal size and type of AlR3 [27].
Previously we reported β-iminophosphonamine ligated complexes can realize the switchable stereoselectivity of polyisoprene from 3,4-selectivity of 94% to trans-1,4-selectivity of 95% by changing the metal size from Lu3+ (84.8 pm) to La3+ (106.1 pm) [2]. In addition to the type of metal and co-catalyst, the strategy of using the flexible side arm on ligand to adjust the electronic/steric properties of complex and thus finely change the capability of catalysts has also attracted much attention. As introduced by Yong Tang et.al, the additional “side-arm” groups near the catalytic center to regulate the electronic properties and spatial shape of the metal center, thus affecting the coordination of the monomer, and achieving the purpose of regulating the activity and catalytic properties of the catalyst [28,29]. However, there are few of report that fine modulate the stereoselectivity of catalyst by “side-arm strategy”.
Herein, we devised an aniline-imine ligated rare-earth metal complexes and introduced flexible heteroatom groups near the metal center to further control the polymerization process of isoprene. The skeleton of aniline-imine ligands might well be interpreted as a hybridization of classic salicylaldiminate and β-diketiminate ligands, which are widely applied to construct catalysts for olefin polymerization [30,31]. Through combining the “side-arm strategy” and suitable metal size influence, we have successfully realized the switchable stereoselectivity of polyisoprene from cis-1,4 to trans-1,4 selectivity. Furthermore, the efficient chain transfer polymerization for high trans-1,4 polymerization system was fine developed.

2. Results and Discussion

2.1. Synthesis and Characterization of Rare-Earth-Metal Complexes

The deprotonation of anilido-imine ligands with 1 equiv of rare-earth metal tris(alkyl)s afforded a series of rare-earth metal–bis(alkyl) complexes in high yields (Scheme 1). All the complexes are soluble in polar solvent like toluene and THF, but insoluble in unpolar hexane. The solid-state structures of complexes 1a, 1c, 2a and 2c were confirmed by X-ray diffraction measurements (Figure 1, Figure 2, Figure 3 and Figure 4). And the crystallographic data were summarized in Table S1 (see Supplementary Materials). Except for the coordinated bidentate ligands, the fluorine group and methoxy substituent on the phenyl group also coordinate to the corresponding central metal. The Y–N (amido) distances [2.313(5) Å for 1c, 2.310(2) Å for 2c] are shorter than the Y–N (imine) distances [2.451(5) Å for 1c, 2.476(2) Å for 2c], respectively. Except for 2a, additional one THF molecule was involved in their crystal structures, which implies that 2a owns the most crowded coordinating environment. A twisted angle between two planes formed by chelated heteroatoms (N, N, F for HL1 and N, N, O for HL2) were observed. Interestingly, the highest cis-1,4-selectivity of 1c owns the largest dihedral angle (53.3°) among them, and the highest trans-1,4-selectivity of 2a shows the smallest dihedral angle (44.32°), which are depicted below in detail. For complex 2a, O, N1 and N2 all coordinate with the central metal scandium forming a chelating complex. The O–Sc bond length is 2.2387(13), Sc–N1 bond length is 2.1415(15), Sc–N2 bond length is 2.2540(15), which are close to the bond lengths between scandium and coordinated nitrogen atoms in the chelating catalyst [(S,S)-BOPA]Sc(CH2SiMe3)2 (BOPA = (S,S)-bis(oxazolinylphenyl)amido) reported by Xiaofang Li. Therefore, their catalytic behaviors are similar, and they both show high trans-1,4 selectivity for isoprene polymerization [26].
The 1H NMR spectrum of complex 1a (Figure S3) was indicative of the formation of bis(alkyl) species and the resonance at 0.25 ppm is attributed to the methylene protons of scandium alkyl species. Compared with many other rare earth metal alkyl species, the resonance shifts down due to the stronger Lewis acidity of the Sc3+ ion [2]. Correspondingly, the methylene protons of Lu–CH2SiMe3 and Y–CH2SiMe3 in 1b and 1c appear at −0.75 and −0.51 ppm at higher field, respectively (Figures S5 and S7).

2.2. Polymerization of Isoprene

Isoprene polymerization was investigated in detail and the representative data are summarized in Table 1. The isoprene polymerization was initiated immediately as the complex was activated by [Ph3C][B(C6F5)4] and aluminum alkyls. The monomer was completely consumed within 0.5 h at room temperature (Table 1, entries 1–9).
Bearing the ligand L1, the scandium complex 1a had medium cis-1,4-selectivity (57.3%), while yttrium complex 1c and 2c provided predominant cis-1,4-selectivity (84.0%, 83.3%) (Table 1, entries 1, 3, 7). This may be due to the fact that the larger yttrium provided a more open space, which is beneficial to the η4-cis-coordinating mode of monomer. Bearing the same metal center, the complex 2b had medium trans-1,4-selectivity (58.3%), while the complex 1b had medium cis-1,4-selectivity (64.2%) (Table 1, entries 2, 6). This may be attributed to the small steric hindrance of fluorine group facilitating η4-coordination. On the other hand, the electronic absorption of fluorine group increased the Lewis acidity of Lu3+ ion, which could increase the chance of η4-coordination [18]. Surprisingly, scandium complex 2a displayed high trans-1,4-selectivity (93.0%), and exhibited a promising catalytic activity even at a low temperature (−30 °C, 98% yield in 4 h) with an increased trans-1,4-selectivity (97.0%) (Table 1, entries 4–5). Compared with 1a, the side arm of ligand L2 adopted methoxy group with larger steric hindrance, making the ligand have a bulky space shielding on the metal center. The more planar structure of complex 2a together with bulky space may only allow η2-trans-coordination mode of isoprene, on the other hand, the consecutive insertion of trans-monomer into the allyl–metal bond of active species with syn-prenyl moiety and anti-syn isomerization prior to monomer insertion lead to trans-1,4 unit [32,33,34,35,36,37].
It is well known that AliBu3 is not only used as a cocatalyst for the activation of catalyst precursors, but also as a chain transfer agent to regulate the molecular weight. In general, the molecular weight distribution increases with the amount of AliBu3 due to the deactivation of active sites by excessive AliBu3 [38,39,40]. In contrast, using the 2a/AliBu3/[Ph3C][B(C6F5)4] system, the molecular weight of PIP decreased inversely with the increase of [Al]/[2a] ratio, meanwhile the molecular weight distribution were almost the same, showing the quasi-living chain transfer mode. For example, with the increase of [Al]/[2a] ratios from 2:1 to 40:1, the molecular weight of polyisoprene decreased from 4.58 × 104 to 0.85 × 104 g/mol, together with a slightly lower selectivity was observed (Table 1, entries 4, 8–11).

3. Materials and Methods

3.1. General Information

All manipulations were performed under nitrogen atmosphere using standard high-vacuum Schlenk techniques or in a glovebox (Braun, Germany). All solvents were purified with a SPS system (Braun, Germany). The NMR data of the organometallic samples were obtained on a Bruker AV500 spectrometer (Switzerland) in chloroform-d or benzene-d6 at room temperature. The molecular weight and molecular weight distribution of the polyisoprene were measured with a HLC-8420 GPC (Tosoh corporation, Shunan, Yamaguchi, Japan) at 40 °C using THF as eluent (the flow rate was 0.35 mL/min) against polystyrene standards. Differential scanning calorimetry analyses were carried out on a Q 100 DSC (METTLER TOLEDO, Switzerland) from TA instrument under a nitrogen atmosphere. Any thermal history difference in the polymers was eliminated by first heating the specimen to above 80 °C, then cooling to −80 °C at 10 °C/min, and finally recording the second DSC scan from −80 to 80 °C at 10 °C/min. Isoprene was dried over CaH2 with stirring for 48 h and distilled under vacuum before use. [Ph3C][B(C6F5)4] was synthesized following the literature [41].

3.2. Synthesis

3.2.1. Synthesis of o-C6H4NH(C6H4-F-o)(CH=NC6H3-i-Pr2-2,6) (L1)

Anilido-imine ligands L1–L2 were prepared according to the literature procedures [30,31]. Taking the synthesis method of ligand L1 as an example: o-fluorobenzaldehyde (6.2 g, 50 mmol), 2,6-diisopropylaniline (8.9 g, 50 mmol) and MgSO4 were mixed in hexane and stirred for 1 h. The mixture was filtered, and the solvent was removed to obtain the yellow solid o-C6H4F(CHNC6H3-i-Pr2-2,6), which was recrystallized in hexane to obtain pure product. A solution of n-BuLi (6 mL, 14.7 mmol) in hexane was added into hexane solution of o-fluoroaniline (1.4 mL, 14 mmol) at −78 °C, and the white lithium salt LiNHAr precipitated immediately, then the mixture was warmed to room temperature and stirred for 2 h, then it was added into a solution of o-C6H4F(CHNC6H3-i-Pr2-2,6) (4 g, 14 mmol) in THF at 25 °C. After stirring 5 h, the reaction was terminated by adding water, extracted with hexane and the solvent was removed to obtain the yellow solid crude product. The pure product was obtained by recrystallization in hexane (3 g, 60%). 1H NMR (500 MHz, CDCl3, 7.26 ppm, 25 °C): δ 1.19 (d, 12H, CH(CH3)2), 3.07 (m, 2H, CH(CH3)2), 6.88 (t, 1H, Ph-H), 7.03–7.4 (m, 9H, Ph-H), 7.56 (t, 1H, Ph-H), 8.33 (s, 1H, CH = NAr), 11.12 (s, 1H, NH). Analytical Calculated (Anal. Calcd) for C25H27FN2 (%): C, 80.18; H, 7.27; N, 7.48. Found: C, 80.45; H, 7.01; N, 7.73.

3.2.2. Synthesis of o-C6H4NH(C6H4-OMe-o)(CH=NC6H3-i-Pr2-2,6) (L2)

The preparation method of ligand L2 is similar to that of ligand L1. 1H NMR (500 MHz, CDCl3, 7.26 ppm, 25 °C): δ 1.19 (d, 12H, CH(CH3)2), 3.10 (m, 2H, CH(CH3)2), 3.77 (s, 3H, OCH3), 6.82 (t, 1H, Ph-H), 6.92–6.98 (m, 2H, Ph-H), 7.04 (t, 1H, Ph-H), 7.12–7.31 (m, 4H, Ph-H), 7.37 (d, 1H, Ph-H), 7.44 (d, 1H, Ph-H), 7.56 (d, 1H, Ph-H), 8.3 (s, 1H, CH = NAr), 10.95 (s, 1H, NH). Anal. Calcd for C26H30N2O (%): C, 80.79; H, 7.82; N, 7.25. Found: C, 80.55; H, 8.08; N, 7.51.

3.2.3. Synthesis of L1Sc(CH2SiMe3)2(THF) (1a)

The hexane solution (4.0 mL) of Sc(CH2SiMe3)3(THF)2 (0.22 g, 0.5 mmol) was added dropwise to the ligand L1 solution (0.187 g, 0.5 mmol in 4 mL hexane) at 0 °C. The mixture was stirred for 1 h and then cooling to −30 °C for 1 day afforded crystalline solids, which dried in vacuo to give orange solids of 1a (0.23 g, 69%). Single crystals suitable for X-ray analysis were obtained from hexane at −30 °C. 1H NMR (500 MHz, C6D6, 7.16 ppm, 25 °C): δ 0.11 (s, 18H, CH2SiMe3), 0.25 (s, 4H, CH2SiMe3), 0.99 (d, 6H, CH(CH3)2), 1.24 (m, 4H, THF), 1.29 (d, 6H, CH(CH3)2), 3.31 (br s, 2H, CH(CH3)2), 3.59 (m, 4H, THF), 6.43–7.39 (m, 11H, Ph-H), 8.00 (s, 1H, CH = NAr). 13C NMR (125 MHz, C6D6, 128 ppm, 25 °C): 3.22, 25.6, 29.0, 45.9, 68.5, 114.9, 118.7, 119.5, 120.5, 121.8, 124.4, 126.5, 135.2, 136.2, 138.4, 141.5, 146.6, 151.8, 157.3, 159.1, 170.4. Anal. Calcd for C40H64FN2OScSi2 (%): C, 67.75; H, 9.10; N, 3.95. Found: C, 67.53; H, 8.90; N, 4.20.

3.2.4. Synthesis of L1Lu(CH2SiMe3)2(THF) (1b)

The preparation method of complex 1b is similar to that of complex 1a (67%). 1H NMR (500 MHz, C6D6, 7.16 ppm, 25 °C): δ −0.77 (s, 4H, CH2SiMe3), 0.08 (s, 18H, CH2SiMe3), 0.92 (d, 6H, CH(CH3)2), 1.23 (m, 4H, THF), 1.34 (d, 6H, CH(CH3)2), 3.21 (m, 2H, CH(CH3)2), 3.85 (m, 4H, THF), 6.36–7.48 (m, 11H, Ph-H), 7.96 (s, 1H, CH = NAr). 13C NMR (125 MHz, C6D6, 128 ppm, 25 °C): 4.28, 25.0, 25.4, 28.9, 41.2, 70.3, 114.4, 115, 116, 120.5, 124.1, 125.6, 126.1, 127.2, 133.4, 134.8, 140.6, 141.5, 147.9, 149.2, 156.2, 158.1, 168.6. Anal. Calcd for C40H64FLuN2OSi2 (%): C, 57.26; H, 7.69; N, 3.34. Found: C, 57.56; H, 7.90; N, 3.08.

3.2.5. Synthesis of L1Y(CH2SiMe3)2(THF) (1c)

The preparation method of complex 1c is similar to that of complex 1a (65%). Single crystal suitable for X-ray analysis was obtained from hexane at −30 °C. 1H NMR (500 MHz, C6D6, 7.16 ppm, 25 °C): δ −0.51 (s, 4H, CH2SiMe3), 0.11 (s, 18H, CH2SiMe3), 0.94 (d, 6H, CH(CH3)2), 1.22 (m, 4H, THF), 1.36 (d, 6H, CH(CH3)2), 3.21 (m, 2H, CH(CH3)2), 3.84 (m, 4H, THF), 6.38–7.49 (m, 11H, Ph-H), 7.98 (s, 1H, CH = NAr). 13C NMR (125 MHz, C6D6, 128 ppm, 25 °C): 4.13, 25.33, 28.95, 35.30, 70.30, 114.22, 115.45, 120.66, 124.15, 125.90, 126.07, 127.13, 133.29, 134.88, 140.45, 141.33, 147.73, 148.53, 156.32, 158.12, 168.18. Anal. Calcd for C40H64FN2OSi2Y (%): C, 63.80; H, 8.57; N, 3.72. Found: C, 63.59; H, 8.81; N, 3.95.

3.2.6. Synthesis of L2Sc(CH2SiMe3)2 (2a)

The hexane solution (4 mL) of Sc(CH2SiMe3)3(THF)2 (0.22 g, 0.5 mmol) was added dropwise to the ligand L2 solution (0.193 g, 0.5 mmol in 4 mL hexane) at 0 ℃. The mixture was stirred for 1 h and then cooling to −30 °C for 1 day afforded crystalline solids, which dried in vacuo to give orange solids of 2a (0.22 g, 73%). 1H NMR (500 MHz, C6D6, 7.16 ppm, 25 °C): δ −0.05 (s, 4H, CH2SiMe3), 0.00 (s, 18H, CH2SiMe3), 1.01–1.33 (m, 12H, CH(CH3)2), 3.37 (m, 2H, CH(CH3)2), 3.90 (s, 3H, OCH3), 6.48–7.45 (m, 11H, Ph-H), 8.07 (s, 1H, CH = NAr). 13C NMR (125 MHz, C6D6, 128 ppm, 25 °C): 3.25, 26.07, 58.54, 111.28, 117.43, 118.51, 120.46, 120.99, 121.24, 124.14, 124.44, 135.15, 136.36, 139.96, 141.81, 147.12, 149.88, 153.11, 170.63. Anal. Calcd for C37H60N2OScSi2 (%): C, 68.37; H, 9.30; N, 4.31. Found: C, 68.72; H, 9.04; N, 4.58.

3.2.7. Synthesis of L2Lu(CH2SiMe3)2(THF) (2b)

The preparation method of complex 2b is similar to that of complex 2a (70%). 1H NMR (500 MHz, C6D6, 7.16 ppm, 25 °C): δ −0.59, −0.84 (4H, CH2SiMe3), 0.00 (s, 18H, CH2SiMe3), 0.96–1.24 (m, 12H, CH(CH3)2), 1.37 (m, 4H, THF), 3.22 (m, 2H, CH(CH3)2), 3.55 (m, 4H, THF), 3.75 (s, 3H, OCH3), 6.44–7.45 (m, 11H, Ph-H), 7.94 (s, 1H, CH=NAr). 13C NMR (125 MHz, C6D6, 128 ppm, 25 °C): 3.88, 25.71, 29.06, 58.01, 67.99, 111.60, 117.28, 120.12, 120.45, 120.62, 121.47, 124.43, 135.06, 136.57, 140.90, 141.64, 145.97, 149.98, 154.17, 171.05. Anal. Calcd for C41H67LuN2O2Si2 (%): C, 57.86; H, 7.93; N, 3.29. Found: C, 57.63; H, 8.14; N, 3.53.

3.2.8. Synthesis of L2Y(CH2SiMe3)2(THF) (2c)

The preparation method of complex 2c is similar to that of complex 2a (74%). Single crystals suitable for X-ray analysis were obtained from hexane at −30 °C. 1H NMR (500 MHz, C6D6, 7.16 ppm, 25 °C): δ −0.46, −0.68 (4H, CH2SiMe3), 0.06 (s, 18H, CH2SiMe3), 1.00–1.33 (m, 12H, CH(CH3)2), 1.37 (m, 4H, THF), 3.19 (m, 2H, CH(CH3)2), 3.61 (m, 4H, THF), 3.78 (s, 3H, OCH3), 6.52–7.63 (m, 11H, Ph-H), 7.94 (s, 1H, CH = NAr). 13C NMR (125 MHz, C6D6, 128 ppm, 25 °C): 4.21, 25.38, 29.00, 31.97, 34.57, 34.88, 57.66, 69.37, 110.62, 116.03, 117.53, 119.03, 123.94, 124.19, 124.50, 127.20, 128.35, 133.77, 135.46, 140.82, 142.63, 147.58, 150.02, 151.29, 168.84. Anal. Calcd for C41H67N2O2Si2Y (%): C, 64.37; H, 8.83; N, 3.66. Found: C, 64.65; H, 8.59; N, 3.96.

3.3. Polymerization of Isoprene

In a glovebox, a toluene solution (3 mL) of complex 1a (6.6 mg, 10 μmol), 1 equiv. of [Ph3C][B(C6F5)4] (9.2 mg, 10 μmol) and 0.1 mmol of AliBu3 were added to a 10 mL flask. The toluene solution of isoprene (1 mL of toluene, 10 mmol, 0.68 g) was added to the catalytic system and stirred for 0.5 h. Then, the polymerization was terminated by injecting methanol. The mixture was poured into methanol containing an appropriate amount of BHT to precipitate the products. The obtained polymer was collected by filtration, washed with methanol, and dried to constant weight in vacuum at 35 °C. Then the conversion can be calculated from the mass of the polymer (Table 1, Entry 1).

3.4. X-ray Crystallographic Studies

Data collections of Single crystals of complexes were performed at −100 °C on a Bruker SMART APEX diffractometer with a CCD area detector, using graphite-monochromated Mo Kα radiation (λ = 0.71073 Å). The SMART program package was used to determine the unit-cell parameters. The absorption correction was applied using SADABS. The structures were solved by direct methods and refined on F2 by full-matrix least-squares techniques with anisotropic thermal parameters for non-hydrogen atoms. Hydrogen atoms were placed at calculated positions and were included in the structure calculation without further refinement of the parameters. All calculations were carried out using the Olex 2 program. Molecular structures were generated using the ORTEP program.

4. Conclusions

In summary, a series of rare-earth metal complexes (Ln = Sc, Lu, Y) bearing anilido-imine ligand were successfully synthesized and defined, which were activated with [Ph3C][B(C6F5)4] and aluminum alkyls to initiate polymerization of isoprene. The choice of central metal size and the introduction of heteroatom coordination groups with different steric hindrance on the side arm of the ligand are the key to adjust the stereoselectivity of polyisoprene. Therefore, by adjusting these two factors, the stereoselectivity of polyisoprene was successfully switched from cis-1,4 (84%) to trans-1,4 (93%). In addition, the quasi-living chain transfer polymerization for high trans-1,4 polyisoprene system was developed.

Supplementary Materials

The following are available online: Figure S1: 1H NMR spectrum (500 MHz, CDCl3, 25 °C) of ligand HL1, Figure S2: 1H NMR spectrum (500 MHz, CDCl3, 25 °C) of ligand HL2, Figure S3: 1H NMR spectrum (500 MHz, C6D6, 25 °C) of complex 1a, Figure S4: 13C NMR spectrum (125 MHz, C6D6, 25 °C) of complex 1a, Figure S5: 1H NMR spectrum (500 MHz, C6D6, 25 °C) of complex 1b, Figure S6: 13C NMR spectrum (125 MHz, C6D6, 25 °C) of complex 1b, Figure S7: 1H NMR spectrum (500 MHz, C6D6, 25 °C) of complex 1c, Figure S8: 13C NMR spectrum (125 MHz, C6D6, 25 °C) of complex 1c, Figure S9: 1H NMR spectrum (500 MHz, C6D6, 25 °C) of complex 2a, Figure S10: 13C NMR spectrum (125 MHz, C6D6, 25 °C) of complex 2a, Figure S11: 1H NMR spectrum (500 MHz, C6D6, 25 °C) of complex 2b, Figure S12: 13C NMR spectrum (125 MHz, C6D6, 25 °C) of complex 2b, Figure S13: 1H NMR spectrum (500 MHz, C6D6, 25 °C) of complex 2c, Figure S14: 13C NMR spectrum (125 MHz, C6D6, 25 °C) of complex 2c, Figure S15: 1H NMR spectrum (500 MHz, CDCl3, 25 °C) of cis-1,4 PIP (Table 1, entry 3), Figure S16: 13C NMR spectrum (125 MHz, CDCl3, 25 °C) of cis-1,4 PIP (Table 1, entry 3), Figure S17: 1H NMR spectrum (500 MHz, CDCl3, 25 °C) of trans-1,4 PIP (Table 1, entry 4), Figure S18: 13C NMR spectrum (125 MHz, CDCl3, 25 °C) of trans-1,4 PIP (Table 1, entry 4), Table S1: Crystal data and structure refinement for complexes 1a, 1c, 2a, 2c.

Author Contributions

Conceptualization, X.L. and D.C.; methodology, X.L.; software, X.L.; validation, X.L.; formal analysis, Y.W.; investigation, Y.W.; resources, D.C.; data curation, Y.W.; writing—original draft preparation, Y.W.; writing—review and editing, X.L. and D.C.; supervision, D.C.; project administration, X.L.; funding acquisition, D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by NSFC for project Nos. 21634007 and 21674108, Science and Technology Development Project of Jilin Province (No. 20190201067JC).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data is available in the main text, in the supplementary materials, or on reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rong, W.; Liu, D.; Zuo, H.; Pan, Y.; Jian, Z.; Li, S.; Cui, D. Rare-Earth-Metal Complexes Bearing Phosphazene Ancillary Ligands: Structures and Catalysis toward Highly Trans-1,4-Selective (Co)Polymerizations of Conjugated Dienes. Organometallics 2012, 32, 1166–1175. [Google Scholar] [CrossRef]
  2. Liu, B.; Sun, G.; Li, S.; Liu, D.; Cui, D. Isoprene Polymerization with Iminophosphonamide Rare-Earth-Metal Alkyl Complexes: Influence of Metal Size on the Regio- and Stereoselectivity. Organometallics 2015, 34, 4063–4068. [Google Scholar] [CrossRef]
  3. Evans, W.J.; Giarikos, D.G. Chloride Effects in Lanthanide Carboxylate Based Isoprene Polymerization. Macromolecules 2004, 37, 5130–5132. [Google Scholar] [CrossRef]
  4. Evans, W.J.; Giarikos, D.G.; Ziller, J.W. Lanthanide Carboxylate Precursors for Diene Polymerization Catalysis: Syntheses, Structures, and Reactivity with Et2AlCl. Organometallics 2001, 20, 5751–5758. [Google Scholar] [CrossRef]
  5. Fischbach, A.; Klimpel, M.G.; Widenmeyer, M.; Herdtweck, E.; Scherer, W.; Anwander, R. Stereospecific Polymerization of Isoprene with Molecular and MCM-48-Grafted Lanthanide(III) Tetraalkylaluminates. Angew. Chem. Int. Ed. 2004, 43, 2234–2239. [Google Scholar] [CrossRef] [PubMed]
  6. Ajellal, N.; Furlan, L.; Thomas, C.M.; Jr, O.L.C.; Carpentier, J.-F. Mixed Aluminum-Magnesium-Rare Earth Allyl Catalysts for Controlled Isoprene Polymerization: Modulation of Stereocontrol. Macromol. Rapid Commun. 2006, 27, 338–343. [Google Scholar] [CrossRef]
  7. Zhang, L.; Suzuki, T.; Luo, Y.; Nishiura, M.; Hou, Z. Cationic Alkyl Rare-Earth Metal Complexes Bearing an Ancillary Bis(phosphinophenyl)amido Ligand: A Catalytic System for Livingcis-1,4-Polymerization and Copolymerization of Isoprene and Butadiene. Angew. Chem. Int. Ed. 2007, 46, 1909–1913. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, L.; Nishiura, M.; Yuki, M.; Luo, Y.; Hou, Z. Isoprene Polymerization with Yttrium Amidinate Catalysts: Switching the Regio- and Stereoselectivity by Addition of AlMe3. Angew. Chem. Int. Ed. 2008, 47, 2642–2645. [Google Scholar] [CrossRef]
  9. Zhang, G.; Deng, B.; Wang, S.; Wei, Y.; Zhou, S.; Zhu, X.; Huang, Z.; Mu, X. Di and trinuclear rare-earth metal complexes supported by 3-amido appended indolyl ligands: Synthesis, characterization and catalytic activity towards isoprene 1,4-cis polymerization. Dalton Trans. 2016, 45, 15445–15456. [Google Scholar] [CrossRef]
  10. Pan, Y.; Li, W.; Wei, N.-N.; So, Y.-M.; Li, Y.; Jiang, K.; He, G. Anilido-oxazoline-ligated rare-earth metal complexes: Synthesis, characterization and highly cis-1,4-selective polymerization of isoprene. Dalton Trans. 2019, 48, 3583–3592. [Google Scholar] [CrossRef]
  11. Li, L.; Wu, C.; Liu, D.; Li, S.; Cui, D. Binuclear Rare-Earth-Metal Alkyl Complexes Ligated by Phenylene-Bridged β-Diketiminate Ligands: Synthesis, Characterization, and Catalysis toward Isoprene Polymerization. Organometallics 2013, 32, 3203–3209. [Google Scholar] [CrossRef]
  12. Jian, Z.; Tang, S.; Cui, N. Highly Regio- and Stereoselective Terpolymerization of Styrene, Isoprene and Butadiene with Lutetium-Based Coordination Catalyst. Macromolecules 2011, 44, 7675–7681. [Google Scholar] [CrossRef]
  13. Gao, W.; Cui, D. Highly cis-1,4 Selective Polymerization of Dienes with Homogeneous Ziegler–Natta Catalysts Based on NCN-Pincer Rare Earth Metal Dichloride Precursors. J. Am. Chem. Soc. 2008, 130, 4984–4991. [Google Scholar] [CrossRef]
  14. Lv, K.; Cui, D. CCC-Pincer Bis(carbene) Lanthanide Dibromides. Catalysis on Highly cis-1,4-Selective Polymerization of Isoprene and Active Species. Organometallics 2010, 29, 2987–2993. [Google Scholar] [CrossRef]
  15. Wang, L.; Cui, D.; Hou, Z.; Li, W.; Li, Y. Highly Cis-1,4-Selective Living Polymerization of 1,3-Conjugated Dienes and Copolymerization with ε-Caprolactone by Bis(phosphino)carbazolide Rare-Earth-Metal Complexes. Organometallics 2011, 30, 760–767. [Google Scholar] [CrossRef]
  16. Huang, J.; Liu, Z.; Cui, D.; Liu, X. Precisely Controlled Polymerization of Styrene and Conjugated Dienes by Group 3 Single-Site Catalysts. ChemCatChem 2017, 10, 42–61. [Google Scholar] [CrossRef]
  17. Yao, C.; Liu, D.; Li, P.; Wu, C.; Li, S.; Liu, B.; Cui, D. Highly 3,4-Selective Living Polymerization of Isoprene and Copolymerization with ε-Caprolactone by an Amidino N-Heterocyclic Carbene Ligated Lutetium Bis(alkyl) Complex. Organometallics 2014, 33, 684–691. [Google Scholar] [CrossRef]
  18. Li, S.; Cui, D.; Li, D.; Hou, Z. Highly 3,4-Selective Polymerization of Isoprene with NPN Ligand Stabilized Rare-Earth Metal Bis(alkyl)s. Structures and Performances. Organometallics 2009, 28, 4814–4822. [Google Scholar] [CrossRef]
  19. Zhang, L.; Luo, Y.; Hou, Z. Unprecedented Isospecific 3,4-Polymerization of Isoprene by Cationic Rare Earth Metal Alkyl Species Resulting from a Binuclear Precursor. J. Am. Chem. Soc. 2005, 127, 14562–14563. [Google Scholar] [CrossRef] [Green Version]
  20. Yang, Y.; Lv, K.; Wang, L.; Wang, Y.; Cui, D. Isoprene polymerization with aminopyridinato ligand supported rare-earth metal complexes. Switching of the regio- and stereoselectivity. Chem. Commun. 2010, 46, 6150–6152. [Google Scholar] [CrossRef]
  21. Liu, D.; Cui, D. Highly trans-1,4 selective (co-)polymerization of butadiene and isoprene with quinolyl anilido rare earth metal bis(alkyl) precursors. Dalton Trans. 2011, 40, 7755–7761. [Google Scholar] [CrossRef]
  22. Bonnet, F.; Jones, C.E.; Semlali, S.; Bria, M.; Visseaux, M.; Roussel, P.; Arnold, P.L. Tuning the catalytic properties of rare earth borohydrides for the polymerisation of isoprene. Dalton Trans. 2013, 42, 790–801. [Google Scholar] [CrossRef] [Green Version]
  23. Bonnet, F.; Visseaux, M.; Pereira, A.; Barbier-Baudry, D. Highlytrans-Stereospecific Isoprene Polymerization by Neodymium Borohydrido Catalystst. Macromolecules 2005, 38, 3162–3169. [Google Scholar] [CrossRef] [Green Version]
  24. Litlabø, R.; Enders, M.; Törnroos, K.W.; Anwander, R. Bis(tetramethylaluminate) Complexes of Yttrium and Lanthanum Supported by a Quinolyl-Substituted Cyclopentadienyl Ligand: Synthesis and Performance in Isoprene Polymerization. Organometallics 2010, 29, 2588–2595. [Google Scholar] [CrossRef]
  25. Diether, D.; Tyulyunov, K.; Maichle-Mössmer, C.; Anwander, R. Fluorenyl Half-Sandwich Bis(tetramethylaluminate) Complexes of the Rare-Earth Metals: Synthesis, Structure, and Isoprene Polymerization. Organometallics 2017, 36, 4649–4659. [Google Scholar] [CrossRef]
  26. Liu, H.; He, J.; Liu, Z.; Lin, Z.; Du, G.; Zhang, S.; Li, X. Quasi-Livingtrans-1,4-Polymerization of Isoprene by Cationic Rare Earth Metal Alkyl Species Bearing a Chiral (S,S)-Bis(oxazolinylphenyl)amido Ligand. Macromolecules 2013, 46, 3257–3265. [Google Scholar] [CrossRef]
  27. Kaita, S.; Yamanaka, M.; Horiuchi, A.C.; Wakatsuki, Y. Butadiene Polymerization Catalyzed by Lanthanide Metallocene–Alkylaluminum Complexes with Cocatalysts: Metal-Dependent Control of 1,4-Cis/Trans Stereoselectivity and Molecular Weight. Macromolecules 2006, 39, 1359–1363. [Google Scholar] [CrossRef]
  28. Yang, X.; Wang, Z.; Sun, X.; Tang, Y. Synthesis, characterization, and catalytic behaviours of β-carbonylenamine-derived [ONS]TiCl3 complexes in ethylene homo- and copolymerization. Dalton Trans. 2009, 41, 8945–8954. [Google Scholar] [CrossRef]
  29. Liao, S.; Sun, X.-L.; Tang, Y. Side Arm Strategy for Catalyst Design: Modifying Bisoxazolines for Remote Control of Enantioselection and Related. Accounts Chem. Res. 2014, 47, 2260–2272. [Google Scholar] [CrossRef] [PubMed]
  30. Xu, T.; An, H.; Gao, W.; Mu, Y. Chromium(III) Complexes with Chelating Anilido-Imine Ligands: Synthesis, Structures, and Catalytic Properties for Ethylene Polymerization. Eur. J. Inorg. Chem. 2010, 2010, 3360–3364. [Google Scholar] [CrossRef]
  31. Gao, H.; Guo, W.; Bao, F.; Gui, G.; Zhang, J.; Zhu, A.F.; Wu, Q. Synthesis, Molecular Structure, and Solution-Dependent Behavior of Nickel Complexes Chelating Anilido–Imine Donors and Their Catalytic Activity toward Olefin Polymerization. Organometallics 2004, 23, 6273–6280. [Google Scholar] [CrossRef]
  32. Li, S.; Miao, W.; Tang, T.; Dong, W.; Zhang, X.; Cui, D. New Rare Earth Metal Bis(alkyl)s Bearing an Iminophosphonamido Ligand. Synthesis and Catalysis toward Highly 3,4-Selective Polymerization of Isoprene. Organometallics 2008, 27, 718–725. [Google Scholar] [CrossRef]
  33. Kang, X.; Luo, Y.; Zhou, G.; Wang, X.; Yu, X.; Hou, Z.; Qu, J. Theoretical Mechanistic Studies on thetrans-1,4-Specific Polymerization of Isoprene Catalyzed by a Cationic La–Al Binuclear Complex. Macromolecules 2014, 47, 4596–4606. [Google Scholar] [CrossRef]
  34. Tobisch, S. Mechanistic insight into the selective trans-1,4-polymerization of butadiene by terpyridine–iron(II) complexes—A computational study. Can. J. Chem. 2009, 87, 1392–1405. [Google Scholar] [CrossRef]
  35. Porri, L.; Giarrusso, A.; Ricci, G. Recent views on the mechanism of diolefin polymerization with transition metal initiator systems. Prog. Polym. Sci. 1991, 16, 405–441. [Google Scholar] [CrossRef]
  36. Wang, D.; Li, S.; Liu, X.; Gao, W.; Cui, D. Thiophene-NPN Ligand Supported Rare-Earth Metal Bis(alkyl) Complexes. Synthesis and Catalysis toward Highly trans-1,4 Selective Polymerization of Butadiene. Organometallics 2008, 27, 6531–6538. [Google Scholar] [CrossRef]
  37. Wang, Z.; Liu, D.; Cui, D. Highly Selective Polymerization of 1,3-Conjugated Dienes by Rare Earth Organometallic Complexes. Acta Polym. Sin. 2015, 989–1009. [Google Scholar] [CrossRef]
  38. Jian, Z.; Cui, D.; Hou, Z.; Li, X. Living catalyzed-chain-growth polymerization and block copolymerization of isoprene by rare-earth metal allyl precursors bearing a constrained-geometry-conformation ligand. Chem. Commun. 2010, 46, 3022–3024. [Google Scholar] [CrossRef] [PubMed]
  39. Wu, C.; Liu, B.; Lin, F.; Wang, M.; Cui, D. cis -1,4-Selective Copolymerization of Ethylene and Butadiene: A Compromise between Two Mechanisms. Angew. Chem. Int. Ed. 2017, 56, 6975–6979. [Google Scholar] [CrossRef]
  40. Liu, B.; Wang, L.; Wu, C.; Cui, D. Sequence-controlled ethylene/styrene copolymerization catalyzed by scandium complexes. Polym. Chem. 2018, 10, 235–243. [Google Scholar] [CrossRef]
  41. Piers, W.E.; Chivers, T. Pentafluorophenylboranes: From obscurity to applications. Chem. Soc. Rev. 1997, 26, 345–354. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of complexes 1a1c and 2a2c.
Scheme 1. Synthesis of complexes 1a1c and 2a2c.
Molecules 26 04154 sch001
Figure 1. X-ray molecular structure of complex 1a with 35% probability thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (degree): Sc1–F1 = 2.493(3), Sc1–N1 = 2.156(4), Sc1–N2 = 2.331(4), Sc1–O1 = 2.210(3), Sc1–C1 = 2.227(4), Sc1–C2 = 2.234(4), Si1–C1–Sc1 = 129.3(2), Si2–C2–Sc1 = 135.2(2).
Figure 1. X-ray molecular structure of complex 1a with 35% probability thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (degree): Sc1–F1 = 2.493(3), Sc1–N1 = 2.156(4), Sc1–N2 = 2.331(4), Sc1–O1 = 2.210(3), Sc1–C1 = 2.227(4), Sc1–C2 = 2.234(4), Si1–C1–Sc1 = 129.3(2), Si2–C2–Sc1 = 135.2(2).
Molecules 26 04154 g001
Figure 2. X-ray molecular structure of complex 1c with 35% probability thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (degree): Y1–F1 = 2.639(4), Y1–N1 = 2.451(5), Y1–N2 = 2.313(5), Y1–O1 = 2.361(5), Y1–C1 = 2.381(7), Y1–C2 = 2.365(8), Si1–C1–Y1 = 135.2(4), Si2–C2–Y1 = 135.8(5).
Figure 2. X-ray molecular structure of complex 1c with 35% probability thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (degree): Y1–F1 = 2.639(4), Y1–N1 = 2.451(5), Y1–N2 = 2.313(5), Y1–O1 = 2.361(5), Y1–C1 = 2.381(7), Y1–C2 = 2.365(8), Si1–C1–Y1 = 135.2(4), Si2–C2–Y1 = 135.8(5).
Molecules 26 04154 g002
Figure 3. X-ray molecular structure of complex 2a with 35% probability thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (degree): Sc1–O1 = 2.2387(13), Sc1–N1 = 2.1415(15), Sc1–N2 = 2.2540(15), Sc1–C1 = 2.217(2), Sc1–C2 = 2.221(2), Si1–C1–Sc1 = 129.79(11), Si2–C2–Sc1 = 121.75(10).
Figure 3. X-ray molecular structure of complex 2a with 35% probability thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (degree): Sc1–O1 = 2.2387(13), Sc1–N1 = 2.1415(15), Sc1–N2 = 2.2540(15), Sc1–C1 = 2.217(2), Sc1–C2 = 2.221(2), Si1–C1–Sc1 = 129.79(11), Si2–C2–Sc1 = 121.75(10).
Molecules 26 04154 g003
Figure 4. X-ray molecular structure of complex 2c with 35% probability thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (degree): Y1–O1 = 2.4025(19), Y1–O2 = 2.5277(19), Y1–N1 = 2.476(2), Y1–N2 = 2.310(2), Y1–C1 = 2.389(4), Y1–C2 = 2.417(4), Si1–C1–Y1 = 131.1(2), Si2–C2–Y1 = 132.38(18).
Figure 4. X-ray molecular structure of complex 2c with 35% probability thermal ellipsoids. Hydrogen atoms are omitted for clarity. Selected bond lengths (Å) and angles (degree): Y1–O1 = 2.4025(19), Y1–O2 = 2.5277(19), Y1–N1 = 2.476(2), Y1–N2 = 2.310(2), Y1–C1 = 2.389(4), Y1–C2 = 2.417(4), Si1–C1–Y1 = 131.1(2), Si2–C2–Y1 = 132.38(18).
Molecules 26 04154 g004
Table 1. Polymerization of isoprene by using catalytic precursors under various conditions a.
Table 1. Polymerization of isoprene by using catalytic precursors under various conditions a.
EntryCatAliBu3/CatT (°C)Time (h)Conv. (%)Microstructure (%)
cis-1,4/trans-1,4/3,4
Mnb(104)PDI bTg/Tmc(°C)
11a10250.5>9957.3/25.6/17.14.031.62−55.9/--
21b10250.5>9964.2/21.7/14.15.831.64−60.3/--
31c10250.5>9984.0/6.7/9.36.241.12−59.9/--
42a10250.5>991.7/93.0/5.32.051.54−69.7/36.7
52a10−30498-/97.0/3.02.031.34−68.3/37.8
62b10250.5>9936.3/58.3/5.45.761.29−66.5/--
72c10250.5>9983.3/15.2/1.56.051.16−64.8/--
82a2250.5>991.4/94.1/4.54.581.51−69.6/36.5
92a5250.5>991.5/93.6/4.93.081.61−69.3/37.6
102a20250.5>991.8/92.8/5.41.341.56−69/36.4
112a40252>992.0/92.5/5.50.851.53−69.5/36.8
a General polymerization conditions: rare earth complex 10 μmol, [IP]/[Cat]/[Ph3C][B(C6F5)4] = 1000/1/1 (mol/mol), toluene 4 mL. b Determined by gel permeation chromatography (GPC) in THF at 40 °C against polystyrene standard. c Determined by differential scanning calorimetry (DSC).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, Y.; Liu, X.; Cui, D. Side-Arm Assisted Anilido-Imine Based Rare-Earth Metal Complexes for Isoprene Stereoselective Polymerization. Molecules 2021, 26, 4154. https://doi.org/10.3390/molecules26144154

AMA Style

Wu Y, Liu X, Cui D. Side-Arm Assisted Anilido-Imine Based Rare-Earth Metal Complexes for Isoprene Stereoselective Polymerization. Molecules. 2021; 26(14):4154. https://doi.org/10.3390/molecules26144154

Chicago/Turabian Style

Wu, Yi, Xinli Liu, and Dongmei Cui. 2021. "Side-Arm Assisted Anilido-Imine Based Rare-Earth Metal Complexes for Isoprene Stereoselective Polymerization" Molecules 26, no. 14: 4154. https://doi.org/10.3390/molecules26144154

Article Metrics

Back to TopTop