Evolution of aluminum aminophenolate complexes in the ring-opening polymerization of ε-caprolactone: electronic and amino-chelating effects

Prasanna Kumar Ganta a, Fei Huang a, Taoufik Ben Halima b, Rajiv Kamaraj a, Yu-Ting Chu ac, Hsi-Ching Tseng d, Shangwu Ding ac, Kuo-Hui Wu *e and Hsuan-Ying Chen *acfg
aDepartment of Medicinal and Applied Chemistry, Drug Development and Value Creation Research Center, Kaohsiung Medical University, Kaohsiung, Taiwan 80708, Republic of China. E-mail: hchen@kmu.edu.tw; Fax: +886-7-3125339; Tel: +886-7-3121101-2585
bDepartment of Chemistry & Biomolecular Sciences, University of Ottawa, Ottawa, Canada
cDepartment of Chemistry, National Sun Yat-Sen University, Kaohsiung, Taiwan 80424, Republic of China
dCollege of Science Instrumentation Center, National Taiwan University, Taipei, Taiwan 106319, Republic of China
eDepartment of Chemistry, National Central University, Taoyuan, Taiwan 32001, Republic of China
fDepartment of Medical Research, Kaohsiung Medical University Hospital, Kaohsiung 80708, Taiwan, Republic of China
gNational Pingtung University of Science and Technology, Pingtung, Taiwan 91201, Republic of China

Received 19th October 2024 , Accepted 28th November 2024

First published on 3rd December 2024


Abstract

A series of aluminum complexes bearing phenolate (O–Al and O2–Al), biphenolate (OO–Al type), aminophenolate (ON–Al), aminobiphenolate (ONO–Al), bis(phenolato)bis(amine) (NNOO–Al), and Salan (ONNO–Al) type ligands were synthesized. ε-Caprolactone (CL) polymerization using these aluminum complexes as catalysts was investigated. The overall polymerization rates of Al catalysts with different ligands were found to be in the following order (kobs values): ONBr–Al (0.124 min−1) ≥ OBr2–Al (0.121 min−1) > ONNOBr–Al (0.054 min−1) > NNOBr–Al (0.044 min−1) ≥ ONOBr–Al (0.043 min−1) > OBr–Al (0.033 min−1) > NNOOBr–Al (0.015 min−1) ≥ BuONNOBu–Al (0.001 min−1) = OOBr–Al (0.001 min−1). In addition, Al complexes with electron-donating substituents on ligands exhibited higher catalytic activity than those with bromo substituents. Density functional theory (DFT) calculations revealed that a dinuclear Al complex with two bridging methoxides had to rearrange to a phenolate bridged dinuclear Al complex with terminal methoxides. This is due to the low initiating ability of two bridging benzyl alkoxides. Combining the polymerization data and DFT results, it was concluded that the electron-donating substituents on the phenolate ring and chelating amino group enhance the electron density of the Al center. This may prevent the formation of a less active dinuclear Al complex with two bridging alkoxides (initiators) or facilitate its structural rearrangement. OOMe–Al has been established as a powerful candidate with a high polymerization rate and it exhibits well-controlled polymerization for synthesizing the mPEG-b-PCL copolymer.


Introduction

Poly-ε-caprolactone (PCL) is an advantageous biomaterial in various fields.1 This is mainly due to its biodegradability,2 biocompatibility,3 and permeability.4 Given the increasing demand for PCL, synthesizing PCL effectively remains a topic of concern. Compared to the traditional polycondensation,5 ring-opening polymerization using metal complexes6 as catalysts exhibited greater catalytic activity and controllability. Aluminum complexes7 proved to be effective catalysts for cyclic ester ROP combining both high conversion and ease of synthesis. Ligand design is closely related to the catalytic activity of Al catalysts due to the influence of ligands through inductive, steric, and chelating effects. The investigation of the inductive effect on the catalytic activity of cyclic ester polymerization by Al catalysts is illustrated in Fig. 1, and lots of Al complexes, including enolic Salen,8 Schiff base,9 8-quinolinolate,10 ketiminate,11 β-diketiminate,12 amidinate,13 β-quinolyl-enaminate,14 and phenoxy-thiother15 ligands, showed higher activity with electron-withdrawing groups. However, Tolman's group16 reported a series of five-coordinate Al complexes bearing bis(phenolato)bis(amine) ligands (NNOO type) and their application in CL polymerization (Fig. 2A), and the polymerization results revealed that Al complexes with electron-donating groups, such as tert-butyl and methoxy groups, on the aromatic position para to the phenoxide donor oxygen exhibited higher catalytic activity compared to Al complexes with electron-withdrawing groups, such as bromo substituents. The same catalytic behavior of five-coordinate Al complexes bearing bis(phenolato)bis(amine) ligands for rac-lactide (rac-LA) polymerization was also reported by Gibson's group17 (Fig. 2B). These results attracted the attention of numerous researchers because many electron-deficient Al complexes9c–e,10,12,13,18 with electron-withdrawing substituents of ligands exhibited higher catalytic activity for cyclic ester ROP compared to electron-rich Al complexes with ligands holding electron-donating substituents.
image file: d4dt02923b-f1.tif
Fig. 1 Classification of the catalytic activity of cyclic esters using Al complexes by the electronic tendency.

image file: d4dt02923b-f2.tif
Fig. 2 Five-coordinate Al complexes bearing bis(phenolato)bis(amine) ligands and their CL polymerization results reported by (A) Tolman and (B) Gibson groups.

According to the kinetic data,16b Tolman's group reported a plausible mechanism (Fig. 3) to explain why Al complexes with electron-donating groups exhibited higher catalytic activity for CL polymerization. Electron-rich Al complexes could create a vacant coordination site for CL coordination through the amine arm dissociation. However, the amine arm dissociation for electron-deficient Al complexes is hampered because of the enhanced Lewis acidity of the Al center by the electron-withdrawing groups of the ligands, and the sterically crowded Al center is not conducive to CL coordination to form a less stable octahedral intermediate.


image file: d4dt02923b-f3.tif
Fig. 3 Mechanism of five-coordinate Al complexes bearing different bis(phenolato)bis(amine) ligands for CL ROP.

Tolman's study proposed a mechanism and the related kinetic data to explain why electron-rich Al complexes bearing bis(phenolato)bis(amine) ligands showed higher catalytic activity and polymerization properties different from other Al catalysts.9c–e,10,12,13,18 Inspired by Tolman's explanation, we sought to determine whether electron-rich Al complexes still have higher catalytic activity than electron-deficient Al complexes with four-coordinate Al structures, such as ONO–Al, NO–Al, and OO–Al complexes (Fig. 4) for CL ROP. How does the amine arm affect the polymerization rate of the catalyst? What role did the amine arm play at this time? Herein, a series of ONO–Al complexes (Scheme 1A), NO–Al complexes (Scheme 1B), OO–Al complexes (Scheme 1D), O2–Al complexes (Scheme 1E), O–Al complexes (Scheme 1F), NNOO–Al complexes (Scheme 1G), and NNO–Al complexes (Scheme 1H) were synthesized, and their CL polymerization was investigated.


image file: d4dt02923b-f4.tif
Fig. 4 Three- and four-coordinate Al complexes were studied for CL polymerization.

image file: d4dt02923b-s1.tif
Scheme 1 Synthesis of ligands and associated Al complexes in this study.

In addition, ONNO–Salan Al complexes18c,h,19 were reported with cyclic ester polymerization, as shown in Fig. 5. Gibson's, Feijen's, and Hormnirun's results of rac-LA ROP using ONNO–Salan Al complexes as catalysts showed that the Al complex with chloro substituents on the Salan ligand exhibited higher catalytic activity compared to Al complexes with methyl substituents. However, the steric effects between the chloro and methyl groups on the aromatic position ortho to the phenoxide donor oxygen are different,20 and the catalytic comparison of the electronic effect is unfair. To confirm the catalytic properties of ONNO–Salan Al complexes for CL ROP, a series of ONNO–Salan Al complexes (Scheme 1C) were synthesized, and their CL polymerization was studied.


image file: d4dt02923b-f5.tif
Fig. 5 rac-LA ROP using ONNO–Salan Al complexes as catalysts.

Results and discussion

Synthesis and characterization of Al complexes

For the synthesis of ONO-type ligands (Scheme 1A), a solution of two equivalents of 6-tert-butyl-4-substituted-phenol, one equivalent of aqueous formaldehyde, and one equivalent of benzylamine in ethanol was refluxed for 2 days. For the synthesis of ON-type ligands (Scheme 1B), para-formaldehyde, pyrrolidine, and 6-tert-butyl-4-substituted-phenol (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1) were set in ethanol and stirred for 2 days. For the synthesis of ONNO-type ligands (Scheme 1C), two equivalents of 2-hydroxy-5-substituted-benzaldehyde were reacted with one equivalent of ethylenediamine to form the Salen ligands first, and sodium borohydride was used to reduce the Salen ligands to form Salan ligands. The Salan ligands reacted with formaldehyde (37% aqueous solution) in ethanol, and the solution was refluxed for 4 h. Sodium borohydride was transferred to the solution and refluxed for one day. BuONNOOMe–OH and BuONNOBr–OH were synthesized by following the literature procedure.21 For the synthesis of OO type ligands22 (Scheme 1D), 4-chlorobenzaldehyde and 6-tert-butyl-4-substituted-phenol (1[thin space (1/6-em)]:[thin space (1/6-em)]1) were combined in hexane with para-tolyl sulfonic acid as a catalyst and refluxed for two days. The ligand “OOH–H” reacted with bromine to form OOBr–H. NNOOOMe–OH and NNOOBr–OH (Scheme 1G) were synthesized following the literature procedure.16aNNOOMe–OH and NNOBr–OH (Scheme 1H) were synthesized by the reaction of para-formaldehyde, N1-benzyl-N2,N2-dimethylethane-1,2-diamine, and 6-tert-butyl-4-substituted-phenol (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1) in ethanol and stirred for 1 day. All Al complexes were synthesized from ligands with a stoichiometric amount of AlMe3 in toluene in quantitative yield. The 1H NMR peak of the proton on phenolate disappeared after reacting with AlMe3, and the presence of methyl groups on Al atoms was observed. Since L2AlMe type complexes bearing 2,6-di-tert-butyl-4-methoxyphenolate (Scheme 1E) could not be isolated, 2,6-di-tert-butyl-4-methylphenol was used to synthesize OMe2–Al. The crystals of dinuclear ONOOMe–AlOH (CCDC 2247652) were obtained from the NMR tube in which ONOOMe–Al was dissolved in CDCl3 solution, and ONOOMe–AlOH was formed by the reaction of ONOOMe–Al and moisture. The Al atoms revealed a distorted trigonal bipyramidal geometry (τ5 = 0.84)23 with one nitrogen and two oxygen atoms of the ligand and two hydroxyl groups (Fig. 6). The structure of dinuclear ONOOMe–AlOH implies that ONOOMe–AlMe may react with benzyl alcohol (BnOH) to form the dinuclear Al complex with bridging benzyl alkoxide.
image file: d4dt02923b-f6.tif
Fig. 6 Molecular structure of ONOOMe–AlOH depicted as 50% probability ellipsoids (all hydrogen atoms are omitted for clarity).

The high concentration of ONOMe–Al in toluene was placed in the freezer (−20 °C) for one week, and then the crystals of mononuclear ONOMe–Al (CCDC 2247645) were obtained. The Al atoms revealed a distorted tetrahedral geometry (τ4 = 0.92)24 with one nitrogen and one oxygen atom of the ONOMe ligand and two methyl groups (Fig. 7).


image file: d4dt02923b-f7.tif
Fig. 7 Molecular structure of ONOMe–Al depicted as 50% probability ellipsoids (all hydrogen atoms are omitted for clarity).

A high concentration of ONNOOMe–Al in THF was placed in the freezer (4 °C) for one month, and then the crystals of mononuclear ONNOOMe–Al (CCDC 2236052) were obtained. The Al atom revealed a distorted quadrangular pyramidal geometry (τ5 = 0.75)23 with two nitrogen and two oxygen atoms of Salan ligand and one methyl group (Fig. 8).


image file: d4dt02923b-f8.tif
Fig. 8 Molecular structure of ONNOOMe–Al depicted as 50% probability ellipsoids (all hydrogen atoms are omitted for clarity).

The method for obtaining the crystals of mononuclear ONNOBr–Al (CCDC 2247643,Fig. 9) was similar to that for ONNOOMe–Al. The Al atom revealed the distorted quadrangular pyramidal geometry (τ5 = 0.79)23 that was similar to ONNOOMe–Al (Fig. 8). The crystal data revealed that the electron-withdrawing groups, such as the bromo substituent, reduce the electron density of the Al atom and increase the bond strength between the Al and the methyl group.


image file: d4dt02923b-f9.tif
Fig. 9 Molecular structure of ONNOBr–Al depicted as 50% probability ellipsoids (all hydrogen atoms are omitted for clarity).

A high concentration of OOBu–Al in tetrahydrofuran (THF) was placed in the freezer (−20 °C) for three days, and then the crystals of mononuclear OOBu–Al (CCDC 2247640) were obtained. The Al atoms revealed a distorted tetrahedral geometry (τ4 = 0.92)24 with two oxygen atoms of the OOBu ligand, one methyl group, and one oxygen atom of THF (Fig. 10).


image file: d4dt02923b-f10.tif
Fig. 10 Molecular structure of OOBu–Al depicted as 50% probability ellipsoids (all hydrogen atoms are omitted for clarity).

Polymerization of ε-caprolactone

Table 1 presents the catalytic activity results for CL polymerization at 25 °C in toluene in the presence of benzyl alcohol (BnOH). The polymerization results revealed that Al complexes with an electron-donating group exhibited higher catalytic activity than Al complexes with an electron-withdrawing group.
Table 1 ε-Caprolactone polymerization with Al complexes as catalystsa
Entry Cat [CL][thin space (1/6-em)]:[thin space (1/6-em)][Cat][thin space (1/6-em)]:[thin space (1/6-em)][BnOH] Time (min) Conv.b (%) M nCal M nNMR M nGPC Đ k obs × 10−3 (error) min−1
a In general, the reaction was carried out in toluene with [CL] = 2.00 M. b Calculated from the molecular weight of Mw(CL) × [CL]0/[BnOH]0 × conversion yield + Mw(BnOH). c Data were obtained through 1H NMR analysis. MnNMR was calculated as follows: molecular weight of CL × (integration of the peak at 4.0 ppm/integration of the peak at 7.3 ppm) × 5/2 + Mw(BnOH). d Obtained through gel permeation chromatography (GPC). Values of MnGPC were measured by the GPC and multiplied by 0.56.
1 ONOOMe–Al 85 94 10[thin space (1/6-em)]800 7900 17[thin space (1/6-em)]300 1.19 48.5 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
2 ONOBr–Al 70 93 10[thin space (1/6-em)]900 13[thin space (1/6-em)]500 18[thin space (1/6-em)]400 1.13 43.2 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
3 ONOMe–Al 20 >99 5800 9900 7800 1.32 152.0 (3)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2
4 ONOMe–Al 40 90 10[thin space (1/6-em)]400 14[thin space (1/6-em)]900 15[thin space (1/6-em)]500 1.33 57.6 (16)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
5 ONBr–Al 20 94 5500 10[thin space (1/6-em)]500 7500 1.37 124.0 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2
6 ONBr–Al 140 92 10[thin space (1/6-em)]600 9900 10[thin space (1/6-em)]200 1.19 17.3 (7)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
7 ONNOOMe–Al 75 95 10[thin space (1/6-em)]900 11[thin space (1/6-em)]700 17[thin space (1/6-em)]000 1.41 92.4 (2)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
8 ONNOBr–Al 70 92 10[thin space (1/6-em)]600 10[thin space (1/6-em)]400 11[thin space (1/6-em)]200 1.34 53.8 (32)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
9 BuONNOOMe–Al 1500 90 10[thin space (1/6-em)]400 9300 6300 1.05 1.6 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
10 BuONNOBr–Al 1680 90 10[thin space (1/6-em)]400 8800 6700 1.04 1.4 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
11 OOOMe–Al 1620 92 10[thin space (1/6-em)]600 11[thin space (1/6-em)]900 18[thin space (1/6-em)]100 1.17 1.4 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
12 OOBr–Al 2460 90 10[thin space (1/6-em)]400 14[thin space (1/6-em)]100 10[thin space (1/6-em)]400 1.10 1.0 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
13 OOBu–Al 660 93 10[thin space (1/6-em)]600 18[thin space (1/6-em)]000 11[thin space (1/6-em)]800 1.21 4.3 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
14 OMe2–Al 13 90 10[thin space (1/6-em)]400 14[thin space (1/6-em)]100 19[thin space (1/6-em)]800 1.70 244.5 (4)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
15 OBr2–Al 40 95 10[thin space (1/6-em)]900 12[thin space (1/6-em)]800 17[thin space (1/6-em)]300 1.36 121.1 (50)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
16 OOMe–Al 18 90 5200 4700 6300 1.30 143.2 (35)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2
17 OOMe–Al 50 90 10[thin space (1/6-em)]400 11[thin space (1/6-em)]100 12[thin space (1/6-em)]100 1.18 48.6 (9)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
18 OBr–Al 70 90 5200 5400 3100 1.15 33.9 (4)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2
19 OBr–Al 160 92 10[thin space (1/6-em)]600 8800 8900 1.06 16.9 (5)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
20 NNOOOMe–Al 480 90 10[thin space (1/6-em)]400 10[thin space (1/6-em)]800 11[thin space (1/6-em)]400 1.08 4.8 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
21 NNOOBr–Al 1620 90 10[thin space (1/6-em)]400 10[thin space (1/6-em)]800 11[thin space (1/6-em)]400 1.06 1.5 (1)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
22 NNOOMe–Al 40 92 5400 5400 4800 1.09 65.6 (23)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2
23 NNOOMe–Al 140 92 10[thin space (1/6-em)]600 10[thin space (1/6-em)]500 10[thin space (1/6-em)]300 1.13 18.8 (9)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
24 NNOBr–Al 60 92 5400 5900 4800 1.12 43.9 (9)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]2
25 NNOBr–Al 220 93 10[thin space (1/6-em)]700 12[thin space (1/6-em)]800 8500 1.22 12.3 (5)
100[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1


The overall polymerization rates of Al catalysts with different ligands (ranked by their kobs values, Fig. 11) are in the following order: ONBr–Al > OBr2–Al > ONNOBr–Al > NNOBr–Al > ONOBr–Al > OBr–Al > NNOOBr–Al > BuONNOBr–Al > OOBr–Al. OBr2–Al exhibited higher catalytic activity than OOBr–Al because the flexibility25 of the coordination from phenolate to Al atom allows the Al atom to adjust its electron density to facilitate CL coordination to the Al center and alkoxide initiation. According to Tolman16b results, the order of the catalytic activity of Al complexes should be OBr–Al > ONBr–Al > NNOBr–Al and OOBr–Al > ONOBr–Al; however, in fact, OOBr–Al and OBr–Al exhibited lower catalytic activity compared with aminophenolate Al complexes. Although the amino group of ONOBr–Al still occupies a coordination site and competes with CL coordination compared to OOBr–Al, ONOBr–Al exhibited higher catalytic activity than OOBr–Al, and the amino group of ONOBr–Al seems to have other functions in the catalytic process to enhance the catalytic activity. The phenomenon of NOBr–Al exhibiting higher catalytic activity than OBr–Al and OBr2–Al supports the earlier speculation regarding the role played by the amino group. In addition, ONNOBr–Al exhibited higher catalytic activity than BuONNOBr–Al, implying that two bulky t-butyl groups prevent CL from approaching the Al center. The Al complexes with electron-donating substituents on ligands exhibited higher catalytic activity than those with bromo substituents. In addition, the catalytic trend of these Al complexes is similar to those with bromo substituents, except that OMe2–Al exhibited higher catalytic activity than ONOMe–Al, and the catalytic activity of OOMe–Al is higher than those of NNOOMe–Al and ONOOMe–Al. This phenomenon is unexplained because the functional groups between OMe2–Al (with the methyl group in the phenolate ring) and ONOMe–Al (with the methoxy group in the phenolate ring) are different. The catalytic activity of ONOOMe–Al is higher than that of NNOOOMe–Al because the dimethyl amino group16b of NNOOOMe–Al occupies a coordination site to compete against CL coordination and reduces the polymerization rate.


image file: d4dt02923b-f11.tif
Fig. 11 Comparison of the catalytic activity of CL polymerization by using Al complexes bearing various substituted ligands.

The Đ values of producing PCL become broader when Al complexes exhibit higher catalytic activity. Despite the low coordination number of Al complexes, such as O–Al, O2–Al, and ON–Al type complexes, they exhibited higher catalytic activity, and a reduced control over CL polymerization was observed when using these Al complexes. It may be attributed to the fact that when there are more exposed spaces around the uncrowded Al center, the chance of CL coordination increases along with the polymerization rate. However, it also increases the possibility of the coordination of the polymer to the metal center and then accelerates the occurrence of transesterification.26

Catalytic polymerization ability test of the mPEG-b-PCL copolymer

The advantage of alkyl Al catalysts is that the initiators (alcohol precursors) can be easily exchanged with the alkyl group on the Al atom, and this advantage is very important for synthesizing PCL copolymers for drug delivery systems. Therefore, these Al complexes with mPEG 1900 as an initiator were tested for catalytic polymerization of the mPEG-b-PCL copolymer, and the optimal controllability of producing the mPEG-b-PCL copolymer is shown in Table 2.
Table 2 ε-Caprolactone polymerization with Al complexes as catalysts with the mPEG initiator ([CL][thin space (1/6-em)]:[thin space (1/6-em)][Cat][thin space (1/6-em)]:[thin space (1/6-em)][mPEG] = 45[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1)a
Entry [CL][thin space (1/6-em)]:[thin space (1/6-em)][Cat][thin space (1/6-em)]:[thin space (1/6-em)][mPEG] Time (h) Conv.b (%) M nCal M nNMR M nGPC Đ
a Generally, the reaction was carried out in toluene with [CL] = 0.90 M at 25 °C. b Calculated from the molecular weight of Mw(CL) × [CL]0/[mPEG]0 × conversion yield + Mw(mPEG). c Data were obtained through 1H NMR analysis. MnNMR was calculated as follows: molecular weight of CL × (integration of the peak at 4.0 ppm/integration of the peak at 3.62 ppm) × 86 + Mw(mPEG) (an example in Fig. S60†). d Obtained through gel permeation chromatography (GPC). e [CL] = 1.55 M in toluene at 25 °C. f [CL] = 3.11 M in toluene at 25 °C. g [CL] = 3.51 M in toluene at 25 °C. h [CL] = 4.47 M in toluene at 25 °C.
1 ONOOMe–Al 6 92 6600 6500 5700 1.16
2 ONOMe–Al 25 min 95 6800 6700 6100 1.46
3 BuONNOOMe–Al 24 94 6700 6600 4800 1.06
4 OOOMe–Al 2 90 6500 6300 5900 1.13
5 OMe2–Al 2 94 6700 6800 6200 1.10
6 OOMe–Al 40 min 95 6800 6200 7000 1.19
7 NNOOOMe 8 90 6500 6200 4700 1.06
8e OOMe–Al 40 min 90 5100 5500 4800 1.13
31[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
9f OOMe–Al 1.5 90 9900 9200 12[thin space (1/6-em)]500 1.18
78[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
10g OOMe–Al 2.5 90 10[thin space (1/6-em)]900 10[thin space (1/6-em)]400 16[thin space (1/6-em)]400 1.19
88[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1
11h OOMe–Al 5 85 12[thin space (1/6-em)]700 14[thin space (1/6-em)]600 19[thin space (1/6-em)]300 1.19
112[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1


Among these Al complexes, ONOMe–Al exhibited the highest polymerization rate (25 min, conv. = 95%) but lower control over polymerization (Đ = 1.46), and BuONNOOMe–Al still exhibited the lowest catalytic activity (24 h, conv. = 94%). Because the catalytic activity of OOMe–Al ranked second with greater controllability, OOMe–Al was used as a catalyst to test its controllability, as shown in entries 8–11 of Table 2. When the linear relation MnGPC and ([monomer]0 × conv.)/[BnOH] (entries 6 and 8–11 in Table 2), as presented in Fig. 12, was considered, OOMe–Al as a catalyst was believed to exhibit well-controlled polymerization with narrow Đ values (1.13–1.20) for CL polymerization. The polymerization results revealed that OOMe–Al is a good candidate for synthesizing the mPEG-b-PCL copolymer.


image file: d4dt02923b-f12.tif
Fig. 12 Linear plots of MnGPCversus ([CL]0 × conv.)/[mPEG], with Đ indicated by red solid dots (entries 6 and 8–11) in Table 2.

CL polymerization mechanism using OOBu–Al as a catalyst

Although the literature22c,27 reported that Al complexes bearing biphenolates formed mononuclear Al complexes, dinuclear Al complexes bridged by benzyl alkoxide27a,b were also reported (such as the structure of (OOBu–AlOBn)2 in Scheme 2) and have maintained dinuclear geometric structures during the CL polymerization process. To determine whether the dimeric structure in the solution exists, the diffusion-order NMR spectroscopy (DOSY) analysis of OOBu–Al and (OOBu–AlOBn)2 in CDCl3 was carried out. The DOSY data (Fig. S61 and S62) revealed that the diffusion coefficient (D) of OOBu–Al was 10.84 × 10−10 m2 s−1, and its hydrodynamic radius (r) was 3.66 Å. However, the D value of (OOBu–AlOBn)2 was 6.01 × 10−10 m2 s−1, and its hydrodynamic radius (r) was 6.71 Å, larger than that of OOBu–Al. The DOSY data exhibited the dimeric structure of (OOBu–AlOBn)2 in solution. To understand the difference in the catalytic activity between dinuclear Al complexes bridged by benzyl alkoxide and the in situ reaction of mononuclear Al complexes with BnOH, (OOBu–AlOBn)2 was obtained through the reaction of OOBu–Al and BnOH (1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio, Scheme 2) and used to study the CL polymerization compared with the CL polymerization by using the in situ reaction of mononuclear OOBu–Al with BnOH. The polymerization results revealed that (OOBu–AlOBn)2 exhibited lower catalytic activity for CL polymerization (2880 min, conv. = 90%, kobs = 0.0008 min−1) compared to the in situ reaction of mononuclear OOBu–Al with BnOH (660 min, conv. = 93%, kobs = 0.0043 min−1). These results implicated that Al catalysts with bridging benzyl alkoxide exhibited low catalytic activity because only one lone pair can initiate CL polymerization, and the bridging oxygen atom between two aluminum atoms fixes the orientation of the lone pair. This makes the initiation of CL polymerization by alkoxide difficult. Therefore, the electronic effect and amino chelating effect may affect the structures of Al catalysts with bridging benzyl alkoxide, thereby converting them into a more active structural state. Many Al complexes such as these bearing (2,6-bis(hydroxyalkyl)pyridine),28 2,2′-oxybis(N-cyclopentylanilinate),29 amine-pyridine-bis(phenolate),30 β-diketonate,31 biphenolate phosphine,32 β-quinolyl enolato,33 catalen,34 catam,7e,35 ketiminate,11,36 NHC bis-phenolate,37N-phenylpicolinimidate,38 phenolate,39 pyrrolyl amide,40 Salen,41 and triphenolate22c,42 were reported with bridging alkoxide structures that are inactive11,22c,43 during the polymerization process. Herein, biphenolate-associated Al catalysts with bridging benzyl alkoxide were used as a model, and the CL polymerization mechanism was investigated by density functional theory (DFT) calculations carried out using the Gaussian 09E program package.44
image file: d4dt02923b-s2.tif
Scheme 2 Reaction of OOBu–Al with BnOH to form (OOBu–AlOBn)2.

Confirming the active species of the Al catalyst is the first necessary investigation. For dimer–tetramer interconversions (Scheme 3), the OOOMe ligand structures in the calculation are simplified to reduce the calculation loading using methoxide ligands to replace benzyl alkoxides. All the calculations were carried out at the B3LYP level with the 6-31G(d) basis sets. All DFT data regarding optimizing the most suitable active species of Al catalysts have been presented in the ESI. They revealed that the syn-Dimer-B is likely the catalytic site to accept CL coordination to form the intermediate I (the CL coordination on the inner site of syn-Dimer-B as shown in Fig. 13) for the ring-opening reaction.


image file: d4dt02923b-s3.tif
Scheme 3 The interconversion from Dimer-A to Dimer-B through tetranuclear species T.

image file: d4dt02923b-f13.tif
Fig. 13 Structures of intermediate I (the syn-Dimer-B with CL coordination).

In the polymerization catalytic cycle (Fig. 14), the catalyst (syn-Dimer-B) first accepts one CL to form intermediate I. The Al(1) of intermediate I (Fig. 13) has a twisted trigonal bipyramidal structure with CL and bridging methoxide on the axial sites. The bond angle between two axial ligands (Ocar–Al(1)–Oμ-Me) is 162.06°. The bond length between Ocar and Al(1) is 1.989 Å, slightly longer than those reported in the literature (1.8–1.9 Å),45 indicating that the Al center weakly binds the CL. The terminal methoxide's oxygen atom shows a hydrogen bond-like interaction with the α-hydrogen of CL with a Hα⋯Otm-Me distance of 2.174 Å, which is shorter than the sum of equilibrium van der Waals radii (eVDWR) of hydrogen and oxygen (∼2.6 Å),46 indicating that a strong hydrogen bond-like interaction acts as a second attractive force to bind CL to the catalyst.


image file: d4dt02923b-f14.tif
Fig. 14 The catalytic cycle of ring-opening polymerization catalyzed by syn-Dimer-B.

Intermediate I converts to intermediate II through transition structure TS-1 (Fig. 15), and the terminal OMe ligand is added to the CL molecule. The Al(1) center maintains a trigonal bipyramidal structure with the bridged phenoxide and terminal OMe as the axial ligands. The CL leads to the more Lewis-acidic equatorial site being further activated. The carbonyl group of CL rotates to face the terminal methoxide ligand with a Ccarbonyl⋯OOMe distance of 2.005 Å, which is between the sum (oxygen and carbon atoms) of eVDWR (3.75 Å) and single-bond covalent radii (1.38 Å),47 indicating that the Ccarbonyl–Ocarbonyl bond is forming. Moreover, the addition occurs at the space between the two ortho-tBu groups on the bidentate ligand. This demonstrates that the catalyst possesses a protected open site for the catalytic reaction to prevent transesterification.


image file: d4dt02923b-f15.tif
Fig. 15 The transition structures of the catalytic reaction.

After forming the CL–OMe adduct, the complex transforms into intermediate II; the adduct chelates on Al(1) through its Ocarbonyl and OOMe atoms with a small bit angle of 67.67°. The axial coordinating atoms lead to Oμ-Ph and OOMe. The Ocarbonyl-Al(1) length is 1.788 Å, slightly longer than the OOMe–Al(1) bond of I (1.784 Å). The ether oxygen of the adduct is weakly coordinated to Al with an OOMe–Al length of 2.112 Å, which can easily dissociate from the Al center. The OOMe atom dissociates from Al(1) to form intermediate III. The Al(1) center, due to losing one coordination ligand, converts to a tetrahedral geometry. Interestingly, the central H of the chelating ligand moves closer to Al(1), and the H⋯Al(1) distance is 2.760 Å, which is shorter than the sum of van der Waals radii of H and Al (2.94 Å), offering an agnostic-like interaction. The Ocarbonyl–Al(1) bond is 1.736 Å, which elongates slightly compared to intermediate II (1.788 Å), due to the decrease of coordination points of the adduct. The Ocarbonyl–Ccarbonyl bond rotates, and the Oest atoms (the oxygen atom of the seven rings) of the adduct coordinate with Al(1) to form intermediate IV. The Oest–Ccarbonyl bond length is elongated from 1.429 Å (III) to 1.456 Å (IV) because the coordination of the Oest atom on Al helps weaken the bond. Moreover, the structure of the adduct, due to the assistance of the complex, can dissociate the Oest–Ccarb bond. The complex then undergoes a CL ring opening through transition structure TS-2 (Fig. 15). The Ccarbonyl⋯Oest distance is further elongated to 1.981 Å, which is much longer than the C–O bond of IV. Moreover, the Ccarbonyl is nearly converted to sp2 hybridization (the sum of the three bond angles around it is 352.89°). This indicates that the structure of the adduct on TS-2 is close to the ring-opened oligomer. Finally, the CL ring opens to form intermediate V. The newly generated carbonyl group is located at an equatorial position of the Al(1) center, and the newly generated alkoxide ligand is in an axial position. Intermediate V contrasts with intermediate I, where the CL carbonyl oxygen is on an axial site, and the terminal methoxide is on an equatorial site.

The coordination of CL on syn-Dimer-B only requires 4.11 kcal mol−1 (free energy), and this process can release 8.84 kcal mol−1 heat (enthalpy change), indicating that the coordination site is readily accessible for CL. To consider the free energy changes of the mechanism (Fig. 16), after forming I, the reaction goes through TS-1, overcoming the highest energy barrier (16.65 kcal mol−1) of the catalytic reaction. This is a moderate energy barrier caused by the bulky ortho-tBu group on the phenol moieties, partially disturbing the addition of the terminal alkoxide to CL. After the addition to form II, the isomerization between II, III, and IV only shows slight energy differences due to the small structural differences. Then, the CL ring opens to form V through TS-2, facing a relatively small energy barrier of 11.92 kcal mol−1. Finally, because of the large chelating ring of the oligomer, V is the intermediate with the highest free energy. It can revert to structure I easily by accepting one more CL molecule. Compared with the Br-substituted OOBr–Al catalyst, the highest energy barrier from I to TS-1 is about 1 kcal mol−1 lower than that from I–Br to TS-1–Br (0 to 17.625 kcal mol−1), indicating that this mechanism also can explain the relatively high reactivity of OOOMe–Al well, as shown in Table 1.


image file: d4dt02923b-f16.tif
Fig. 16 The energy profile of the catalytic reaction.

According to the DFT results, Al complexes bearing phenolate ligands can easily form bridging benzyl alkoxide-associated dinuclear complexes with low catalytic activity. Thus, they undergo rearrangement to form active terminal benzyl alkoxide-associated dinuclear complexes for the ring-opening polymerization process. Many studies7e,f,11,16b,22c,28–30,32–38,42,48 have also reported that Al complexes (especially those containing phenolate ligands) easily form bridging benzyl alkoxide-associated dinuclear complexes. The bridging benzyl alkoxide-associated dinuclear Al complexes with electron-donating groups have relatively weak bonds between bridging benzyl alkoxide and aluminum atoms. Their bonds can be easily broken to rearrange and form the active terminal benzyl alkoxide-associated dinuclear complexes and increase their catalytic activity. However, the phenolate ligands, with electron-donating groups, may provide more electrons to the Al atoms to increase their electron density. Al catalysts with high electron density on Al atoms are expected to reduce the chance of benzyl alkoxide bridging to form the terminal benzyl alkoxide-associated Al complexes. Like the electron-donating groups in the phenolate ligands, the chelating amino group also provides electrons to the Al atoms. Using this concept to examine the catalytic activity of the Al complexes used in this study, the following conclusions can be drawn:

1. OBr–Al and OBr2–Al exhibit a higher ability to rearrange from the bridging benzyl alkoxide-associated dinuclear complexes to active terminal benzyl alkoxide-associated dinuclear complexes compared to OOBr–Al because OOBr–Al with a rigid metal–ligand eight-membered ring has a poor ability25 to enhance the electron density of the Al center through the oxygen pi-donation of the OOBr ligand; therefore, OOBr–Al exhibits the lowest catalytic activity [Fig. 17(1)].


image file: d4dt02923b-f17.tif
Fig. 17 Comparison of the catalytic activity of CL polymerization by using Al complexes.

2. From four-coordinate OBr–Al to four-coordinate NOBr–Al and five-coordinate NNOBr–Al [Fig. 17(2)], the catalytic activity of NOBr–Al and NNOBr–Al is higher compared to OBr–Al since the chelating amino group may increase the electron density of the Al center to carry out the rearrangement. However, five-coordinate NNOBr–Al revealed a more crowded environment around the Al center than four-coordinate NOBr–Al and reduced the ability of the CL coordination to the Al center.

3. Three-coordinate OBr2–Al can rearrange to form terminal benzyl alkoxide-associated Al complexes because of the oxygen pi-donation of two terminal OBr ligands.25 In the case of ONOBr–Al, adding an amino group to link two phenolates did not improve ONOBr–Al's ability to rearrange because of the formation of a rigid seven-membered ring of ONOBr–Al. As the number of amino groups in phenolate increases, the ability of the CL coordination to the Al center decreases [Fig. 17(3)]. In addition, five-coordinate NNOOBr–Al and BuONOOBr–Al with the high coordination number of Al complexes prevent bridging benzyl alkoxide structure, and the dimethyl amino group of less bulky NNOOBr–Al is more flexible compared with BuONOOBr–Al to dissociate from the Al center that easily accepts the CL coordination. Moreover, the electron donating group in the phenolate ring may increase the electron density of the Al center to afford a facile dimethyl amino group dissociation and improve the catalytic activity (Tolman's research).16b

Conclusions

A series of Al complexes bearing phenolates and aminophenolates with different coordination numbers were synthesized, and their CL polymerization was investigated. Most Al complexes with low coordination numbers, such as ON–Al, O2–Al, and O–Al type complexes, exhibited high catalytic activity in CL polymerization, except OO–Al and ONO–Al type complexes due to the rigid ring between Al–ligand structures that reduces the ability of the phenolates to control the electron density of the Al center. The ON–Al system exhibited a higher catalytic activity (1.1–3.9-fold for kobs) than the O–Al type complex, implying that the amino group could improve the catalytic activity of three-coordinate Al complexes. Five-coordinate Al complexes, such as NNOO–Al and ONNO–Al type complexes, exhibited lower catalytic activity than four-coordinate ones because of the crowded Al center, which is not conducive to CL coordination. In addition, ONO–Al type complexes exhibited higher catalytic activity compared to NNOO–Al and ONNO–Al type complexes (9.9 and 37.3-fold for kobs). Removing the amino group from high-coordinate Al complexes can enhance catalytic activity by reducing the Al center's repulsion to accept CL coordination. Moreover, Al complexes with an electron-donating substituent on the phenolate ring exhibited higher catalytic activity, contrasting with those with an electron-withdrawing one. DFT calculations revealed that the dinuclear Al complex with two bridging alkoxides exhibited low activity for CL polymerization because of the low initiating ability of two bridging methoxides. The dinuclear Al complex with two bridging alkoxides had to rearrange to a phenolate bridged dinuclear Al complex with terminal alkoxide. Then, CL coordinated to one Al center with terminal alkoxide. The electron-donating substituents on the phenolate ring can increase the bridging ability of the phenolate group and then increase the CL polymerization rate. In addition, the amino group can also make the Al center electron-rich by chelating to the Al center, facilitating the structural rearrangement of the dinuclear Al complex with two bridging alkoxides. For synthesizing the mPEG-b-PCL copolymer, OOMe–Al is an efficient candidate with a high polymerization rate and controllability.

Experimental section

Standard Schlenk techniques and an N2-filled glovebox were used throughout the isolation and handling of compounds. Solvents, ε-caprolactone, and deuterated solvents were purified before use. ε-Caprolactone was purified by distillation with magnesium sulfate anhydrous. The organic solvent was purified by distillation in the presence of sodium and benzophenone. Deuterated chloroform was purified by distillation in the presence of calcium hydride. Deuterated chloroform, ε-caprolactone, and trimethyl aluminum were purchased from Aldrich. Methoxy poly(ethylene glycol) 1900, ethane-1,2-diamine, N,N′-dimethylethane-1,2-diamine, 2-tert-buty-4-bromo-phenol, 3-tert-butyl-4-hydroxyanisole, benzylamine, paraformaldehyde, pyrrolidine, triethylamine, 5-bromosalicylaldehyde, ethylenediamine, sodium borohydride, 2-tert-butylphenol, 2,4-di-tert-butylphenol, 4-chlorobenzaldehyde, methyl magnesium bromide solution (1.0 M) in dibutyl ether, bromine, and 4-bromo-2,6-di-tert-butylphenol were purchased from Sigma-Aldrich. 2-Hydroxy-5-methoxybenzaldehyde was purchased from Nova-Matls. 1H and 13C NMR spectra were recorded on a JEOL JNM-ECS400 (400 MHz for 1H and 100 MHz for 13C) spectrometer with chemical shifts in ppm from the internal TMS or center line of CDCl3. For 1H PFGNMR diffusion measurement, all experiments were performed on a JEOL 400 MHz liquid state NMR spectrometer operating at 9.4 Tesla (with a proton resonance frequency of 400.052 MHz). Each freshly synthesized sample was tightly sealed in a 7 mm liquid state NMR tube so that the humidity of the sample did not change throughout the experiment (about 10 minutes each). The sample lengths were all longer than 40 mm to minimize the magnetic susceptibility effect from the sample. The sample was static (zero spinning speed) and with 2H lock used during the diffusion measurement to maintain the homogeneity of the magnetic field within 2 Hz inside the sample region and less than 1 Hz of fluctuation overnight. The 90° pulse width for 1H was 9.5 μs. The spectral widths were 15 ppm for all samples. Standard NMR reference TMS was used and all experiments were performed at room temperature (25 °C). The spectrometer is equipped with a magnetic field gradient coil that enables the generation of a magnetic field gradient up to 30 G cm−1 (300 mT m−1). The pulse sequence used for measuring diffusion coefficient is DOSY with a diffusion time (Δ) of 60 ms and a gradient interval (δ) of 5 ms, respectively. The pulsed field gradients (mT m−1) were 0, 60, 120, 240, and 300. The scan number was 8 and the recycle delay was 10 s for all samples. Microanalyses were performed using a Heraeus CHN-O-RAPID instrument. GPC measurements were performed on a Jasco PU-2080 PLUS HPLC pump system equipped with a differential Jasco RI-2031 PLUS refractive index detector using THF (HPLC grade) as an eluent (flow rate 1.0 mL min−1, at 40 °C). The chromatographic column was JORDI Gel DVB 103 Å, and primary polystyrene standards made the calibration curve to calculate MnGPC. All X-ray diffraction data were accumulated using Rigaku Oxford Diffraction single crystal X-ray diffractometers with Mo Kα radiation (λ = 0.71073 Å). Data collection was executed using the CrysAlisPro 1.171.41.56a program. Cell refinement and data reduction were achieved using the CrysAlisPro 1.171.41.56a program. The structure was determined using the Olex2/ShelXL program and refined using full-matrix least squares. All non-hydrogen atoms were refined anisotropically, whereas hydrogen atoms were placed at calculated positions and included in the final refinement stage with fixed parameters. NNOOOMe–OH,16aNNOOBr–OH,16aONNOOMe–OH,49ONNOBr–OH,49bBuONNOOMe–OH,21OMe2–Al,50OBr2–Al,51NNOOOMe–Al,16a and NNOOBr–Al[thin space (1/6-em)]16a were synthesized by following the literature procedure.

Synthesis of ONOOMe–OH

In a 100 mL round-bottom flask, 3-tert-butyl-4-hydroxyanisole (3.60 g, 20.0 mmol), paraformaldehyde (37 wt% in water, 2 mL), and benzylamine (1.07 g, 10.0 mmol) were added and dissolved in 60 mL of EtOH. The mixture was refluxed, and the reaction was monitored by TLC, which revealed that the reaction was complete after 4 h. The mixture was then warmed to room temperature, and the volatiles were removed under vacuum. The pure product was obtained after purification via silica gel column chromatography (10/1 hexane/ethyl acetate). Yield: 3.92 g (80%). 1H NMR (400 MHz, CDCl3) δ 7.42–7.30 (m, 3H, meta-H + para-H of PhCH2), 7.05 (br, 2H, ortho-H of PhCH2), 6.83 (d, 2H, J = 2 Hz, para-H of PhtBu), 6.54 (d, 2H, J = 2 Hz, ortho-H of PhtBu), 3.76 (s, 6H, OCH3), 3.64 (s, 4H, ArCH2N), 3.58 (s, 2H, PhCH2N), 1.41 (s, 18H, C(CH3)3). 13C NMR (100 MHz, CDCl3, δ): 152.37 (MeO–C of Ar), 148.57 (HO–C of Ar), 138.51 (tBu–C of ArtBu), 137.45 (NCH2C of Ph), 129.62 (ortho-C of NCH2Ph), 129.12 (meta-C of NCH2Ph), 128.10 (para-C of NCH2Ph), 122.90 (NCH2C of Ar), 113.52 (ortho-C of NCH2Ar), 112.82 (para-C of NCH2Ar), 58.54 (NCH2Ph), 56.78 (OCH3), 55.77 (NCH2Ar), 34.99 (C(CH3)3), 29.54 (C(CH3)3). Anal. calc. (found) for C31H41NO4: C 75.73 (75.52), H 8.41 (8.25), N 2.85 (2.68).

Synthesis of ONOBr–OH

This compound was prepared following the same procedure described for ONOOMe–H by using 2-(tert-butyl)-4-bromo-phenol in place of 3-tert-butyl-4-hydroxyanisole. Yield: 4.57 g (78%). 1H NMR (400 MHz, CDCl3) δ 7.25–7.13 (br, 7H, ortho-H–ArCH2 + C6H5CH2), 6.99 (br, 2H, ortho-H–ArtBu), 3.51 (s, 4H, ArCH2N), 3.46 (s, 2H, PhCH2N), 1.29 (s, 18H, C(CH3)3). 13C NMR (100 MHz, CDCl3, δ): 156.62 (HO–C of Ar), 139.50 (tBu–C of Ar), 136.85 (NCH2C of Ph), 130.82 (ortho-C of ArtBu), 129.97 (meta-C of NCH2Ph), 129.46 (ortho-C of NCH2Ph), 129.27 (NCH2C of Ar), 128.35 (para-C of NCH2Ph), 124.15 (ortho-C of NCH2Ar), 111.71 (Br–C of Ar), 58.49 (NCH2Ph), 55.94 (NCH2Ar), 34.96 (C(CH3)3), 29.32 (C(CH3)3). Anal. calc. (found) for C29H35Br2NO2: C 59.10 (59.22), H 5.99 (5.79), N 2.38 (2.22).

Synthesis of ONOMe–OH

To a 100 mL round-bottom flask, 3-tert-butyl-4-hydroxyanisole (3.60 g, 20.0 mmol), paraformaldehyde (37 wt% in water, 2 mL), and pyrrolidine (1.85 g, 26.0 mmol) were added and dissolved in 50 mL of EtOH. The mixture was refluxed, and the reaction was monitored by TLC, revealing that the reaction was complete after 4 h. The mixture was then warmed to room temperature, and the volatiles were removed under vacuum. The pure product was obtained after purification via silica gel column chromatography using a 10/1 hexane/ethyl acetate mixture as an eluent system. Yield: 4.31 g (82%). 1H NMR (400 MHz, CDCl3) δ 6.80 (d, 1H, J = 2 Hz, para-H of PhtBu), 6.43 (d, 1H, J = 2 Hz, ortho-H of PhtBu), 3.78 (s, 2H, ArCH2N), 3.76 (s, 3H, OCH3), 2.62 (br, 4H, N(CH2CH2)2), 1.86 (br, 4H, N(CH2CH2)2), 1.43 (s, 9H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 151.56 (MeO–C of Ar), 150.92 (HO–C of Ar), 137.53 (tBu–C of Ar), 123.10 (NCH2C of Ar), 112.40 (ortho-C of NCH2Ar), 110.59 (para-C of NCH2Ar), 59.54 (OCH3), 55.73 (NCH2Ar), 53.33 (N(CH2CH2)2), 34.90 (C(CH3)3), 29.48 (C(CH3)3), 23.83 (N(CH2CH2)2). Anal. calc. (found) for C16H25NO2: C 72.97 (72.66), H 9.57 (9.49), N 5.32 (5.21).

Synthesis of ONBr–OH

This compound was prepared following the same procedure described for ONOMe–H by using 2-(tert-butyl)-4-bromo-phenol was used in place of 3-tert-butyl-4-hydroxyanisole. Yield: 5.10 g (82%). 1H NMR (400 MHz, CDCl3) δ 7.25 (d, 1H, J = 2 Hz, para-H of PhtBu), 6.96 (d, 1H, J = 2 Hz, ortho-H of PhtBu), 3.76 (s, 2H, ArCH2N), 2.61 (br, 4H, N(CH2CH2)2), 1.85 (br, 4H, N(CH2CH2)2), 1.39 (s, 9H, C(CH3)3). 13C NMR (100 MHz, CDCl3, δ): 156.56 (HO–C of Ar), 138.71 (tBu–C of Ar), 128.61 (para-C of NCH2Ar), 128.35 (ortho-C of NCH2Ar), 124.52 (NCH2C of Ar), 110.11 (Br–C of Ar), 58.67 (NCH2Ar), 53.18 (N(CH2CH2)2), 34.84 (C(CH3)3), 29.21 (C(CH3)3), 23.72 (N(CH2CH2)2). C15H22BrNO: C 57.70 (57.58), H 7.10 (6.88), N 4.49 (4.19).

Synthesis of BuONNOBr–OH

In a 100 mL round-bottom flask, 4-bromo-2-tert-butylphenol (4.56 g, 20.0 mmol), paraformaldehyde (37 wt% in water, 2 mL), and N,N′-dimethylethane-1,2-diamine (0.88 g, 10.0 mmol) were added and dissolved in 30 mL of EtOH. The mixture was refluxed, and the reaction was monitored by TLC, which revealed that the reaction was complete after 8 h. The mixture was then warmed to room temperature, and the volatiles were removed under vacuum. The white powder can be obtained by recrystallization in hexane. Yield: 3.40 g (60%). 1H NMR (400 MHz, CDCl3) δ 10.92 (s, 2H, OH), 7.25 (d, 1H, J = 2 Hz, para-H of PhtBu), 6.93 (br, 1H, ortho-H–ArtBu), 3.62 (s, 4H, PhCH2N), 2.58 (br, 4H, NCH2CH2N), 2.23 (s, 6H, NCH3), 1.34 (s, 18H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 155.91 (HO–C of Ar), 139.15 (C of Ar–CtBu), 129.12 (ortho-C of ArtBu), 129.02 (ortho-C of NCH2Ph), 123.69 (NCH2C of Ar), 110.50 (Br–C of Ar), 61.81 (NCH2Ar), 53.30 (NCH2CH2N), 41.31 (NCH3), 34.88 (C(CH3)3), 29.19 (C(CH3)3). Anal. calc. (found) for C26H38Br2N2O2: C 54.75 (54.61), H 6.72 (6.57), N 4.91 (4.84).

Synthesis of OOH–OH

A stirred solution of 2-(tert-butyl) phenol (4.5 g, 30 mmol) in 60 mL of diethyl ether under nitrogen was cooled to 0 °C. To this solution, methyl magnesium bromide (10.0 mL, 3.0 M in diethyl ether) was slowly added. After gas evolution had ceased, the solvent was removed in vacuo. The resulting residue was dissolved in 150 mL of toluene. 4-Chlorobenzaldehyde (2.1 g, 15 mmol) was added, and the reaction mixture was refluxed for 12 h. After cooling to room temperature, the reaction mixture was washed with saturated NH4Cl. The layers were separated, and the aqueous layer was extracted using diethyl ether. The combined organic extracts were dried with MgSO4, and the volatile materials were removed at reduced pressure. Yielded: 1.0 g (42%) of OOH–H. 1H NMR (400 MHz, CDCl3) δ 7.32 (d, 2H, J = 8 Hz, ortho-H of PhCl), 7.28 (d, 2H, J = 2 Hz, para-H of PhtBu), 7.12 (d, 2H, J = 8 Hz, meta-H of PhCl), 6.84 (t, 2H, J = 8 Hz, para-H of Ar–OH), 6.67 (d, 2H, J = 2 Hz, ortho-H of PhtBu), 5.71 (s, 1H, CH(ArOH)2), 4.94 (s, 2H, OH), 1.40 (s, 18H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 152.89 (HO–C of Ar), 139.61 (para-C of ArCl), 137.62 (tBu–C of ArtBu), 133.28 (Cl–C), 130.93 (meta-C of ArCl), 129.20 (ortho-C of ArCl), 128.36 (CH–C of ArOH), 127.61 (para-C of ArtBu), 126.59 (para-C of ArOH), 120.87 (para-C of CHAr), 45.66 (CH(ArOH)2), 34.73 (C(CH3)3), 29.95 (C(CH3)3). Anal. calc. (found) for C27H31ClO2: C 76.67 (76.56), H 7.39 (7.22).

Synthesis of OOBr–OH

A solution of OOH–H (4.2 g, 10 mmol) in methanol (50 mL) was stirred at 0 °C and treated dropwise with neat Br2 (3.2 g, 20 mmol). After 30 min, the cooling bath was removed, and the mixture was kept at 25 °C for 12 h. The mixture was dissolved in ethyl acetate and washed with water and brine, and volatiles were removed from the organic phase by evaporation under reduced pressure. Flash chromatography (silica, hexane) of the residue provided an orange crystalline solid of OOBr–H (2.95 g, 51%). 1H NMR (400 MHz, CDCl3) δ 7.36 (d, 2H, J = 2 Hz, para-H of PhtBu), 7.34 (d, 2H, J = 8 Hz, ortho-H of PhCl), 7.07 (d, 2H, J = 8 Hz, meta-H of PhCl), 6.72 (d, 2H, J = 2 Hz, para-H of PhCH), 5.66 (s, 1H, CH(ArOH)2), 4.87 (s, 2H, OH), 1.38 (s, 18H, C(CH3)3). 13C NMR (100 MHz, CDCl3, δ): 151.62 (HO–C of Ar), 139.76 (tBu–C of Ar), 138.20 (para-C of ArCl), 133.86 (Cl–C), 130.37 (meta-C of ArCl), 129.98 (ortho-C of ArCl), 129.83 (CH–C of ArOH), 129.54 (para-C of ArtBu), 113.75 (Br–C), 44.78 (CH(ArOH)2), 34.86 (C(CH3)3), 29.78 (C(CH3)3). Anal. calc. (found) for C27H29Br2ClO2: C 55.84 (55.51), H 5.03 (4.77).

Synthesis of OOOMe–OH

A method similar to that for OOH–H was used except that 2-(tert-butyl)-4-methoxyphenyl was used in place of 2-(tert-butyl) phenol. Yield: 5.42 g (75%). 1H NMR (400 MHz, CDCl3) δ 7.31 (d, 2H, J = 8 Hz, ortho-H of PhCl), 7.12 (d, 2H, J = 8 Hz, meta-H of PhCl), 6.84 (d, 2H, J = 2 Hz, para-H of PhtBu), 6.72 (d, 2H, J = 2 Hz, para-H of PhCH), 5.65 (s, 1H, CH(ArOH)2), 4.47 (s, 2H, OH), 3.62 (s, 6H, OCH3), 1.37 (s, 18H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 153.47 (MeO–C of Ar), 146.69 (HO–C of Ar), 139.46 (tBu–C of ArtBu), 139.26 (para-C of ArCl), 133.39 (Cl–C), 130.89 (meta-C of ArCl), 129.38 (ortho-C of ArCl), 129.24 (CH–C of ArOH), 112.93 (para-C of ArtBu), 111.85 (ortho-C of ArOMe, ArOH), 55.56 (OCH3), 46.29 (CH(ArOH)2), 34.97 (C(CH3)3), 29.82 (C(CH3)3). Anal. calc. (found) for C29H35ClO4: C 72.11 (71.88), H 7.30 (7.05).

Synthesis of OOBu–OH

A method similar to that for OOH–H was used except that 2-(tert-butyl)-4-methoxyphenol was used instead of 2,4-di-tert-butylphenol. Yield: 5.42 g (73%). 1H NMR (400 MHz, CDCl3) δ 7.33 (d, 2H, J = 8 Hz, ortho-H of PhCl), 7.30 (d, 2H, J = 8 Hz, meta-H of PhCl), 7.11 (d, 2H, J = 2 Hz, para-H of PhCH), 6.63 (d, 2H, J = 2 Hz, para-H of PhtBu), 5.63 (s, 1H, CH(ArOH)2), 4.75 (s, 2H, OH), 1.38, 1.16 (s, 36H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 150.45 (HO–C of Ar), 143.05, 136.74 (tBu–C of ArtBu), 139.86 (para-C of ArCl), 136.74 (C of ArtBu, ortho-C of ArOH), 133.08 (Cl–C), 130.93 (meta-C of ArCl), 129.05 (ortho-C of ArCl), 127.61 (CH–C of ArOH), 124.39 (para-C of ArtBu, meta-C of ArOH), 123.26 (ortho-C of ArtBu, meta-C of ArOH), 46.48 (CH(ArOH)2), 34.95, 34.44 (C(CH3)3), 31.55, 29.99 (C(CH3)3). Anal. calc. (found) for C35H47ClO2: C 78.55 (78.23), H 8.85 (8.49).

Synthesis of NNOOMe–OH

One equivalent of benzaldehyde and one equivalent of N1,N1-dimethylethane-1,2-diamine were dissolved in ethanol and refluxed to form an imine compound, and sodium borohydride was used to reduce the imine compound to form an N-benzyl-N′,N′-dimethylethane-1,2-diamine ligand. The stirred solution of formaldehyde, 2-(tert-butyl)-4-methoxyphenol, and N-benzyl-N′,N′-dimethylethane-1,2-diamine in ethanol was refluxed for 1 h, and sodium borohydride was transferred and refluxed for one day. After the reaction was completed, the resulting mixture was cooled at room temperature, and water was added dropwise to quench the reaction and DCM. After stirring for 1 h, more water was added, and the organic layer was separated. The aqueous layer was extracted by DCM twice. The organic layer was dried over MgSO4 and the solvent was removed by evaporation to produce the colorless gel product. The pure product was obtained using column chromatography (10/1 hexane/ethyl acetate). Yield: 1.30 g (65%). 1H NMR (400 MHz, CDCl3) δ 7.30–7.21 (m, 5H, H–Ph), 6.79 (d, J = 2.0 Hz, 1H, ortho-H of tBuAr), 6.45 (d, J = 2.0 Hz, 1H, para-H of tBuAr), 3.73 (s, 3H, OCH3), 3.71 (s, 2H, ArCH2N), 3.56 (s, 2H, PhCH2N), 2.56 (t, J = 8.0 Hz, 2H, NCH2CH2NMe2), 2.43 (t, J = 8.0 Hz, 2H, NCH2CH2NMe2), 2.10 (s, 6H, N(CH3)2), 1.43 (s, 9H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 151.63 (C of MeO–Ar), 150.49 (C of HO–Ar), 138.01 (ipso-C of CH2Ph), 137.83 (C of tBuAr), 129.64 (ortho-C of CH2Ph), 128.40 (meta-C of CH2Ph), 127.40 (para-C of CH2Ar), 123.30 (C of CH2Ar), 112.81 (ortho-C of tBuAr), 111.98 (para-C of tBuAr), 58.45 (NCH2Ph), 57.95 (NCH2Ar), 56.53 (OCH3–Ar), 55.82 (NCH2CH2N(CH3)2), 50.46 (NCH2CH2N(CH3)2), 45.42 (N(CH3)2), 35.01 (C(CH3)3), 29.49 (C(CH3)3). Anal. calc. (found) for C23H34N2O2: C 74.55 (74.31), H 9.25 (9.21), N 7.56 (7.50).

Synthesis of NNOBr–OH

A method similar to that used for NNOOMe–OH, except that 2-tert-buty-4-bromo-phenol was used instead of 2-tert-butyl-4-methoxyphenol. Yield: 1.10 g (60%). 1H NMR (400 MHz, CDCl3) δ 7.28–7.20 (m, 5H, H–PhCH2), 7.22 (d, J = 2.0 Hz, 1H, ortho-H of tBuAr), 7.01 (d, J = 2.0 Hz, 1H, para-H of tBuAr), 3.66 (s, 2H, ArCH2N), 3.53 (s, 2H, PhCH2N), 2.54 (t, J = 8.0 Hz, 2H NCH2CH2N(CH3)2), 2.44 (t, J = 8.0 Hz, 2H, NCH2CH2N(CH3)2), 2.10 (s, 6H, N(CH3)2), 1.42 (s, 9H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 155.76 (C of HO–Ar), 139.16 (ipso-C of Ph), 137.67 (C of tBuAr), 128.96 (ortho-C of tBuAr), 129.48 (ortho-C of CH2Ph), 128.96 (para-C of tBuAr), 128.30 (meta-C of CH2Ph), 127.35 (para-C of CH2Ph), 127.35 (C of CH2Ar), 110.06 (Br–C of Ar), 58.30 (NCH2Ph), 56.46 (NCH2Ar), 56.11 (NCH2CH2N(CH3)2), 49.96 (NCH2CH2N(CH3)2), 45.10 (N(CH3)2), 35.00 (C(CH3)3), 29.25 (C(CH3)3). Anal. calc. (found) for C22H31BrN2O: C 63.00 (62.89), H 7.45 (7.41), N 6.68 (6.50).

Synthesis of ONOOMe–Al

A mixture of ONOOMe–OH (2.45 g, 5.0 mmol) and trimethylaluminum (2.5 mL, 2 M in toluene) in toluene (30 mL) was stirred for 5 h at room temperature. After ONOOMe–OH disappeared, as monitored by 1H NMR spectroscopy, volatile materials were removed under vacuum to obtain the yellow mud. Hexane (40 mL) was used to wash the yellow mud to produce a white powder. Yield: 2.12 g (80%). 1H NMR (400 MHz, CDCl3) δ 7.50–7.41 (m, 5H, meta-H + para-H + ortho-H of Ph), 6.88 (d, 2H, J = 2 Hz, para-H of PhtBu), 6.39 (d, 2H, J = 2 Hz, ortho-H of PhtBu), 4.02 (s, 2H, PhCH2N), 3.72 (s, 6H, OCH3), 3.67 (s, 4H, ArCH2N), 1.41 (s, 18H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 151.95 (MeO–C of Ar), 151.24 (HO–C of Ar), 140.84 (tBu–C of ArtBu), 132.86 (NCH2C of Ar), 129.67 (NCH2C of Ph), 128.95 (ortho-C of NCH2Ph), 128.73 (meta-C of NCH2Ph), 121.40 (para-C of NCH2Ph), 114.92 (ortho-C of NCH2Ar), 112.87 (para-C of NCH2Ar), 56.14 (OCH3), 56.10 (NCH2Ph), 55.02 (NCH2Ar), 34.41 (C(CH3)3), 29.72 (C(CH3)3). Anal. calc. (found) for C32H42AlNO4: C 72.29 (75.52), H 7.96 (8.25), N 2.63 (2.68).

Synthesis of ONOBr–Al

ONOBr–Al was prepared following the same procedure as described for ONOOMe–Al, except that ONOBr–OH was used instead of ONOOMe–OH. Yield: 2.50 g (80%). 1H NMR (400 MHz, CDCl3) δ 7.53–7.50 (br, 3H, para-H + ortho-H of Ph), 7.35–7.34 (br, 4H, meta-H of PhC + ortho-H–ArtBu), 6.95 (d, 2H, J = 2 Hz, ortho-H–ArtBu), 3.99 (s, 2H, PhCH2N), 3.65 (br, 4H, ArCH2N), 1.38 (s, 18H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 157.00 (HO–C of Ar), 142.22 (C of Ar–CtBu), 132.64 (ortho-C of NCH2Ph), 131.19 (NCH2C of Ar), 130.75 (meta-C of NCH2Ph), 130.00 (NCH2C of Ph), 129.19 (para-C of NCH2Ar), 128.03 (para-C of NCH2Ph), 123.19 (ortho-C of NCH2Ar), 109.96 (Br–C of Ar), 56.17 (NCH2Ph), 54.26 (NCH2Ar), 35.43 (C(CH3)3), 29.57 (C(CH3)3). Anal. calc. (found) for C30H36AlBr2NO2: C 57.25 (56.89), H 5.77 (5.56), N 2.23 (2.01).

Synthesis of ONOMe–Al

A mixture of ONOMe–OH (2.63 g, 5.0 mmol) and trimethylaluminum (2.5 mL, 2 M in toluene) in toluene (15 mL) was stirred for 3 h at room temperature. After ONOMe–OH disappeared, as monitored by 1H NMR spectroscopy, volatile materials were removed under a vacuum to obtain the white powder. Yield: 3.10 g (97%). 1H NMR (400 MHz, CDCl3) δ 6.87 (d, 1H, J = 2 Hz, para-H of PhtBu), 6.41 (d, 1H, J = 2 Hz, ortho-H of PhtBu), 3.77 (s, 2H, ArCH2N), 3.74 (s, 3H, OCH3), 3.15, 2.62 (br, 4H, N(CH2CH2)2), 1.95 (br, 4H, N(CH2CH2)2), 1.39 (s, 9H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 153.31 (HO–C of Ar), 150.08 (MeO–C of Ar), 140.08 (NCH2C of Ar), 122.19 (tBu–C of ArtBu), 114.74 (ortho-C of NCH2Ar), 112.48 (para-C of NCH2Ar), 60.11 (OCH3), 56.57 (NCH2Ar), 54.78 (N(CH2CH2)2), 35.31 (C(CH3)3), 29.72 (C(CH3)3), 22.96 (N(CH2CH2)2). Anal. calc. (found) for C18H30AlNO2: C 67.68 (67.41), H 9.47 (9.25), N 4.39 (4.11).

Synthesis of ONBr–Al

ONBr–Al was prepared following the same procedure described for ONOMe–Al, except that ONBr–OH was used instead of ONOMe–OH. Yield: 3.42 g (93%). 1H NMR (400 MHz, CDCl3) δ 7.30 (d, 1H, J = 2 Hz, para-H of PhtBu), 6.95 (d, 1H, J = 2 Hz, ortho-H of PhtBu), 3.74 (s, 2H, ArCH2N), 3.13, 2.62 (br, 4H, N(CH2CH2)2), 1.96 (br, 4H, N(CH2CH2)2), 1.37 (s, 9H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 158.26 (HO–C of Ar), 141.71 (tBu–C of ArtBu), 130.52 (para-C of NCH2Ar), 129.88 (ortho-C of NCH2Ar), 124.24 (NCH2C of Ar), 108.11 (Br–C of Ar), 59.32 (NCH2Ar), 54.85 (N(CH2CH2)2), 35.33 (C(CH3)3), 29.57 (C(CH3)3), 22.97 (N(CH2CH2)2). C17H27AlBrNO: C 55.44 (55.11), H 7.39 (7.21), N 3.80 (3.57).

Synthesis of ONNOOMe–Al

A mixture of ONNOOMe–OH (1.80 g, 5.0 mmol) and trimethylaluminum (2.5 mL, 2 M in toluene) in toluene (30 mL) was stirred for 3 h at 45 °C. After ONNOOMe–OH disappeared, as monitored by 1H NMR spectroscopy, volatile materials were removed under vacuum to obtain the white mud. Hexane (40 mL) was used to wash the white mud to produce white powder. Yield: 1.50 g (75%). 1H NMR (400 MHz, CDCl3) δ 6.81–6.73 (m, 4H, para-H + meta-H of NCH2Ar), 6.51 (d, 2H, J = 2 Hz, ortho-H of NCH2Ar), 4.26 (br, 2H, ArCH2N), 3.73 (s, 6H, OCH3), 3.25–2.70 (br, 4H, ArCH2N + N(CH2CH2)2), 2.78–2.68 (br, 2H, N(CH2CH2)2), −0.84 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 153.96 (AlO–C of Ar), 150.84 (MeO–C of Ar), 121.97 (NCH2C of Ar), 120.24 (ortho-C of ArO), 115.17 (ortho-C of NCH2Ar), 114.72 (para-C of NCH2Ar, ortho-C of ArOMe), 63.48 (OCH3), 61.67 (NCH2Ar), 56.10, 55.18 (N(CH2CH2)2), 44.44 (NCH3), −11.22 (CH3). Anal. calc. (found) for C21H29AlN2O4: C 62.99 (62.67), H 7.30 (7.21), N 7.00 (6.61).

Synthesis of ONNOBr–Al

ONNOBr–Al was prepared following the same procedure described for ONNOOMe–Al, except that ONNOBr–OH was used instead of ONNOOMe–OH. Yield: 1.98 g (80%). 1H NMR (400 MHz, CDCl3) δ 7.26 (d, 2H, J = 8 Hz, ortho-H of NCH2Ar), 7.02 (d, 2H, J = 2 Hz, ortho-H of NCH2Ar), 6.76 (br, 2H, meta-H of BrAr), 4.24, 3.20 (d, 4H, J = 14 Hz, ArCH2N), 3.45, 3.05, 2.83, 2.65 (br, 4H, NCH2CH2N), 2.32 (s, 6H, (NCH3)2), −0.83 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 158.98 (HO–C of Ar), 132.89 (ortho-C of ArO), 131.29 (NCH2C of Ar), 123.50 (ortho-C of OAr, meta-C of NCH2Ar), 121.89 (para-C of NCH2Ar, ortho-C of BrAr), 108.21 (Br–C of Ar), 62.82, 60.98 (N(CH2CH2)2), 55.09 (NCH2Ar), 44.63 (NCH3), −11.24 (CH3). Anal. calc. (found) for C19H23AlBr2N2O2: C 45.81 (45.49), H 4.65 (4.19), N 5.62 (5.31).

Synthesis of BuONNOOMe–Al

BuONNOOMe–Al was prepared following the same procedure described for ONNOOMe–Al, except that BuONNOOMe–OH was used instead of ONNOOMe–OH. Yield: 2.18 g (85%). 1H NMR (400 MHz, CDCl3) δ 6.89 (d, 2H, J = 2 Hz, ortho-H of tBuAr), 6.42 (d, 2H, J = 2 Hz, ortho-H of ArCH2), 4.23, 3.03 (d, 4H, J = 12 Hz, ArCH2N), 3.73 (s, 6H, OCH3), 3.03, 2.81 (m, 4H, NCH2CH2N), 2.23 (s, 6H, (NCH3)2), 1.45 (s, 18H, (CH3)3C), −0.89 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 152.99 (MeO–C of Ar), 149.68 (AlO–C of Ar), 140.09 ((CH3)3C–C of ArO), 121.76 (NCH2C of Ar), 123.50 (para-C of (CH3)3–CAr), 121.89 (para-C of NCH2Ar), 61.96 (CH3O), 55.93 (NCH2Ar), 55.32 (N(CH2CH2)2), 43.93 (NCH3), 35.16 ((CH3)3C), 29.79 ((CH3)3C), −11.56 (AlCH3). Anal. calc. (found) for C29H45AlN2O4: C 67.94 (67.66), H 8.85 (8.54), N 5.46 (5.12).

Synthesis of BuONNOBr–Al

BuONNOBr–Al was prepared following the same procedure described for ONNOOMe–Al, except that BuONNOBr–OH was used in place of ONNOOMe–OH. Yield: 1.07 g (84%). 1H NMR (400 MHz, CDCl3) δ 7.31 (d, 2H, J = 2 Hz, ortho-H of tBuAr), 6.94 (d, 2H, J = 2 Hz, ortho-H of ArCH2), 4.35, 3.05 (d, 4H, J = 12 Hz, ArCH2N), 3.02, 2.85 (m, 4H, NCH2CH2N), 2.23 (s, 6H, (NCH3)2), 1.41 (s, 18H, (CH3)3C), −0.87 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 157.91 (O–C of Ar), 141.63 ((CH3)3C–C of ArO), 130.12 (NCH2C of Ar), 129.42 (para-C of (CH3)3–CAr), 123.66 (para-C of NCH2Ar), 107.46 (Br–C of ArO), 61.13 (NCH2Ar), 55.18 (N(CH2CH2)2), 44.10 (NCH3), 35.18 ((CH3)3C), 29.62 ((CH3)3C), −11.54 (AlCH3). Anal. calc. (found) for C27H39AlBr2N2O2: C 53.13 (52.97), H 6.44 (6.42), N 4.59 (4.56).

Synthesis of OOOMe–Al

A mixture of OOOMe–OH (2.41 g, 5.0 mmol) and trimethylaluminum (2.5 mL, 2 M in toluene) in THF (20 mL) was stirred for 1 h at 0 °C and overnight at room temperature. After OOOMe–OH disappeared, as monitored by 1H NMR spectroscopy, volatile materials were removed under a vacuum to obtain the white powder. Yield: 2.43 g (82%). 1H NMR (400 MHz, CDCl3) δ 7.07 (d, 2H, J = 6 Hz, ortho-H of PhCl), 7.02 (d, 2H, J = 6 Hz, meta-H of PhCl), 6.62 (d, 2H, J = 2 Hz, para-H of PhtBu), 6.49 (d, 2H, J = 2 Hz, para-H of PhCH), 5.84 (s, 1H, CH(ArO)2), 4.07 (t, 4H, J = 4 Hz, O(CH2CH2)2), 3.55 (s, 6H, OCH3), 1.62 (br, 4H, O(CH2CH2)2), 1.27 (s, 18H, C(CH3)3) −0.70 (s, 3H, CH3). 13C NMR (100 MHz, CDCl3) δ 151.15 (MeO–C of Ar), 150.13 (AlO–C of Ar), 143.42 (para-C of ArCl), 139.46 (tBu–C of ArtBu), 132.63 (CH–C of ArO), 131.21 (Cl–C), 130.72 (ortho-C of ArCl), 127.75 (meta-C of ArCl), 111.91 (para-C of ArtBu), 111.14 (ortho-C of ArOMe, ArOAl), 71.63 (O(CH2CH2)2), 55.77 (OCH3), 41.32 (CH(ArO)2), 35.12 (C(CH3)3), 29.75 (C(CH3)3), 25.11 (O(CH2CH2)2), −14.04 (AlCH3). Anal. calc. (found) for C34H44AlClO5: C 68.62 (68.52), H 7.45 (7.63).

Synthesis of OOBu–Al

A method similar to that for OOOMe–Al was used, except that OOBu–OH was used instead of OOOMe–OH. Yield: 3.10 g (96%). 1H NMR (400 MHz, CDCl3) δ 7.18 (d, 2H, J = 8 Hz, ortho-H of PhCl), 7.14–7.11 (m, 4H, meta-H of PhCl + meta-H of AlOAr), 5.89 (s, 1H, CH(ArO)2), 4.24 (br, 4H, O(CH2CH2)2), 1.91 (br, 4H, O(CH2CH2)2) 1.41, 1.22 (s, 36H, C(CH3)3), −0.59 (s, 3H, AlCH3). 13C NMR (100 MHz, CDCl3) δ 153.50 (AlO–C of Ar), 144.77 (para-C of ArCl), 139.91, (tBu–C of ArtBu), 132.19 (C of ArtBu, ortho-C of ArOH), 137.17 (Cl–C), 130.97 (meta-C of ArCl), 130.85 (ortho-C of ArCl), 127.65 (CH–C of ArOAl), 124.26 (para-C of ArtBu, meta-C of ArO), 121.37 (ortho-C of ArtBu, meta-C of ArO), 71.61 (O(CH2CH2)2), 41.13 (CH(ArO)2), 35.29, 34.32 (C(CH3)3), 31.77, 30.13 (C(CH3)3), −13.95 (AlCH3). Anal. calc. (found) for C40H56AlClO3: C 74.22 (74.00), H 8.72 (8.65).

Synthesis of OOBr–Al

A method similar to that for OOOMe–Al was used, except that OOBr–OH was used in place of OOOMe–OH. Hexane (40 mL) was used to wash the white mud to give the product as a white powder. Yield: 2.76 g (80%). 1H NMR (400 MHz, CDCl3) δ 7.23 (d, 2H, J = 8 Hz, ortho-H of ClPh), 7.21 (d, 2H, J = 2 Hz, para-H of tBuAr), 7.11 (d, 2H, J = 2 Hz, para-H of CHAr), 7.07 (d, 2H, J = 8 Hz, meta-H of ClPh), 5.86 (s, 1H, CH(AlOAr)2), 4.16 (t, 4H, J = 8 Hz, O(CH2CH2)2), 1.98 (br, 4H, O(CH2CH2)2), 1.37 (s, 18H, C(CH3)3), −0.56 (s, 3H, AlCH3). 13C NMR (100 MHz, CDCl3) δ 154.87 (AlO–C of Ar), 141.12 (tBu–C of tBuAr), 142.16 (para-C of ClAr), 131.79 (Cl–C), 131.79 (meta-C of ClAr), 130.38 (ortho-C of ClAr), 134.40 (CH–C of ArOAl), 128.16 (para-C of CHAr), 128.13 (para-C of tBuAr), 110.73 (Br–C of Ar), 71.72 (O(CH2CH2)2), 40.71 (CH(ArO)2), 35.16 (C(CH3)3), 29.64 (C(CH3)3), −14.11 (AlCH3). Anal. calc. (found) for C32H38AlBr2ClO3: C 55.47 (55.67), H 5.53 (5.70).

Synthesis of OOMe–Al

A mixture of OOMe–OH (1.18 g, 5.0 mmol) and trimethylaluminum (2.5 mL, 2 M in toluene) in THF (50 mL) was stirred for 1 h at 0 °C and overnight at room temperature. After OOMe–OH disappeared, as monitored by 1H NMR spectroscopy, volatile materials were removed under vacuum to obtain the white powder. Yield: 1.46 g (80%). 1H NMR (400 MHz, CDCl3, δ): 6.81 (s, 2H, ortho-H of MeOAr), 4.20 (m, 4H, O(CH2CH2)2), 3.77 (s, 3H, OCH3), 2.08 (m, 4H, O(CH2CH2)2), 1.41 (s, 18H, C(CH3)3), −0.67 (s, 6H, CH3). 13C NMR (100 MHz, CDCl3) δ 151.43 (MeO–C of Ar), 149.87 (AlO–C of Ar), 139.31 (tBu–C of ArtBu), 110.82 (ortho-C of ArOMe), 71.24 (O(CH2CH2)2), 55.72 (OCH3), 35.17 (O(CH2CH2)2), 30.59 (C(CH3)3), 25.20 (C(CH3)3), −6.97 (AlCH3). Anal. calc. (found) for C21H37AlO3: C 69.20 (68.79), H 10.23 (9.75).

Synthesis of OBr–Al

A method similar to that for OOMe–Al was used, except that OBr–OH was used instead of OOMe–OH. Yield: 1.81 g (85%). 1H NMR (400 MHz, CDCl3) δ 7.26 (s, 2H, ortho-H of BrAr), 4.22 (m, 4H, O(CH2CH2)2), 2.10 (m, 4H, O(CH2CH2)2), 1.38 (s, 18H, C(CH3)3), −0.67 (s, 6H, CH3). 13C NMR (100 MHz, CDCl3) δ 156.44 (O–C of Ar), 141.12 (Br–C of Ar), 127.83 (tBu–C of ArtBu), 109.49 (ortho-C of BrAr), 71.13 (O(CH2CH2)2), 34.97 (O(CH2CH2)2), 30.36 (C(CH3)3), 25.13 (C(CH3)3), −7.13 (AlCH3). Anal. calc. (found) for C20H34AlBrO2: C 58.11 (57.93), H 8.29 (8.08).

Synthesis of NNOOMe–Al

NNOOMe–Al was prepared following the same procedure described for ONOMe–Al, except that NNOOMe–OH was used instead of ONOMe–OH. Yield: 1.15 g (78%). 1H NMR (400 MHz, CDCl3) δ 7.36–7.23 (m, 5H, Ph), 6.89 (d, J = 8.0 Hz, 1H, para-H of tBuAr), 6.32 (d, J = 2.0 Hz, 1H, meta-H of tBuAr), 3.92 (s, 2H, NCH2Ph), 3.78–3.66 (m, 5H, NCH2Ar + OCH3), 2.78 (br, 4H, NCH2CH2N), 2.25 (s, 6H, N(CH3)2), 1.40 (s, 9H, C(CH3)3), −0.70 (s, 6H, Al(CH3)2). 13C NMR (100 MHz, CDCl3) δ 153.30 (CH3O–C of Ar), 150.11 (HO–C of Ar), 140.22 (ipso-C of CH2Ar), 132.10 (ortho-C of tBuAr), 129.95 (ortho-C of CH2Ar), 129.30 (meta-C of CH2Ar), 128.65 (para-C of CH2Ar), 119.13 (ortho-C of tBuAr), 114.79 (meta-C of tBuAr), 112.34 (meta-C of tBuAr), 58.70 (OCH3), 56.11 (NCH2Ar), 53.20 (NCH2Ph), 52.88 (NCH2CH2N), 49.56 (NCH2CH2N), 45.71 (NCH3), 35.11 C(CH3)3, 29.33 C(CH3)3, −9.67 (Al(CH3)2). Anal. calc. (found) for C25H39AlN2O2: C 70.39 (70.12), H 9.22 (8.96), N 6.57 (6.26).

Synthesis of NNOBr–Al

NNOBr–Al was prepared following the same procedure described for NNOOMe–Al, except that NNOBr–Al was used instead of NNOOMe–OH. Yield: 1.13 g (79%). 1H NMR (400 MHz, CDCl3) δ 7.39–7.37 (m, 3H, Ph), 7.30 (d, J = 2.0 Hz, 1H, meta-H of tBuAr), 7.24–7.18 (m, 2H, Ph), 6.86 (d, J = 2.0 Hz, 1H, para-H of tBuAr), 3.90 (s, 2H, PhCH2N), 3.69 (s, 2H, ArCH2N), 2.76 (br, 4H, (NCH2CH2N)), 2.24 (s, 6H, N(CH3)2), 1.37 (s, 9H, C(CH3)3), −0.71 (s, 6H Al(CH3)2). 13C NMR (100 MHz, CDCl3) δ 158.34 (O–C of tBuAr), 141.53 (ipso-C of CH2Ph), 132.00 (C of tBuAr), 130.55 (ortho-C of tBuAr), 130.20 (meta-C of tBuAr), 129.69 (ortho-C of CH2Ph), 129.36 (meta-C of CH2Ph), 128.74 (para-C of CH2Ph), 121.69 (NCH2C of Ar), 107.73 (Br–C of Ar), 57.86 (NCH2Ar), 53.69 (NCH2Ph), 52.79 (NCH2CH2N), 50.43 (NCH2CH2N), 46.16 (N(CH3)2), 35.13 C(CH3)3, 29.19 C(CH3)3, −9.76 (Al(CH3)2). Anal. calc. (found) for C24H36AlBrN2O: C 60.63 (60.45), H 7.63 (7.54), N 5.89 (5.63).

Synthesis of (OOBu–AlOBn)2

BnOH (0.22 g, 0.2 mmol) was added to the solution of OOBu–Al (1.29 g, 2.0 mmol) in toluene (20 mL) at 0 °C for one day. The volatile materials were removed under vacuum to obtain the yellow powder. The powder was washed with hexane (30 mL) to obtain the pale yellow powder. Yield: 1.00 g (48%). 1H NMR (400 MHz, CDCl3) δ 7.27 (m, 4H, ortho-H of PhCl), 7.18 (d, 4H, J = 8 Hz meta-H of PhCl), 7.10, 7.01 (br, 10H, H–Bn), 6.94, 6.65 (br, 8H, H of ArO), 5.23 (s, 4H, CH2Ph), 5.11 (s, 2H, CH(ArO)2), 1.39, 1.17 (s, 36H, C(CH3)3). 13C NMR (100 MHz, CDCl3) δ 151.86 (AlO–C of Ar), 142.29 (para-C of ArCl), 141.23, 137.24, 131.31, 130.88, 130.29, 129.13, 128.32, 127.36, 125.50, 124.28, 121.80, 68.46 (OCH2Ph), 42.85 (CH(ArO)2), 35.38, 34.30 (C(CH3)3), 31.59, 30.44 (C(CH3)3). Anal. calc. (found) for C84H104Al2Cl2O6: C 75.60 (75.52), H 7.85 (7.53).

General procedures for the polymerization of ε-caprolactone

Synthesis of entry 1 (Table 2) illustrates a representative polymerization procedure using complex ONOOMe–Al as a catalyst. The polymerization conversion was examined by 1H NMR spectroscopy studies. Toluene (5.0 mL) was transferred to a mixture of complex ONOOMe–Al (0.1 mmol) and ε-caprolactone (10 mmol) at 25 °C. At indicated time intervals, 0.05 mL aliquots were removed, trapped with CDCl3 (1.0 mL), and analyzed by 1H NMR. After the solution was stirred for 85 min, the reaction was quenched by adding ethanol (10.0 mL), and the polymer precipitated as a white solid when poured into n-hexane (60.0 mL). The isolated white solid was dissolved in CH2Cl2 (10.0 mL), and water (10.0 mL) was used to wash the organic solution. Volatile materials were removed under vacuum to give a purified crystalline solid with a yield of 1.05 g (92%).

Author contributions

P. K. G. and F. H. performed the experiments; T. B. H. performed the additional edits and proofreading; R. K. prepared the graphical abstract; Y.-T. C. performed all crystallographic measurements; H.-C. T. performed the NMR experiments; S. D. performed additional edits and proofreading; K.-H. W. performed the DFT experiments; H.-Y. C. conceptualized the project and wrote and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Data availability

The authors confirm that the data supporting the findings of this study are available within the article and its ESI.

Conflicts of interest

The authors declare no conflicts of interest.

Acknowledgements

This study was supported by the National Science and Technology Council of Taiwan (Grant NSTC 113-2113-M-037-001, 111-2314-B-037-094-MY3, and 113-2113-M-037-009) and the Kaohsiung Medical University “NSYSU-KMU JOINT RESEARCH PROJECT” (NSYSU-KMU-113-P24 and KMU-DK109004). The authors thank the Center for Research Resources and Development at Kaohsiung Medical University for instrumentation and equipment support and the National Center for High-performance Computing (NCHC) of National Applied Research Laboratories (NARLabs) in Taiwan for providing computational and storage resources. The authors gratefully acknowledge the use of EA000600 of MOST 110-2731-M006-001, which belongs to the Core Facility Center of National Cheng Kung University.

References

  1. (a) R. Diez-Ahedo, X. Mendibil, M. C. Marquez-Posadas, I. Quintana, F. Gonzalez, F. J. Rodriguez, L. Zilic, C. Sherborne, A. Glen, C. S. Taylor, F. Claeyssens, J. W. Haycock, W. Schaafsma, E. Gonzalez, B. Castro and S. Merino, Polymers, 2020, 12, 971 CrossRef CAS PubMed ; (b) M. Abrisham, M. Noroozi, M. Panahi-Sarmad, M. Arjmand, V. Goodarzi, Y. Shakeri, H. Golbaten-Mofrad, P. Dehghan, A. Seyfi Sahzabi, M. Sadri and L. Uzun, Eur. Polym. J., 2020, 131, 109701–109716 CrossRef CAS ; (c) R. Al-Wafi, New J. Chem., 2022, 46, 17055–17065 RSC ; (d) S. Bhowmick, A. V. Thanusha, A. Kumar, D. Scharnweber, S. Rother and V. Koul, RSC Adv., 2018, 8, 16420–16432 RSC ; (e) A. Yadav, S. Ghosh, A. Samanta, J. Pal and R. K. Srivastava, Chem. Commun., 2022, 58, 1468–1480 RSC ; (f) K. Liu, X. Jiang and P. Hunziker, Nanoscale, 2016, 8, 16091–16156 RSC ; (g) C. G. Palivan, R. Goers, A. Najer, X. Zhang, A. Car and W. Meier, Chem. Soc. Rev., 2016, 45, 377–411 RSC ; (h) J. Y. Park, G. Gao, J. Jang and D. W. Cho, J. Mater. Chem. B, 2016, 4, 7521–7539 RSC .
  2. (a) M. Nevoralová, M. Koutný, A. Ujčić, Z. Starý, J. Šerá, H. Vlková, M. Šlouf, I. Fortelný and Z. Kruliš, Front. Mater., 2020, 7, 141 CrossRef ; (b) A. Nawaz, F. Hasan and A. A. Shah, FEMS Microbiol. Lett., 2014, 362, 1–7 CrossRef PubMed ; (c) A. Ohtaki, N. Sato and K. Nakasaki, Polym. Degrad. Stab., 1998, 61, 499–505 CrossRef CAS ; (d) H. Lu, S. A. Madbouly, J. A. Schrader, G. Srinivasan, K. G. McCabe, D. Grewell, M. R. Kessler and W. R. Graves, ACS Sustainable Chem. Eng., 2014, 2, 2699–2706 CrossRef CAS ; (e) T. Tábi, eXPRESS Polym. Lett., 2022, 16, 115–115 CrossRef ; (f) Y. Tokiwa, B. P. Calabia, C. U. Ugwu and S. Aiba, Int. J. Mol. Sci., 2009, 10, 3722–3742 CrossRef CAS .
  3. (a) M. Abedalwafa, F. Wang, L. Wang and C. Li, Rev. Adv. Mater. Sci., 2013, 34, 123–140 CAS ; (b) D. da Silva, M. Kaduri, M. Poley, O. Adir, N. Krinsky, J. Shainsky-Roitman and A. Schroeder, Chem. Eng. J., 2018, 340, 9–14 CrossRef CAS PubMed ; (c) E. M. Elmowafy, M. Tiboni and M. E. Soliman, J. Pharm. Invest., 2019, 49, 347–380 CrossRef CAS ; (d) Y. Ramot, M. Haim-Zada, A. J. Domb and A. Nyska, Adv. Drug Delivery Rev., 2016, 107, 153–162 CrossRef CAS ; (e) C. L. Salgado, E. M. Sanchez, C. A. Zavaglia and P. L. Granja, J. Biomed. Mater. Res., Part A, 2012, 100, 243–251 CrossRef ; (f) E. S. Permyakova, J. Polčak, P. V. Slukin, S. G. Ignatov, N. A. Gloushankova, L. Zajíčková, D. V. Shtansky and A. Manakhov, Mater. Des., 2018, 153, 60–70 CrossRef CAS .
  4. (a) K. Halász, Y. Hosakun and L. Csóka, Int. J. Polym. Sci., 2015, 2015, 1–6 CrossRef ; (b) J. Bota, M. Vukoje, M. Brozović and Z. Hrnjak-Murgić, Chem. Biochem. Eng. Q., 2018, 31, 417–424 CrossRef ; (c) A. Bhatia, R. K. Gupta, S. N. Bhattacharya and H. J. Choi, J. Nanomater., 2012, 2012, 1–11 Search PubMed ; (d) G. Colomines, V. Ducruet, C. Courgneau, A. Guinault and S. Domenek, Polym. Int., 2010, 59, 818–826 CrossRef CAS .
  5. (a) X.-G. Yang and L.-J. Liu, Polym. Bull., 2008, 61, 177–188 CrossRef CAS ; (b) S. Inkinen, M. Hakkarainen, A. C. Albertsson and A. Sodergard, Biomacromolecules, 2011, 12, 523–532 CrossRef CAS .
  6. (a) O. Santoro and C. Redshaw, Catalysts, 2020, 10, 210 CrossRef CAS ; (b) O. Santoro, X. Zhang and C. Redshaw, Catalysts, 2020, 10, 800 CrossRef CAS ; (c) M. Strianese, D. Pappalardo, M. Mazzeo, M. Lamberti and C. Pellecchia, Dalton Trans., 2020, 49, 16533–16550 RSC ; (d) H. Li, R. M. Shakaroun, S. M. Guillaume and J. F. Carpentier, Chem. – Eur. J., 2020, 26, 128–138 CrossRef CAS ; (e) S. Kaler and M. D. Jones, Dalton Trans., 2022, 51, 1241–1256 RSC ; (f) D. M. Lyubov, A. O. Tolpygin and A. A. Trifonov, Coord. Chem. Rev., 2019, 392, 83–145 CrossRef CAS ; (g) L.-J. Wu, W. Lee, P. Kumar Ganta, Y.-L. Chang, Y.-C. Chang and H.-Y. Chen, Coord. Chem. Rev., 2023, 475, 21487–21509 CrossRef ; (h) M. Chen and C. Chen, Chin. J. Chem., 2020, 38, 282–286 CrossRef CAS .
  7. (a) A. M. McCollum, A. M. Longo, A. E. Stahl, A. S. Butler, A. L. Rheingold, T. R. Cundari, D. B. Green, K. R. Brereton and J. M. Fritsch, Polyhedron, 2021, 204, 115233–115245 CrossRef CAS ; (b) E. N. Cooper, B. Averkiev, V. W. Day and P. E. Sues, Organometallics, 2021, 40, 3185–3200 CrossRef CAS ; (c) S. R. Kosuru, Y. L. Chang, P. Y. Chen, W. Lee, Y. C. Lai, S. Ding, H. Y. Chen, H. Y. Chen and Y. C. Chang, Inorg. Chem., 2022, 61, 3997–4008 CrossRef CAS ; (d) S. R. Kosuru, F. J. Lai, Y. L. Chang, C. Y. Li, Y. C. Lai, S. Ding, K. H. Wu, H. Y. Chen and Y. H. Lo, Inorg. Chem., 2021, 60, 10535–10549 CrossRef CAS PubMed ; (e) S. Gesslbauer, G. Hutchinson, A. J. P. White, J. Burés and C. Romain, ACS Catal., 2021, 11, 4084–4093 CrossRef CAS ; (f) E. Yue, F. Cao, J. Zhang, W. Zhang, Y. Jiang, T. Liang and W. H. Sun, RSC Adv., 2021, 11, 13274–13281 RSC ; (g) X. Zhang, K. Chen, M. Chicoma, K. Goins, T. Prior, T. Nile and C. Redshaw, Catalysts, 2021, 11, 1090 CrossRef CAS .
  8. X. Pang, H. Du, X. Chen, X. Wang and X. Jing, Chem. – Eur. J., 2008, 14, 3126–3136 CrossRef CAS .
  9. (a) N. Nomura, T. Aoyama, R. Ishii and T. Kondo, Macromolecules, 2005, 38, 5363–5366 CrossRef CAS ; (b) J. Liu, N. Iwasa and K. Nomura, Dalton Trans., 2008, 3978–3988 RSC ; (c) N. Iwasa, M. Fujiki and K. Nomura, J. Mol. Catal. A: Chem., 2008, 292, 67–75 CrossRef CAS ; (d) D. Pappalardo, L. Annunziata and C. Pellecchia, Macromolecules, 2009, 42, 6056–6062 CrossRef CAS ; (e) N. Iwasa, S. Katao, J. Liu, M. Fujiki, Y. Furukawa and K. Nomura, Organometallics, 2009, 28, 2179–2187 CrossRef CAS .
  10. W. H. Sun, M. Shen, W. Zhang, W. Huang, S. Liu and C. Redshaw, Dalton Trans., 2011, 40, 2645–2653 RSC .
  11. H. C. Tseng, M. Y. Chiang, W. Y. Lu, Y. J. Chen, C. J. Lian, Y. H. Chen, H. Y. Tsai, Y. C. Lai and H. Y. Chen, Dalton Trans., 2015, 44, 11763–11773 RSC .
  12. W. A. Ma, L. Wang and Z. X. Wang, Dalton Trans., 2011, 40, 4669–4677 RSC .
  13. F. Qian, K. Liu and H. Ma, Dalton Trans., 2010, 39, 8071–8083 RSC .
  14. P. Wang, J. Chao and X. Chen, Dalton Trans., 2018, 47, 4118–4127 RSC .
  15. M. Lamberti, I. D'Auria, M. Mazzeo, S. Milione, V. Bertolasi and D. Pappalardo, Organometallics, 2012, 31, 5551–5560 CrossRef CAS .
  16. (a) L. M. Alcazar-Roman, B. J. O'Keefe, M. A. Hillmyer and W. B. Tolman, Dalton Trans., 2003, 3082–3087,  10.1039/b303760f ; (b) K. Ding, M. O. Miranda, B. Moscato-Goodpaster, N. Ajellal, L. E. Breyfogle, E. D. Hermes, C. P. Schaller, S. E. Roe, C. J. Cramer, M. A. Hillmyer and W. B. Tolman, Macromolecules, 2012, 45, 5387–5396 CrossRef CAS .
  17. Z. Tang and V. C. Gibson, Eur. Polym. J., 2007, 43, 150–155 CrossRef CAS .
  18. (a) J. Liu and H. Ma, J. Polym. Sci., Part A:Polym. Chem., 2014, 52, 3096–3106 CrossRef CAS ; (b) J. Liu, N. Iwasa and K. Nomura, Dalton Trans., 2008, 3978–3988 RSC ; (c) P. Hormnirun, E. L. Marshall, V. C. Gibson, A. J. White and D. J. Williams, J. Am. Chem. Soc., 2004, 126, 2688–2689 CrossRef CAS ; (d) N. Iwasa, J. Liu and K. Nomura, Catal. Commun., 2008, 9, 1148–1152 CrossRef CAS ; (e) P. Wang, J. Chao and X. Chen, Dalton Trans., 2018, 47, 4118–4127 RSC ; (f) Y. Liu, W. S. Dong, J. Y. Liu and Y. S. Li, Dalton Trans., 2014, 43, 2244–2251 RSC ; (g) Z. Qu, R. Duan, X. Pang, B. Gao, X. Li, Z. Tang, X. Wang and X. Chen, J. Polym. Sci., Part A:Polym. Chem., 2014, 52, 1344–1352 CrossRef CAS ; (h) H. Du, A. H. Velders, P. J. Dijkstra, J. Sun, Z. Zhong, X. Chen and J. Feijen, Chem. – Eur. J., 2009, 15, 9836–9845 CrossRef CAS PubMed ; (i) C. Bakewell, R. H. Platel, S. K. Cary, S. M. Hubbard, J. M. Roaf, A. C. Levine, A. J. P. White, N. J. Long, M. Haaf and C. K. Williams, Organometallics, 2012, 31, 4729–4736 CrossRef CAS ; (j) X. Pang, R. Duan, X. Li and X. Chen, Polym. Chem., 2014, 5, 3894–3900 RSC ; (k) S. Gong and H. Ma, Dalton Trans., 2008, 3345–3357 RSC ; (l) J. Liu and H. Ma, Dalton Trans., 2014, 43, 9098–9110 RSC .
  19. P. Sumrit and P. Hormnirun, Macromol. Chem. Phys., 2013, 214, 1845–1851 CrossRef CAS .
  20. V. Belot, D. Farran, M. Jean, M. Albalat, N. Vanthuyne and C. Roussel, J. Org. Chem., 2017, 82, 10188–10200 CrossRef CAS PubMed .
  21. G. Ciancaleoni, N. Fraldi, P. H. M. Budzelaar, V. Busico and A. Macchioni, Organometallics, 2011, 30, 3096–3105 CrossRef CAS .
  22. (a) M.-L. Hsueh, B. H. Huang and C.-C. Lin, Macromolecules, 2002, 5763–5768 CrossRef CAS ; (b) M. T. Jiang, S. R. Kosuru, Y. H. Lee, W. Y. Lu, J. K. Vandavasi, Y. C. Lai, M. Y. Chiang and H. Y. Chen, eXPRESS Polym. Lett., 2018, 12, 126–135 CrossRef CAS ; (c) H.-Y. Chen, Y.-H. Lee, M. Y. Chiang, W.-Y. Lu, H.-C. Tseng, H.-Y. Tsai, Y.-H. Chen, Y.-C. Lai and H.-Y. Chen, RSC Adv., 2015, 5, 82018–82026 RSC .
  23. D. C. Crans, M. L. Tarlton and C. C. McLauchlan, Eur. J. Inorg. Chem., 2014, 2014, 4450–4468 CrossRef CAS .
  24. L. Yang, D. R. Powell and R. P. Houser, Dalton Trans., 2007, 955–964 RSC .
  25. H.-Y. Chen, L. Mialon, K. A. Abboud and S. A. Miller, Organometallics, 2012, 31, 5252–5261 CrossRef CAS .
  26. (a) Y. Yu, E. J. Fischer, G. Storti and M. Morbidelli, Ind. Eng. Chem. Res., 2014, 53, 7333–7342 CrossRef CAS ; (b) A. Duda, Z. Florjanczyk, A. Hofman, S. Slomkowski and S. Penczek, Macromolecules, 1990, 23, 1640–1646 CrossRef CAS .
  27. (a) M.-L. Hsueh, B.-H. Huang and C.-C. Lin, Macromolecules, 2002, 35, 5763–5768 CrossRef CAS ; (b) Y.-C. Liu, B.-T. Ko and C.-C. Lin, Macromolecules, 2001, 34, 6196–6201 CrossRef CAS ; (c) H.-L. Chen, B. T. Ko, B.-H. Huang and C.-C. Lin, Organometallics, 2001, 20, 5076–5083 CrossRef CAS .
  28. M. M. Kireenko, E. A. Kuchuk, K. V. Zaitsev, V. A. Tafeenko, Y. F. Oprunenko, A. V. Churakov, E. Lermontova, G. S. Zaitseva and S. S. Karlov, Dalton Trans., 2015, 44, 11963–11976 RSC .
  29. F. Hild, N. Neehaul, F. Bier, M. Wirsum, C. Gourlaouen and S. Dagorne, Organometallics, 2013, 32, 587–598 CrossRef CAS .
  30. M. Hu, F. Wang, F. Han, Q. Deng, W. Ma, H. Yan, G. Dong and W. Song, J. Polym. Sci., Part A:Polym. Chem., 2017, 55, 2084–2091 CrossRef CAS .
  31. M. Bouyahyi, T. Roisnel and J.-F. Carpentier, Organometallics, 2009, 29, 491–500 CrossRef .
  32. Y. N. Chang, P. Y. Lee, X. R. Zou, H. F. Huang, Y. W. Chen and L. C. Liang, Dalton Trans., 2016, 45, 15951–15962 RSC .
  33. P. Wang, X. Hao, J. Cheng, J. Chao and X. Chen, Dalton Trans., 2016, 45, 9088–9096 RSC .
  34. J. Payne, P. McKeown, G. Kociok-Kohn and M. D. Jones, Chem. Commun., 2020, 56, 7163–7166 RSC .
  35. S. Gesslbauer, H. Cheek, A. J. P. White and C. Romain, Dalton Trans., 2018, 47, 10410–10414 RSC .
  36. C. T. Altaf, H. Wang, M. Keram, Y. Yang and H. Ma, Polyhedron, 2014, 81, 11–20 CrossRef CAS .
  37. C. Romain, C. Fliedel, S. Bellemin-Laponnaz and S. Dagorne, Organometallics, 2014, 33, 5730–5739 CrossRef CAS .
  38. R. O. Macrae, C. M. Pask, L. K. Burdsall, R. S. Blackburn, C. M. Rayner and P. C. McGowan, Angew. Chem., Int. Ed., 2011, 50, 291–294 CrossRef CAS .
  39. I. Taden, H.-C. Kang, W. Massa, T. P. Spaniol and J. Okuda, Eur. J. Inorg. Chem., 2000, 441–445 CrossRef CAS .
  40. Y. C. Chen, C. Y. Lin, W. Y. Huang, A. Datta, J. H. Huang, M. S. Tsai, T. Y. Lee, C. Y. Tu and C. H. Hu, Eur. J. Inorg. Chem., 2011, 2011, 2459–2469 CrossRef .
  41. (a) H. Du, X. Pang, H. Yu, X. Zhuang, D. Cui, X. Wang and X. Jing, Macromolecules, 2007, 40, 1904–1913 CrossRef CAS ; (b) E. Rufino-Felipe, N. Lopez, F. A. Vengoechea-Gómez, L. G. Guerrero-Ramírez and M. Á. Muñoz-Hernández, Appl. Organomet. Chem., 2018, 32, e4315 CrossRef ; (c) T. M. Ovitt and G. W. Coates, J. Am. Chem. Soc., 2002, 124, 1316–1326 CrossRef CAS PubMed ; (d) J. Yang, Y. Yu, Q. Li, Y. Li and A. Cao, J. Polym. Sci., Part A:Polym. Chem., 2004, 43, 373–384 CrossRef ; (e) W. Luo, T. Shi, S. Liu, W. Zuo and Z. Li, Organometallics, 2017, 36, 1736–1742 CrossRef CAS .
  42. M. Oishi, Y. Ichinose, N. Iwata and N. Nomura, Organometallics, 2019, 38, 4233–4243 CrossRef CAS .
  43. W. Zhao, Q. Wang, Y. Cui, J. He and Y. Zhang, Dalton Trans., 2019, 48, 7167–7178 RSC .
  44. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision E.01, Wallingford, CT, 2009 Search PubMed.
  45. (a) J. Slaughter, S. A. Molyneux, A. J. Peel and A. E. H. Wheatley, Organometallics, 2019, 38, 395–408 CrossRef CAS ; (b) Q. Wang, W. Zhao, J. He, Y. Zhang and E. Y. X. Chen, Macromolecules, 2017, 50, 123–136 CrossRef CAS ; (c) A. P. Shreve, R. Mulhaupt, W. Fultz, J. Calabrese, W. Robbins and S. D. Ittel, Organometallics, 1988, 7, 409–416 CrossRef CAS .
  46. (a) S. S. Batsanov, Inorg. Mater., 2001, 37, 871–885 CrossRef CAS ; (b) M. Mantina, A. C. Chamberlin, R. Valero, C. J. Cramer and D. G. Truhlar, J. Phys. Chem. A, 2009, 113, 5806–5812 CrossRef CAS ; (c) A. Bondi, J. Phys. Chem., 1964, 68, 441–451 CrossRef CAS .
  47. P. Pyykkö and M. Atsumi, Chem. – Eur. J., 2009, 15, 186–197 CrossRef PubMed .
  48. (a) X. Pang, R. Duan, X. Li, Z. Sun, H. Zhang, X. Wang and X. Chen, Polym. Chem., 2014, 5, 6857–6864 RSC ; (b) Y. Wei, S. Wang, X. Zhu, S. Zhou, X. Mu, Z. Huang and D. Hong, Organometallics, 2016, 35, 2621–2629 CrossRef CAS ; (c) M. J. Stanford and A. P. Dove, Chem. Soc. Rev., 2010, 39, 486–494 RSC ; (d) S. Yoon, S. H. Kim, J. Heo and Y. Kim, Inorg. Chem. Commun., 2013, 29, 157–159 CrossRef CAS ; (e) X. Yang, L. Wang, L. Yao, J. Zhang, N. Tang, C. Wang and J. Wu, Inorg. Chem. Commun., 2011, 14, 1711–1714 CrossRef CAS ; (f) Q. Zhang, W. Zhang, N. M. Rajendran, T. Liang and W. H. Sun, Dalton Trans., 2017, 46, 7833–7843 RSC .
  49. (a) A. Rivera, J. J. Rojas, J. Salazar-Barrios, M. Maldonado and J. Rios-Motta, Molecules, 2010, 15, 4102–4110 CrossRef CAS PubMed ; (b) C. J. Whiteoak, G. J. Britovsek, V. C. Gibson and A. J. White, Dalton Trans., 2009, 2337–2344 RSC .
  50. M. Healy, D. A. Wlerda and A. R. Barron, Organometallics, 1988, 7, 2543–2548 CrossRef CAS .
  51. K. Nonoshita, H. Banno, K. Maruoka and H. Yamamoto, J. Am. Chem. Soc., 1990, 112, 316–322 CrossRef CAS .

Footnotes

Electronic supplementary information (ESI) available: Al complexes and polymer characterization data and details of the kinetic study. CCDC 2236052, 2247640, 2247643, 2247645 and 2247652. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4dt02923b
These authors contributed equally.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.