Next Article in Journal
A Numerical and Experimental Study of Adhesively-Bonded Polyethylene Pipelines
Next Article in Special Issue
Synthesis of Ultrahigh Molecular Weight Polymers Containing Reactive Functionality with Low PDIs by Polymerizations of Long-Chain α-Olefins in the Presence of Their Nonconjugated Dienes by Cp*TiMe2(O-2,6-iPr2C6H3)–Borate Catalyst
Previous Article in Journal
Effect of the Chain Structure of Self-Emulsifying Polyester Sizing Agent on ILSS of Carbon Fiber/Unsaturated Polyester Resin Composites
Previous Article in Special Issue
Biodegradable Oligoesters of ε-Caprolactone and 5-Hydroxymethyl-2-Furancarboxylic Acid Synthesized by Immobilized Lipases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanistic Insight into the Ring-Opening Polymerization of ε-Caprolactone and L-Lactide Using Ketiminate-Ligated Aluminum Catalysts

1
Department of Fragrance and Cosmetic Science, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
2
Department of Medicinal and Applied Chemistry, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
*
Author to whom correspondence should be addressed.
Polymers 2019, 11(9), 1530; https://doi.org/10.3390/polym11091530
Submission received: 14 August 2019 / Revised: 11 September 2019 / Accepted: 17 September 2019 / Published: 19 September 2019
(This article belongs to the Special Issue Catalytic Polymerization)

Abstract

:
The reactivity and the reaction conditions of the ring-opening polymerization of ε-caprolactone (ε-CL) and L-lactide (LA) initiated by aluminum ketiminate complexes have been shown differently. Herein, we account for the observation by studying the mechanisms on the basis of density functional theory (DFT) calculations. The calculations show that the ring-opening polymerization of ε-CL and LA are rate-determined by the benzoxide insertion and the C–O bond cleavage step, respectively. Theoretical computations suggest that the reaction temperature of L–LA polymerization should be higher than that of ε-CL one, in agreement with the experimental data. To provide a reasonable interpretation of the experimental results and to give an insight into the catalyst design, the influence of the electronic, steric, and thermal effects on the polymerization behaviors will be also discussed in this study.

Graphical Abstract

1. Introduction

Utilization of the polyester such as poly(ε-caprolactone) (PCL) and poly(L-lactide) (PLA) as eco-friendly materials has been paid more and more attention [1,2,3,4,5,6]. Therefore, development of efficient synthetic strategies becomes of great interest to chemists. Among these strategies, ring-opening polymerization (ROP) catalyzed by the less toxic and inexpensive aluminum complexes is one of the useful and well-controlled methods [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30]. The ring-opening polymerization catalyzed by an organometallic complex usually proceeds through a coordination-insertion mechanism [31,32]. Recent works have revealed that coordination number [33,34,35], the size effect of the chelate rings [36], steric and electronic effects of the auxiliary ligands surrounding the aluminum complexes [22,25,34,37,38,39,40,41,42,43,44] as well as the co-initiators would alter the reactivity and the properties of the products. Besides the reaction conditions would be very dependent on the ring size and the bulkiness of the monomers [45,46]. For example, L-lactide has been found to be bulkier than caprolactone since the Al–O bond enthalpy of the former are weaker than the later [47]. Several kinds of Al complexes as the initiators for the ring opening polymerization have been designed, and different factors that influence the reactivity have been discussed separately. Nevertheless, systematically studying the structures-activity relationship of the catalysts remains ongoing.
In 2015, Chen et al. reported a series of aluminum complexes bearing ketiminate ligands to process the ring-opening polymerization of both ε-caprolactone (ε-CL) and L-lactide (L-LA) (Figure 1) [48,49]. By screening the pendant groups on the ketiminate skeleton of the aluminum complexes, the steric, electronic and chelating effects of the ring-opening polymerization of ε-caprolactone and L-lactide could be delineated by comparison of the experimental results. Previous experimental data has revealed that in the ketiminate-ligated Al systems, ring-opening polymerization of ε-caprolactone can proceed at room temperature. Nevertheless, if using L-lactide as a monomer, similar reactions cannot occur until raising up temperature to 333 K. Besides, the kobs values of the ring-opening polymerization are influenced by the steric and electronic effects on the pendant N donors of kind of Al complex. It is worth noting that the orders of kobs values in the polymerization of ε-caprolactone and L-LA have different tendencies. For examples, catalysts with the bulky isopropyl substituents facilitates the polymerization of ε-caprolactone, but formation of polylactide is not promoted by this kind Al complex (Figure 2). To get more insight into the mechanism, we decided to employ density functional theory (DFT) calculation to compare the two ring-opening polymerization systems catalyzed by the Al complex bearing a ketiminate ligand. Although only few cases of tetradentate Al catalysts have been evaluated by DFT calculations [50,51,52], it has been proven to be a useful tool to model the reaction intermediates and comprehend the reaction pathways of the tris- and penta-coordinate Al complexes [29,42,44,52,53,54,55,56,57,58,59,60,61,62]. Our aims in this theoretical study are: (a) to realize why polymerization of ε-caprolactone and L-lactide require different reaction temperatures; (b) to understand how the steric and electronic effects on the pendant group of the Al complexes would alter the catalytic abilities. To answer the above questions, we tried to compute the entire reaction pathways and estimate the activation barriers of the rate-determining steps. For the thermodynamic consideration, we also evaluated the bond enthalpy for the monomer coordination to the Al center. In addition, we chose L, LCl, and LiPr (Figure 1) as pendant groups of the catalysts to evaluate the electronic and steric effects in both poly(ε-caprolactone) and poly(L-lactide) synthesis.

2. Materials and Methods

All the energies and the geometries for the intermediates and transition states were evaluated using the Gaussian 16 package (Gaussian Inc., Pittsburgh, PA, USA) [63]. The Becke 3-parameter Lee-Yang Parr (B3LYP) hybrid functional combined with the D3 empirical dispersion correction developed by Grimme and the 6–31g(d) basis set was chosen as the calculation level [64,65,66,67,68,69,70,71]. The polarizable continuum model (PCM) was employed to describe the solvent effect of toluene [72]. To identify all the stationary-point structures, all reaction species were subjected to perform the frequency computation. All the resting states show real values in all types of motional frequencies, while the species with only one imaginary frequency were confirmed as transition states. Finally, the intrinsic reaction coordinate (IRC) or Quasi-IRC calculations were employed to make sure the connection between intermediates and the transition state [73,74,75,76]. The optimized Cartesian coordinates for all structures are provided in the Supplementary Materials.

3. Results and Discussion

Ring-opening polymerization of ε-caprolactone and L-lactide catalyzed by Al complexes would follow the coordination-insertion mechanism. The process starts from monomer coordination to the Al center, followed by benzoxide insertion to the C=O group of the monomer. Then, the resultant intermediate adjusts to the proper conformation to facilitate the breakage of C–O bond, completing the first ring-opening process. In the ketiminate-ligated Al systems, the monomer is approaching the Al center to give a penta-coordinate species. Considering 13 possible isomers can be generated, all of them were submitted as initial geometries and optimized by DFT calculation. (Figure 3). Calculation results show that only isomer I, II, III, and IV can be obtained as the optimal geometries, so mechanisms starting from isomer V–XIII are not considered. In order to obtain a more comprehensive view of the ROP mechanism, we discussed the four possible pathways from four triangular bipyramidal (TBP) isomer I–IV as the starting species (Scheme 1). By scanning the reaction coordinates, we confirmed that for the species N donor resides in the axial site (isomer III and IV), the monomer tends to be driven away from the aluminum center when benzoxide is approaching. As a result, Route 3 and 4 in Scheme 1 can be ruled out. The mechanistic calculations were concentrated on the pathways starting from I and II, which are initiated from the coordination isomers with the axial O donor (Route 1 and 2: the starting species is the isomer that monomer respectively coordinates on the equatorial and axial site).

3.1. Ring-Opening Polymerization of ε-Caprolactone

Computing results showed that ε-caprolactone can occupy either equatorial or axial sites to give coordinated species in all three systems with different pendant groups. The ε-caprolactone equatorially and axially coordinated isomers are respectively denoted as Ceq1a–c and Cax1a-c (a: L, b: LCl, c: Lipr; Figure 4). The Al–O2 (Al–OCL) bond lengths in Ceq1a, Ceq1b, and Ceq1c are respectively 2.013, 2.012, and 2.024 Å, shorter than those in Cax1a–c (Cax1a and Cax1b: 2.069; Cax1c: 2.085 Å). However, the bond enthalpies of Al-caprolactone in Cax1a–c (Cax1a: −12.7; Cax1b: −13.4; Cax1c: −9.1 kcal/mol) are slightly larger than those in Ceq1a–c (Ceq1a: −9.4; Ceq1b: −10.4; Ceq1c: −7.5 kcal/mol), showing coordination of ε-caprolactone at the axial site is energetically favorable. Both the equatorial and axial caprolactone coordination orientation are affected by the steric effect. Bulky substituents of isopropyl groups would hinder the monomer coordination. As a result, the bond enthalpies in the LiPr systems are the smallest among the three systems. On the other hand, the fact the Al–O2 bond enthalpies in the LCl systems are largest can be attributed to the chloro substituent, whose electron withdrawing property would make the Al center more positive and favorably binding to the ε-caprolactone.
The Gibbs free energy profiles are shown in Figure 5, Figure 6 and Figure 7. The calculations reveal that the benzoxide insertion is the rate determining step of the ε-caprolactone ring-opening polymerization. The relative Gibbs energies in Route 1 are computed as 14.4 kcal/mol for CeqTS1a, 14.9 for CeqTS1b, and 13.2 for CeqTS1c (The differences in barrier height between Ceq1a–c and CeqTS1a–c are 10.3, 10.9, and 9.1 kcal/mol). On the other hand, the activation Gibbs energy barriers of CaxTS1a, CaxTS1b and CaxTS1c are 8.2, 8.3, and 7.6 kcal/mol, respectively. The distances of C1–O4 in CeqTS1a, CeqTS1b and CeqTS1c are 1.923, 1.921, and 1.886 Å, while they are 1.768, 1.769, and 1.715 Å in CaxTS1a, CaxTS1b and CaxTS1c. The shorter distances of C1–O4 in CeqTS1c and CaxTS1c are probably due to the steric congestion imposed by the isopropyl groups. The benzoxide insertion steps which give the products of Ceq2a–c (Route 1) and Cax2a–c (Route 2) were evaluated as endergonic reactions. Noteworthily, the LiPr systems of both routes require least input of energies to proceed the insertion step, probably because of the release of crowded space with the benzoxide group approaching to the ε-caprolactone [77]. The distances between Al and O4 of the inserted benzoxide in Ceq2a–c are similar (2.141, 2.148, and 2.147 Å for Ceq2a, Ceq2b and Ceq2c), while the aforementioned distance in Cax2c (2.047 Å) is longer than those in Cax2a (2.021 Å) and Cax2b (2.015 Å). After the benzoxide insertion, the Al-O2 bond in Ceq2a–c/Cax2a–c rotates to yield Ceq3a–c/Cax3a–c, which hold the structures in proper orientations to facilitate the C1-O3 bond breaking. Although Ceq2a–c/Ceq3a–c and Cax2a–c/Cax3a–c can be viewed as two pairs of linkage isomers, the computing result showed that the free energy differences between Ceq3a–c/Cax3a–c and Ceq2a–c/Cax2a–c are very dependent on their configurations and substituents. For the less bulky L and LCl ligand systems, the free energies of Ceq3a–b and Cax3a–b are more favored than those of Ceq2a–b and Cax2a–b with around 3 and 6.5 kcal/mol, respectively. For the LiPr ligand systems, 3.1 kcal/mol is required to drive Ceq2c to Ceq3c, while the pathway from Cax2c to Cax3c is exergonic by 3.7 kcal/mol. Al and O3 are separated with the distance of 2.087, 2.082, and 2.092 Å in Ceq3a, Ceq3b, and Ceq3c and 2.113, 2.123, and 2.123 Å in Cax3a, Cax3b, and Cax3c. The distances reveal the Al…O3 interaction in Ceq3a–c/Cax3a–c, driving the C1–O3 bond cleavage to complete the ring opening. CeqTS2a–c and CaxTS2a–c represent the possible transition-state structures for the ring-opening steps. The C1–O3 are elongated to 1.924, 1.932, and 1.898 Å in CeqTS2a, CeqTS2b, and CeqTS2c. The corresponding distances in CaxTS2a, CaxTS2b, and CaxTS2c are 1.932, 1.922, and 1.781 Å. The reactions then generate the final products Ceq4a–c and Cax4a–c by conquering small energy barriers of ~2 (Route 1) and ~0.5–5 (Route 2) kcal/mol, respectively. In Ceq4a–c and Cax4a–c, the C1–O3 of ε-caprolactone has been broken and reformed to a nine-membered ring with Al. Consequently, the coordination–insertion ring-opening reactions are slightly endergonic by less than 2 kcal/mol in both routes of L and LCl system. For the cases of LiPr systems, Route 1 is predicted to be 3.6 kcal/mol endergonic, whereas Route 2 is 1.8 kcal/mol exergonic.
Comparing the reaction barriers, Route 2 requires much lower free energy than Route 1 to make the reaction take place, indicating that coordination of ε-caprolactone in the axial position is the better direction to undergo nucleophilic addition. The corresponding structures of the transition states in Route 2 display a C1…O4 distance about 1.75 Å, shorter than those in Route 1 (about 1.92 Å). The shorter distance of C1…O4 in Route 2 also supports the fact that the benzoxide fragment can attack the carbonyl group of ε-caprolactone more effectively. As a result, Route 2 is a preferential pathway for the ring-opening polymerization of ε-caprolactone. Moreover, the activation barriers in the Lipr system is about 1 kcal/mol smaller than those in the L and LCl systems, in consistency with the experimental results. It can be attributed to the sterically encumbered groups obliging the benzoxide fragment to get closer to the ε-caprolactone thus speeding up the occurrence of the insertion step, which determines the rate of the ring-opening polymerization of ε-caprolactone.

3.2. Ring-Opening Polymerization of L-Lactide

The ring-opening polymerization mechanism of L-lactide shows different characteristics compared to ε-caprolactone polymerization. Although L-lactide can approach the Al center from the equatorial site, the results of the scan calculation show that benzoxide cannot attack the carbonyl group from this orientation. Hence, Route 2 is the only possible pathway for the ring-opening polymerization of L-lactide. The Al–O2 (Al–OLA) bond lengths of Lax1a, Lax1b, and Lax1c are 2.091, 2.072, and 2.099 Å, respectively (Figure 8). The values are slightly longer than those in Cax1a–c, probably because of the bulkier monomer of L-lactide. The bond enthalpies of Al–O2 in Lax1a–c are calculated to be larger than those in Cax1a−c (Lax1a: −12.6 kcal/mol; Lax1b: −13.9 kcal/mol; Lax1c: –12.3 kcal/mol). Among these species, the electron withdrawing inductive effect in Lax1b would lead to the more stable Al–O2 bond. Nevertheless, the calculation result in Lax1c implies that the steric hindrance does not exhibit obvious influence on the Al–O2 bond enthalpy.
DFT calculated Gibbs energy profiles of the proposed mechanism for the Lax1a-c catalyzed ring-opening polymerization of L-lactide are shown in Figure 9, Figure 10 and Figure 11. Different from the case of ε-caprolactone, the cleavage of the C–O bond is the rate-determining step in the ring-opening polymerization of L-lactide. The activation barriers generated from the nucleophilic addition of benzoxide to the carbonyl group on the L-lactide are about 7 kcal/mol, similar to the case of the ring-opening polymerization of ε-caprolactone. The corresponding transition states are determined to be a distorted square pyramidal structure, and the C1…O4 distances are separated by 1.934 Å in LaxTS1a, 1.932 Å in LaxTS1b, and 1.869 Å in LaxTS1c. The bulky pendant groups in LaxTS1c would lead to confined space and therefore result in the shorter C1…O4 distance. The intermediates produced from the insertion step are the species Lax2a–c with Al…O4 distances of 2.149, 2.135, and 2.195 Å in Lax2a, Lax2b, and Lax2c, respectively. Lax2a–c then undergo a bond rotation and yield Lax3a–c, the conformations more suitable for C–O cleavage. Compared to Lax2a–c, the Al…O3 distances in Lax3a–c are shortened while C1–O4 bonds are elongated. Noteworthily, the Al…O3 distance in Lax3c (2.342 Å) is much longer than in Lax3a (2.215 Å) and Lax3b (2.174 Å). The breakage of the C1–O3 bond involves the transition states LaxTS2a–c which lie 12.1, 11.6, and 15.1 kcal/mol above Lax3a, Lax3b, and Lax3c, respectively. The highest potential Gibbs energies of LaxTS2a–c reveal the C1–O3 breaking step determines the rate of the ring opening polymerization of L-lactide. The calculation results show that the C1–O3 distances are 2.002, 2.001, and 2.069 Å in LaxTS2a, LaxTS2b, and LaxTS2c, respectively. Following the intrinsic reaction coordinate, LaxTS2a–c give rise to the unstable intermediates Lax4a–c which are better described as a distorted tetrahedral species. Finally, Lax4a–c rearrange to the conformations of Lax5a–c, which adopt a TBP structure with the bidentate coordination of the ring-open L-lactide moiety.
The theoretical prediction shows that the LiPr system has the largest activation barrier among the three catalysts. Therefore, it would initiate the ring-opening polymerization of L-lactide more slowly, in agreement with the experimental results. The reaction is rate-determined by the ring opening step, the produced fragment of which would occupy more space. As a consequence, the reaction will be retarded by the bulky substituents. On the other hand, the reaction barriers of the ring-opening polymerization of L-lactide are estimated to be about twice larger than those of ε-caprolactone. The computational results have elucidated the fact that the ring-opening polymerization of L-lactide need to proceed at higher temperature in comparison to the case of ε-caprolactone.

4. Conclusions

In this study, we have shown that the ring-opening polymerization of ε-caprolactone is rate-determined by the step of benzoxide insertion, while the bond cleavage step is the slowest step in the case of ring-opening polymerization of L-lactide. Besides, theoretical analysis revealed that the reaction barriers of the ring-opening polymerization of L-lactide are higher than those of ε-caprolactone, corresponding to the experimental results. Finally, the pendant groups on the ketiminate ligand of the aluminum complexes play an essential role in catalyzing the ring-opening polymerization. In the ε-caprolactone ring-opening polymerization system, the crucial transition state is the stage benzoxide attacking the caprolactone. When the ketimate ligand bears more bulky groups, the congestion can be more largely released. Consequently, the reaction barrier is lowered to facilitate the occurrence of insertion when the pendant groups are more bulky. On the contrary, when L-lactides are loaded as monomers, the aluminum complexes supported by bulky groups are not conducive to the ring-opening polymerization. Since the rate-determining step of bond cleavage would lead to the occupation of space, the bulky substituents would go against the reaction to prevent from losing space. The calculation results are in agreement with the experimental data.

Supplementary Materials

The optimized structures involved in this study are available online at https://www.mdpi.com/2073-4360/11/9/1530/s1, Table S1: Cartesian coordinate of Ceq1a, Table S2: Cartesian coordinate of CeqTS1a, Table S3: Cartesian coordinate of Ceq2a, Table S4: Cartesian coordinate of Ceq3a, Table S5: Cartesian coordinate of CeqTS2a, Table S6: Cartesian coordinate of Ceq4a, Table S7: Cartesian coordinate of Cax1a, Table S8: Cartesian coordinate of CaxTS1a, Table S9: Cartesian coordinate of Cax2a, Table S10: Cartesian coordinate of Cax3a, Table S11: Cartesian coordinate of CaxTS2a, Table S12: Cartesian coordinate of Cax4a, Table S13: Cartesian coordinate of Ceq1b, Table S14: Cartesian coordinate of CeqTS1b, Table S15: Cartesian coordinate of Ceq2b, Table S16: Cartesian coordinate of Ceq3b, Table S17: Cartesian coordinate of CeqTS2b, Table S18: Cartesian coordinate of Ceq5b, Table S19: Cartesian coordinate of Cax1b, Table S20: Cartesian coordinate of CaxTS1b, Table S21: Cartesian coordinate of Cax2b, Table S22: Cartesian coordinate of Cax3b, Table S23: Cartesian coordinate of CaxTS2b, Table S24: Cartesian coordinate of Cax4b, Table S25: Cartesian coordinate of Ceq1c, Table S26: Cartesian coordinate of CeqTS1c, Table S27: Cartesian coordinate of Ceq2c, Table S28: Cartesian coordinate of Ceq3c, Table S29: Cartesian coordinate of CeqTS2c, Table S30: Cartesian coordinate of Ceq4c, Table S31: Cartesian coordinate of Cax1c, Table S32: Cartesian coordinate of CaxTS1c, Table S33: Cartesian coordinate of Cax2c, Table S34: Cartesian coordinate of Cax3c, Table S35: Cartesian coordinate of CaxTS2c, Table S36: Cartesian coordinate of Cax4c, Table S37: Cartesian coordinate of Lax1a, Table S38: Cartesian coordinate of LaxTS1a, Table S39: Cartesian coordinate of Lax2a, Table S40: Cartesian coordinate of Lax3a, Table S41: Cartesian coordinate of LaxTS2a, Table 42: Cartesian coordinate of Lax4a, Table S43: Cartesian coordinate of Lax5a, Table S44: Cartesian coordinate of Lax1b, Table S45: Cartesian coordinate of LaxTS1b, Table S46: Cartesian coordinate of Lax2b, Table S47: Cartesian coordinate of Lax3b, Table S48: Cartesian coordinate of LaxTS2b, Table S49: Cartesian coordinate of Lax4b, Table S50: Cartesian coordinate of Lax5a, Table S51: Cartesian coordinate of Lax1c, Table S52: Cartesian coordinate of LaxTS1c, Table S53: Cartesian coordinate of Lax2c, Table S54: Cartesian coordinate of Lax3c, Table S55: Cartesian coordinate of LaxTS2c, Table S56: Cartesian coordinate of Lax4c, Table S57: Cartesian coordinate of Lax5c.

Author Contributions

Conceptualization, Y.-F.L.; methodology, Y.-F.L.; Calculation, Y.-F.L.; Visualization, Y.-F.L.; Writing—review and editing, Y.-F.L.; Supervision, Y.-F.L.; Calculation, N.Y.-J.

Funding

This research was funded by the Ministry of Science and Technology (MOST), grant number MOST 106-2113-M-037-007 and MOST 107-2113-M-037-005-MY2.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chiellini, E.; Solaro, R. Biodegradable Polymeric Materials. Adv. Mater. 1996, 8, 305–313. [Google Scholar] [CrossRef]
  2. Drumright, R.E.; Gruber, P.R.; Henton, D.E. Polylactic Acid Technology. Adv. Mater. 2000, 12, 1841–1846. [Google Scholar] [CrossRef]
  3. Albertsson, A.-C.; Varma, I.K. Recent Developments in Ring Opening Polymerization of Lactones for Biomedical Applications. Biomacromolecules 2003, 4, 1466–1486. [Google Scholar] [CrossRef] [PubMed]
  4. Auras, R.; Harte, B.; Selke, S. An Overview of Polylactides as Packaging Materials. Macromol. Biosci. 2004, 4, 835–864. [Google Scholar] [CrossRef] [PubMed]
  5. Faÿ, F.; Renard, E.; Langlois, V.; Linossier, I.; Vallée-Rehel, K. Development of Poly(ε-caprolactone-co-l-lactide) and Poly(ε-caprolactone-co-δ-valerolactone) as New Degradable Binder Used for Antifouling Paint. Eur. Polym. J. 2007, 43, 4800–4813. [Google Scholar] [CrossRef]
  6. Slomkowski, S.; Penczek, S.; Duda, A. Polylactides–an Overview. Polym. Adv. Technol. 2014, 25, 436–447. [Google Scholar] [CrossRef]
  7. Nomura, N.; Ishii, R.; Akakura, M.; Aoi, K. Stereoselective Ring-Opening Polymerization of Racemic Lactide Using Aluminum-Achiral Ligand Complexes:  Exploration of a Chain-End Control Mechanism. J. Am. Chem. Soc. 2002, 124, 5938–5939. [Google Scholar] [CrossRef]
  8. Zhong, Z.; Dijkstra, P.J.; Feijen, J. [(salen)Al]-Mediated, Controlled and Stereoselective Ring-Opening Polymerization of Lactide in Solution and without Solvent: Synthesis of Highly Isotactic Polylactide Stereocopolymers from Racemic d,l-Lactide. Angew. Chem. Int. Ed. 2002, 41, 4510–4513. [Google Scholar] [CrossRef]
  9. Alcazar-Roman, L.M.; O’Keefe, B.J.; Hillmyer, M.A.; Tolman, W.B. Electronic Influence of Ligand Substituents on The Rate of Polymerization of ε-Caprolactone by Single-Site Aluminium Alkoxide Catalysts. Dalton Trans 2003, 3082–3087. [Google Scholar] [CrossRef]
  10. Hormnirun, P.; Marshall, E.L.; Gibson, V.C.; White, A.J.P.; Williams, D.J. Remarkable Stereocontrol in the Polymerization of Racemic Lactide Using Aluminum Initiators Supported by Tetradentate Aminophenoxide Ligands. J. Am. Chem. Soc. 2004, 126, 2688–2689. [Google Scholar] [CrossRef]
  11. Ishii, R.; Nomura, N.; Kondo, T. Stereoselective Bulk Polymerization of Racemic Lactide for Stereoblock Poly(racemic lactide) Using an Achiral Aluminum Complex. Polym. J. 2004, 36, 261–264. [Google Scholar] [CrossRef] [Green Version]
  12. Hormnirun, P.; Marshall, E.L.; Gibson, V.C.; Pugh, R.I.; White, A.J.P. Study of Ligand Substituent Effects on the Rate and Stereoselectivity of Lactide Polymerization Using Aluminum Salen-Type Initiators. Proc. Natl. Acad. Sci. USA 2006, 103, 15343–15348. [Google Scholar] [CrossRef] [PubMed]
  13. Nomura, N.; Ishii, R.; Yamamoto, Y.; Kondo, T. Stereoselective Ring-Opening Polymerization of a Racemic Lactide by Using Achiral Salen and Homosalen-Aluminum Complexes. Chem. Eur. J. 2007, 13, 4433–4451. [Google Scholar] [CrossRef] [PubMed]
  14. Chisholm, M.H.; Gallucci, J.C.; Quisenberry, K.T.; Zhou, Z. Complexities in the Ring-Opening Polymerization of Lactide by Chiral Salen Aluminum Initiators. Inorg. Chem. 2008, 47, 2613–2624. [Google Scholar] [CrossRef] [PubMed]
  15. Shen, M.; Zhang, W.; Nomura, K.; Sun, W.-H. Synthesis and Characterization of Organoaluminum Compounds Containing Quinolin-8-amine Derivatives and Their Catalytic Behaviour for Ring-Opening Polymerization of ε-Caprolactone. Dalton Trans. 2009, 41, 9000–9009. [Google Scholar] [CrossRef]
  16. Darensbourg, D.J.; Karroonnirun, O. Stereoselective Ring-Opening Polymerization of rac-Lactides Catalyzed by Chiral and Achiral Aluminum Half-Salen Complexes. Organometallics 2010, 29, 5627–5634. [Google Scholar] [CrossRef]
  17. Sun, W.-H.; Shen, M.; Zhang, W.; Huang, W.; Liu, S.; Redshaw, C. Methylaluminium 8-quinolinolates: Synthesis, characterization and use in ring-opening polymerization (ROP) of ε-caprolactone. Dalton Trans. 2011, 40, 2645–2653. [Google Scholar] [CrossRef]
  18. Whitelaw, E.L.; Loraine, G.; Mahon, M.F.; Jones, M.D. Salalen Aluminium Complexes and their Exploitation for the Ring Opening Polymerisation of rac-Lactide. Dalton Trans. 2011, 40, 11469–11473. [Google Scholar] [CrossRef]
  19. Bakewell, C.; Platel, R.H.; Cary, S.K.; Hubbard, S.M.; Roaf, J.M.; Levine, A.C.; White, A.J.P.; Long, N.J.; Haaf, M.; Williams, C.K. Bis(8-quinolinolato)aluminum ethyl Complexes: Iso-Selective Initiators for rac-Lactide Polymerization. Organometallics 2012, 31, 4729–4736. [Google Scholar] [CrossRef]
  20. Chen, H.-L.; Dutta, S.; Huang, P.-Y.; Lin, C.-C. Preparation and Characterization of Aluminum Alkoxides Coordinated on salen-Type Ligands: Highly Stereoselective Ring-Opening Polymerization of rac-Lactide. Organometallics 2012, 31, 2016–2025. [Google Scholar] [CrossRef]
  21. Ikpo, N.; Barbon, S.M.; Drover, M.W.; Dawe, L.N.; Kerton, F.M. Aluminum Methyl and Chloro Complexes Bearing Monoanionic Aminephenolate Ligands: Synthesis, Characterization, and Use in Polymerizations. Organometallics 2012, 31, 8145–8158. [Google Scholar] [CrossRef]
  22. Castro-Osma, J.A.; Alonso-Moreno, C.; Márquez-Segovia, I.; Otero, A.; Lara-Sánchez, A.; Fernández-Baeza, J.; Rodríguez, A.M.; Sánchez-Barba, L.F.; García-Martínez, J.C. Synthesis, Structural Characterization and Catalytic Evaluation of the Ring-Opening Polymerization of Discrete Five-Coordinate Alkyl Aluminium Complexes. Dalton Trans. 2013, 42, 9325–9337. [Google Scholar] [CrossRef] [PubMed]
  23. Gao, B.; Duan, R.; Pang, X.; Li, X.; Qu, Z.; Tang, Z.; Zhuang, X.; Chen, X. Stereoselective Ring-Opening Polymerization of rac-Lactides Catalyzed by Aluminum Hemi-Salen Complexes. Organometallics 2013, 32, 5435–5444. [Google Scholar] [CrossRef]
  24. Hancock, S.L.; Mahon, M.F.; Jones, M.D. Aluminium Salalen Complexes Based on 1,2-Diaminocyclohexane and their Exploitation for the Polymerisation of rac-Lactide. Dalton Trans. 2013, 42, 9279–9285. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, J.; Ma, H. Aluminum Complexes with Bidentate Amido Ligands: Synthesis, Structure and Performance on Ligand-Initiated Ring-Opening Polymerization of rac-Lactide. Dalton Trans. 2014, 43, 9098–9110. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, W.; Wang, Y.; Wang, L.; Redshaw, C.; Sun, W.-H. Dialkylaluminium 2-Imidazolylphenolates: Synthesis, Characterization and Ring-Opening Polymerization Behavior Towards Lactides. J. Organomet. Chem. 2014, 750, 65–73. [Google Scholar] [CrossRef]
  27. Tabthong, S.; Nanok, T.; Sumrit, P.; Kongsaeree, P.; Prabpai, S.; Chuawong, P.; Hormnirun, P. Bis(pyrrolidene) Schiff Base Aluminum Complexes as Isoselective-Biased Initiators for the Controlled Ring-Opening Polymerization of rac-Lactide: Experimental and Theoretical Studies. Macromolecules 2015, 48, 6846–6861. [Google Scholar] [CrossRef]
  28. Fuoco, T.; Pappalardo, D. Aluminum Alkyl Complexes Bearing Salicylaldiminato Ligands: Versatile Initiators in the Ring-Opening Polymerization of Cyclic Esters. Catalysts 2017, 7, 64. [Google Scholar] [CrossRef]
  29. Nakonkhet, C.; Nanok, T.; Wattanathana, W.; Chuawong, P.; Hormnirun, P. Aluminium Complexes Containing Salicylbenzothiazole Ligands and their Application in the Ring-Opening Polymerisation of rac-Lactide and ε-Caprolactone. Dalton Trans. 2017, 46, 11013–11030. [Google Scholar] [CrossRef]
  30. Huang, H.-C.; Wang, B.; Chen, X.-L.; Pan, L.; Li, Y.-S. Ring-Opening Polymerization of (Macro)lactones by Highly Active Mononuclear Salen-Aluminum Complexes Bearing Cyclic β-Ketoiminato Ligand. J. Polym. Sci. Polym. Chem. 2019, 57, 973–981. [Google Scholar] [CrossRef]
  31. Kricheldorf, H.R.; Berl, M.; Scharnagl, N. Poly(lactones). 9. Polymerization Mechanism of Metal Alkoxide Initiated Polymerizations of Lactide and Various Lactones. Macromolecules 1988, 21, 286–293. [Google Scholar] [CrossRef]
  32. Dubois, P.; Barakat, I.; Jerome, R.; Teyssie, P. Macromolecular Engineering of Polyactones and Polyactides. 12. Study of the Depolymerization Reactions of Poly(.epsilon-caprolactone) with Functional Aluminum Alkoxide end Groups. Macromolecules 1993, 26, 4407–4412. [Google Scholar] [CrossRef]
  33. Yu, R.-C.; Hung, C.-H.; Huang, J.-H.; Lee, H.-Y.; Chen, J.-T. Four- and Five-Coordinate Aluminum Ketiminate Complexes:  Synthesis, Characterization, and Ring-Opening Polymerization. Inorg. Chem. 2002, 41, 6450–6455. [Google Scholar] [CrossRef] [PubMed]
  34. Tabthong, S.; Nanok, T.; Kongsaeree, P.; Prabpai, S.; Hormnirun, P. Monomethylaluminum and Dimethylaluminum Pyrrolylaldiminates for the Ring-Opening Polymerization of rac-Lactide: Effects of Ligand Structure and Coordination Geometry. Dalton Trans. 2014, 43, 1348–1359. [Google Scholar] [CrossRef] [PubMed]
  35. Sumrit, P.; Chuawong, P.; Nanok, T.; Duangthongyou, T.; Hormnirun, P. Aluminum Complexes Containing Salicylbenzoxazole Ligands and their Application in the Ring-Opening Polymerization of rac-Lactide and ε-Caprolactone. Dalton Trans. 2016, 45, 9250–9266. [Google Scholar] [CrossRef]
  36. Lee, C.-L.; Lin, Y.-F.; Jiang, M.-T.; Lu, W.-Y.; Vandavasi, J.K.; Wang, L.-F.; Lai, Y.-C.; Chiang, M.Y.; Chen, H.-Y. Improvement in Aluminum Complexes Bearing Schiff Bases in Ring-Opening Polymerization of ε-Caprolactone: A Five-Membered-Ring System. Organometallics 2017, 36, 1936–1945. [Google Scholar] [CrossRef]
  37. Nomura, N.; Akita, A.; Ishii, R.; Mizuno, M. Random Copolymerization of ε-Caprolactone with Lactide Using a Homosalen−Al Complex. J. Am. Chem. Soc. 2010, 132, 1750–1751. [Google Scholar] [CrossRef] [PubMed]
  38. Koller, J.; Bergman, R.G. Highly Efficient Aluminum-Catalyzed Ring-Opening Polymerization of Cyclic Carbonates, Lactones, and Lactides, Including a Unique Crystallographic Snapshot of an Intermediate. Organometallics 2011, 30, 3217–3224. [Google Scholar] [CrossRef]
  39. Lamberti, M.; D’Auria, I.; Mazzeo, M.; Milione, S.; Bertolasi, V.; Pappalardo, D. Phenoxy-Thioether Aluminum Complexes as ε-Caprolactone and Lactide Polymerization Catalysts. Organometallics 2012, 31, 5551–5560. [Google Scholar] [CrossRef]
  40. Hild, F.; Neehaul, N.; Bier, F.; Wirsum, M.; Gourlaouen, C.; Dagorne, S. Synthesis and Structural Characterization of Various N,O,N-Chelated Aluminum and Gallium Complexes for the Efficient ROP of Cyclic Esters and Carbonates: How Do Aluminum and Gallium Derivatives Compare? Organometallics 2013, 32, 587–598. [Google Scholar] [CrossRef]
  41. McKeown, P.; Davidson, M.G.; Kociok-Köhn, G.; Jones, M.D. Aluminium Salalens vs. Salans: “Initiator Design” for the Isoselective Polymerisation of rac-Lactide. Chem. Commun. 2016, 52, 10431–10434. [Google Scholar] [CrossRef] [PubMed]
  42. Stasiw, D.E.; Mandal, M.; Neisen, B.D.; Mitchell, L.A.; Cramer, C.J.; Tolman, W.B. Why So Slow? Mechanistic Insights from Studies of a Poor Catalyst for Polymerization of ε-Caprolactone. Inorg. Chem. 2017, 56, 725–728. [Google Scholar] [CrossRef] [PubMed]
  43. Beament, J.; Mahon, M.F.; Buchard, A.; Jones, M.D. Aluminum Complexes of Monopyrrolidine Ligands for the Controlled Ring-Opening Polymerization of Lactide. Organometallics 2018, 37, 1719–1724. [Google Scholar] [CrossRef]
  44. Macaranas, J.A.; Luke, A.M.; Mandal, M.; Neisen, B.D.; Marell, D.J.; Cramer, C.J.; Tolman, W.B. Sterically Induced Ligand Framework Distortion Effects on Catalytic Cyclic Ester Polymerizations. Inorg. Chem. 2018, 57, 3451–3457. [Google Scholar] [CrossRef]
  45. Ropson, N.; Dubois, P.; Jerome, R.; Teyssie, P. Macromolecular Engineering of Polylactones and Polylactides. 20. Effect of Monomer, Solvent, and Initiator on the Ring-Opening Polymerization as Initiated with Aluminum Alkoxides. Macromolecules 1995, 28, 7589–7598. [Google Scholar] [CrossRef]
  46. Lewiński, J.; Horeglad, P.; Wójcik, K.; Justyniak, I. Chelation Effect in Polymerization of Cyclic Esters by Metal Alkoxides:  Structure Characterization of the Intermediate Formed by Primary Insertion of Lactide into the Al−OR Bond of an Organometallic Initiator. Organometallics 2005, 24, 4588–4593. [Google Scholar] [CrossRef]
  47. Parenté, V.; Brédas, J.-L.; Dubois, P.; Ropson, N.; Jérôme, R. Theoretical Investigation of the Molecular Structure of Aluminium Triisopropoxide and its Complexes in Ring-Opening Polymerization. Macromol. Theory Simul. 1996, 5, 525–546. [Google Scholar] [CrossRef]
  48. Chen, Y.-H.; Chen, Y.-J.; Tseng, H.-C.; Lian, C.-J.; Tsai, H.-Y.; Lai, Y.-C.; Hsu, S.C.N.; Chiang, M.Y.; Chen, H.-Y. Comparing l-Lactide and ε-Caprolactone Polymerization by Using Aluminum Complexes Bearing Ketiminate Ligands: Steric, Electronic, and Chelating Effects. RSC Adv. 2015, 5, 100272–100280. [Google Scholar] [CrossRef]
  49. Tseng, H.-C.; Chiang, M.Y.; Lu, W.-Y.; Chen, Y.-J.; Lian, C.-J.; Chen, Y.-H.; Tsai, H.-Y.; Lai, Y.-C.; Chen, H.-Y. A Closer Look at ε-Caprolactone Polymerization Catalyzed by Alkyl Aluminum Complexes: The Effect of Induction Period on Overall Catalytic Activity. Dalton Trans. 2015, 44, 11763–11773. [Google Scholar] [CrossRef]
  50. Chang, M.-C.; Lu, W.-Y.; Chang, H.-Y.; Lai, Y.-C.; Chiang, M.Y.; Chen, H.-Y.; Chen, H.-Y. Comparative Study of Aluminum Complexes Bearing N,O- and N,S-Schiff Base in Ring-Opening Polymerization of ε-Caprolactone and l-Lactide. Inorg. Chem. 2015, 54, 11292–11298. [Google Scholar] [CrossRef]
  51. Chang, C.-C.; Chang, Y.-C.; Lu, W.-Y.; Lai, Y.-C.; Wu, K.-H.; Lin, Y.-F.; Chen, H.-Y. Novel Aluminum Complexes Bearing 2-(Aminomethylene)malonate Ligands with High Efficiency and Controllability in Ring-Opening Polymerization of ε-Caprolactone. Eur. Polym. J. 2019, 115, 399–408. [Google Scholar] [CrossRef]
  52. Mandal, M.; Luke, A.M.; Dereli, B.; Elwell, C.E.; Reineke, T.M.; Tolman, W.B.; Cramer, C.J. Computational Prediction and Experimental Verification of ε-Caprolactone Ring-Opening Polymerization Activity by an Aluminum Complex of an Indolide/Schiff-Base Ligand. ACS Catal. 2019, 9, 885–889. [Google Scholar] [CrossRef]
  53. Von Schenck, H.; Ryner, M.; Albertsson, A.-C.; Svensson, M. Ring-Opening Polymerization of Lactones and Lactides with Sn(IV) and Al(III) Initiators. Macromolecules 2002, 35, 1556–1562. [Google Scholar] [CrossRef]
  54. Marshall, E.L.; Gibson, V.C.; Rzepa, H.S. A Computational Analysis of the Ring-Opening Polymerization of rac-Lactide Initiated by Single-Site β-Diketiminate Metal Complexes:  Defining the Mechanistic Pathway and the Origin of Stereocontrol. J. Am. Chem. Soc. 2005, 127, 6048–6051. [Google Scholar] [CrossRef] [PubMed]
  55. Miranda, M.O.; DePorre, Y.; Vazquez-Lima, H.; Johnson, M.A.; Marell, D.J.; Cramer, C.J.; Tolman, W.B. Understanding the Mechanism of Polymerization of ε-Caprolactone Catalyzed by Aluminum Salen Complexes. Inorg. Chem. 2013, 52, 13692–13701. [Google Scholar] [CrossRef] [PubMed]
  56. Vieira, I.d.S.; Whitelaw, E.L.; Jones, M.D.; Herres-Pawlis, S. Synergistic Empirical and Theoretical Study on the Stereoselective Mechanism for the Aluminum Salalen Complex Mediated Polymerization of rac-Lactide. Chem. Eur. J. 2013, 19, 4712–4716. [Google Scholar] [CrossRef]
  57. Roymuhury, S.K.; Chakraborty, D.; Ramkumar, V. Aluminium Complexes Bearing N,O-Aminophenol Ligands as Efficient Catalysts for the Ring Opening Polymerization of Lactide. Eur. Polym. J. 2015, 70, 203–214. [Google Scholar] [CrossRef]
  58. Jitonnom, J.; Molloy, R.; Punyodom, W.; Meelua, W. Theoretical Studies on Aluminum Trialkoxide-Initiated Lactone Ring-Opening Polymerizations: Roles of Alkoxide Substituent and Monomer Ring Structure. Computa. Theor. Chem. 2016, 1097, 25–32. [Google Scholar] [CrossRef]
  59. Chandanabodhi, D.; Nanok, T. A DFT Study of the Ring-Opening Polymerization Mechanism of l-Lactide and ε-Caprolactone Using aluminium Salen-Type Initiators: Towards an Understanding of their Reactivities in Homo- and Copolymerization. Mol. Catal. 2017, 436, 145–156. [Google Scholar] [CrossRef]
  60. Robert, C.; Schmid, T.E.; Richard, V.; Haquette, P.; Raman, S.K.; Rager, M.-N.; Gauvin, R.M.; Morin, Y.; Trivelli, X.; Guérineau, V.; et al. Mechanistic Aspects of the Polymerization of Lactide Using a Highly Efficient Aluminum(III) Catalytic System. J. Am. Chem. Soc. 2017, 139, 6217–6225. [Google Scholar] [CrossRef]
  61. Wei, J.; Riffel, M.N.; Diaconescu, P.L. Redox Control of Aluminum Ring-Opening Polymerization: A Combined Experimental and DFT Investigation. Macromolecules 2017, 50, 1847–1861. [Google Scholar] [CrossRef] [Green Version]
  62. Li, P.; Xi, Y.; Li, L.; Li, H.; Sun, W.-H.; Lei, M. A DFT Study on Ring-Opening Polymerization of ε-Caprolactone Initiated by Mg and Al Complexes. Inorg. Chim. Acta 2018, 477, 34–39. [Google Scholar] [CrossRef]
  63. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Rev. A.03. Wallingford CT. 2016. Available online: https://gaussian.com/citation (accessed on 14 September 2019).
  64. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724–728. [Google Scholar] [CrossRef]
  65. Hehre, W.J.; Ditchfield, R.; Pople, J.A. Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56, 2257–2261. [Google Scholar] [CrossRef]
  66. Hariharan, P.C.; Pople, J.A. The influence of Polarization Functions on Molecular Orbital Hydrogenation Energies. Theoretica chimica acta 1973, 28, 213–222. [Google Scholar] [CrossRef]
  67. Francl, M.M.; Pietro, W.J.; Hehre, W.J.; Binkley, J.S.; Gordon, M.S.; DeFrees, D.J.; Pople, J.A. Self-Consistent Molecular Orbital Methods. XXIII. A Polarization-Type Basis set for Second-Row Elements. J. Chem. Phys. 1982, 77, 3654–3665. [Google Scholar] [CrossRef]
  68. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  69. Becke, A.D. Density-Functional Thermochemistry. III. the Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  70. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Chem. Phys. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  71. Grimme, S. Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  72. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef] [PubMed]
  73. Fukui, K. Formulation of the Reaction Coordinate. J. Phys. Chem. 1970, 74, 4161–4163. [Google Scholar] [CrossRef]
  74. Fukui, K. The Path of Chemical Reactions-the IRC Approach. Acc. Chem. Res. 1981, 14, 363–368. [Google Scholar] [CrossRef]
  75. Gonzalez, C.; Schlegel, H.B. Reaction Path Following in Aass-Weighted Internal Voordinates. J. Phys. Chem. 1990, 94, 5523–5527. [Google Scholar] [CrossRef]
  76. Maeda, S.; Harabuchi, Y.; Ono, Y.; Taketsugu, T.; Morokuma, K. Intrinsic Reaction Coordinate: Calculation, Bifurcation, and Automated Search. Int. J. Quantum Chem. 2015, 115, 258–269. [Google Scholar] [CrossRef]
  77. Chan, M.S.W.; Deng, L.; Ziegler, T. Density Functional Study of Neutral Salicylaldiminato Nickel(II) Complexes as Olefin Polymerization Catalysts. Organometallics 2000, 19, 2741–2750. [Google Scholar] [CrossRef]
Figure 1. A series of aluminum ketiminate complexes bearing different substituents for the ring-opening polymerization of ε-caprolactone and L-lactide.
Figure 1. A series of aluminum ketiminate complexes bearing different substituents for the ring-opening polymerization of ε-caprolactone and L-lactide.
Polymers 11 01530 g001
Figure 2. Experimental data summary for the ring-opening polymerization of ε-caprolactone and L-lactide initiated by aluminum ketiminate complexes.
Figure 2. Experimental data summary for the ring-opening polymerization of ε-caprolactone and L-lactide initiated by aluminum ketiminate complexes.
Polymers 11 01530 g002
Figure 3. Thirteen possible initial geometries considered in this study.
Figure 3. Thirteen possible initial geometries considered in this study.
Polymers 11 01530 g003
Scheme 1. Four possible routes starting from four different triangular bipyramidal (TBP) monomer-coordinated aluminum isomers and follow the coordination-insertion mechanism.
Scheme 1. Four possible routes starting from four different triangular bipyramidal (TBP) monomer-coordinated aluminum isomers and follow the coordination-insertion mechanism.
Polymers 11 01530 sch001
Figure 4. ε-Caprolactone coordinated aluminum complexes which bear L, LCl, and LiPr ligand systems and arrange in a TBP fashion.
Figure 4. ε-Caprolactone coordinated aluminum complexes which bear L, LCl, and LiPr ligand systems and arrange in a TBP fashion.
Polymers 11 01530 g004
Figure 5. Reaction profile and optimized structures of the L system. (Monomer: ε-caprolactone).
Figure 5. Reaction profile and optimized structures of the L system. (Monomer: ε-caprolactone).
Polymers 11 01530 g005
Figure 6. Reaction profile and optimized structures of the LCl system. (Monomer: ε-caprolactone).
Figure 6. Reaction profile and optimized structures of the LCl system. (Monomer: ε-caprolactone).
Polymers 11 01530 g006
Figure 7. Reaction profile and optimized structures of the LiPr system. (Monomer: ε-caprolactone).
Figure 7. Reaction profile and optimized structures of the LiPr system. (Monomer: ε-caprolactone).
Polymers 11 01530 g007
Figure 8. L-lactide coordinated aluminum complexes which bear L, LCl, and LiPr ligand systems and arrange in a TBP fashion.
Figure 8. L-lactide coordinated aluminum complexes which bear L, LCl, and LiPr ligand systems and arrange in a TBP fashion.
Polymers 11 01530 g008
Figure 9. Reaction profile and optimized structures of the L system. (Monomer: L-lactide).
Figure 9. Reaction profile and optimized structures of the L system. (Monomer: L-lactide).
Polymers 11 01530 g009
Figure 10. Reaction profile and optimized structures of the LCl system. (Monomer: L-lactide).
Figure 10. Reaction profile and optimized structures of the LCl system. (Monomer: L-lactide).
Polymers 11 01530 g010
Figure 11. Reaction profile and optimized structures of the LiPr system. (Monomer: L-lactide).
Figure 11. Reaction profile and optimized structures of the LiPr system. (Monomer: L-lactide).
Polymers 11 01530 g011

Share and Cite

MDPI and ACS Style

Lin, Y.-F.; Jheng, N.-Y. Mechanistic Insight into the Ring-Opening Polymerization of ε-Caprolactone and L-Lactide Using Ketiminate-Ligated Aluminum Catalysts. Polymers 2019, 11, 1530. https://doi.org/10.3390/polym11091530

AMA Style

Lin Y-F, Jheng N-Y. Mechanistic Insight into the Ring-Opening Polymerization of ε-Caprolactone and L-Lactide Using Ketiminate-Ligated Aluminum Catalysts. Polymers. 2019; 11(9):1530. https://doi.org/10.3390/polym11091530

Chicago/Turabian Style

Lin, Ya-Fan, and Nai-Yuan Jheng. 2019. "Mechanistic Insight into the Ring-Opening Polymerization of ε-Caprolactone and L-Lactide Using Ketiminate-Ligated Aluminum Catalysts" Polymers 11, no. 9: 1530. https://doi.org/10.3390/polym11091530

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop