You are currently viewing a new version of our website. To view the old version click .
Polymers
  • Review
  • Open Access

12 December 2019

DFT Modeling of Organocatalytic Ring-Opening Polymerization of Cyclic Esters: A Crucial Role of Proton Exchange and Hydrogen Bonding

and
1
Chemistry Department, M.V. Lomonosov Moscow State University, 1 Leninskie Gory Str., Building 3, 119991 Moscow, Russia
2
A.V. Topchiev Institute of Petrochemical Synthesis RAS, 29 Leninsky Pr., 119991 Moscow, Russia
*
Authors to whom correspondence should be addressed.
This article belongs to the Section Polymer Chemistry

Abstract

Organocatalysis is highly efficient in the ring-opening polymerization (ROP) of cyclic esters. A variety of initiators broaden the areas of organocatalysis in polymerization of different monomers, such as lactones, cyclic carbonates, lactides or gycolides, ethylene phosphates and phosphonates, and others. The mechanisms of organocatalytic ROP are at least as diverse as the mechanisms of coordination ROP; the study of these mechanisms is critical in ensuring the polymer compositions and architectures. The use of density functional theory (DFT) methods for comparative modeling and visualization of organocatalytic ROP pathways, in line with experimental proof of the structures of the reaction intermediates, make it possible to establish these mechanisms. In the present review, which continues and complements our recent manuscript that focused on DFT modeling of coordination ROP, we summarized the results of DFT modeling of organocatalytic ROP of cyclic esters and some related organocatalytic processes, such as polyester transesterification.

1. Introduction

The development of biotechnology and medical technology has set higher requirements for biomedical materials. The catalytic ring-opening polymerization (ROP) of cyclic esters and related compounds (Figure 1) provides a basis of the efficient synthesis of biodegradable and biocompatible polymers [,,,,,,,]. Tin (II) carboxylates [,,,] or aluminium alkoxides [,,,] were the first efficient ROP catalysts, in the next decades hundreds of metal complexes were studied in the polymerization of different cyclic substrates [,,,]. However, consideration must be given to the fundamental difference between metal-catalyzed industrial processes, such as Ziegler-Natta polymerization of α-olefins and ROP of cyclic esters. The monomer-to-catalyst molar ratio for the first process is typically in the range 106–108 []; therefore, the removal of the catalyst is not needed and polyolefins can be processed without purification. Instead, the synthesis of polyesters via the ROP of cyclic esters typically requires the monomer-to-catalyst ratios ~103–104 [,,]. The high oxophilicity of the metal complexes makes their removal from oxygen-rich polyesters difficult. Tackling the problem of the metal catalyst removal has become all the more crucial for sensitive domains, such as biomedical, packaging, and microelectronics []. The ability of organocatalysts to be effectively removed from the polymers is a significant asset. The other benefits of organocatalysts include mild reaction conditions, the option to vary initiator/catalyst ratios, and lower toxicity [,,,]. That is why the use of organocatalysts appears to be a reliable alternative to the coordination of ROP catalysts in the synthesis of biomedical-grade polymers.
Figure 1. Coordination (a) and organocatalytic (b) ring-opening polymerization (ROP); (c) cyclic esters, monomers for ROP.
To date, various organocatalysts of ROP have been studied [,,,,,,]. The density functional theory (DFT) methods have been applied for the modeling of the organocatalytic ROP to improve the understanding of the reaction mechanisms for different cyclic substrates, to determine the scope and limitations of ROP in the development of biodegradable polymers with given compositions and architectures. Houk et al. reviewed DFT modeling of the full diversity of organocatalytic processes in 2011 [], the recent review of Jones [] that addressed the contributions of quantum chemistry to the development of ROP also looked at some of the aspects of organocatalytic polymerization of cyclic esters. Ruipérez discussed some aspects of catalytic ROP in a recent review []. Traditionally, organocatalytic ROP mechanisms are divided into base-catalyzed and acid-catalyzed [,,,], but it became clear in recent years that such a split is rather conditional.
In the present review, we tried to summarize and analyze the results of DFT modeling of organocatalytic ROP reported in dozens of publications. In line with different types of organocatalysts (Figure 2), the material that is presented in the review is divided into six main parts bearing in mind the retrospective of the each research topics. This review complements our previous work that was devoted to the DFT modeling of coordination ROP that was recently published in the special issue of Molecules [].
Figure 2. Chronological compilation of the types of ROP organocatalysts studied by the density functional theory (DFT) method.

2. Polymerizability of Cyclic Esters

The ROP thermodynamics plays a crucial role for polymerizability of cyclic esters, as mentioned in a number of publications and reviewed in []. In our recent review [], we only mentioned this subject; in the present manuscript we consider the reaction ability of cyclic esters in more detail.
The strain of the lactone ring can be attributed to the higher stability of s-trans conformation of esters relative to s-cis-conformation [], which was confirmed by DFT optimization of methyl acetate conformers [,,] (Figure 3).
Figure 3. Geometries of s-trans and s-cis methyl acetate and relative Gibbs free energies (kcal/mol) of s-cis methyl acetate determined experimentally and calculated (in parentness). Reprinted with permission from []. Copyright (2008) American Chemical Society.
The strain energies for five- and six-membered lactones, γ-butyrolactone (γBL), and δ-valerolactone (δVL) were found to be ~8 kcal/mol (calorimetry data) []. Although this might be thought to be more than enough to cause the polymerization of these lactones, γBL does not polymerize under the typical ROP conditions, in contrast with easily polymerizable δVL. Houk et al. studied the thermochemistry of γBL and δVL ROP, and the conformational preferences of model molecules that mimic corresponding polymers, to explain these facts []. The geometry optimizations were preformed at B3LYP/6-31G(d) level [,,]; the CBS-QB3 method [,] was used for additional calculations of thermodynamic parameters. The calculations of the thermochemistry of the model transesterification reaction between the lactones and methyl acetate when given the values of ΔG as 1.0 and −1.3 kcal/mol for γBL and δVL, respectively. The conformation effects in the transesterification products also significantly contribute to the thermochemistry of the reactions [,]; a high preference for gauche-coiled conformations of the transesterification products for polymerizable lactones was also detected.
Aleman et al. performed later work on the polymerizability of 1,4-dioxan-2-one (PDO), ε-caprolactone (εCL) [], glycolide (GL), L-lactide (l-LA), and (R,S)-lactide []. All of these monomers were found to be polymerizable due to strain in the ester group. In addition, the predominance of high-energy coiled conformations in linear homopolymers the formed from PDO, εCL and GL was proportionately less than extended conformations, in comparison to the inactive monomer γBL. However, while the prevalence of coiled conformations over extended conformations was expected for poly(LA), this factor does not affected the reactivity of lactides. Very good agreement between the experimental and calculated data for ROP of GL, l-LA and (R,S)-lactide [] should be separately noted.

7. Thiourea-Based Catalysts

The mechanistics aspects of thiourea-based ROP catalysis have been recently reviewed []. In 2005, Waymouth, Hendrick et al. demonstrated unique catalytic behaviour of the amino-functionalized thiourea organocatalyst (Scheme 8) in lactide polymerization []. This catalyst demonstrated the extraordinary selectivity for polymerization relative to transesterification; the proposed reaction mechanism involved the bifunctional activation of the LA carbonyl via hydrogen bonding to the thiourea group and of the initiating/propagating alcohol by the Brønsted basic tertiary amino group (Scheme 8). In the further research of this group, it was demonstrated that the combination of thiourea with terniary amines represented an effective catalyst for the controlled ROP of lactides and lactones [].
Scheme 8. The proposed mechanism of dual activation by amino-functionalized thiourea catalyst [].
In 2009 [], Znang et al. performed DFT modeling of methanolysis of l-LA, which was catalyzed by amino-functionalized thiourea organocatalyst (Scheme 8) at B3LYP/6–311++G(d,p) [,,] level of theory. They found that the activation barrier was about 43 kcal/mol for non-catalytic reaction of l-LA with MeOH, whereas the activation energy of the catalytic reaction via the formation of triple hydrogen-bonded complex was 25.7 kcal/mol (Figure 29). The comparative calculations of the alternative “concerted” catalytic mechanism through CTS2 allowed for excluding the latter reaction pathway by high activation barrier (Figure 29). The zwitterionic mechanism was also modeled and excluded for the same reason.
Figure 29. Energy profile (electronic energies with zero-point corrections) for l-LA methanolysis catalyzed by amino-functionalized thiourea. Reprinted with permission from []. Copyright (2009) Australian Chemical Society.
Hedrick et al. established a very close mechanism for thiourea (TU) and spartein-catalyzed alcohol-initiated ROP of l-LA []. It was found by DFT calculations that it was the lone pair orientation in spartein molecule (Figure 30) provided an effective activation of MeOH molecule. Such an activation resulted in extremely high catalytic activity. Kazakov and Kiesewetter analyzed some aspects of catalytic behaviour of amine-TU catalysts [], using DFT modeling for the optimization of amine-TU complexes.
Figure 30. Thiourea (TU) and (–)-sparteine molecular structures and approximate geometry of the rate limiting step of (–)-sparteine catalyzed l-LA ROP with methanol showing no dependence on catalyst chirality, and the calculated distance between N atoms along with the lone pair angle extending from the N–N plane. Reprinted with permission from []. Copyright (2010) American Chemical Society.
Finally, Waymouth et al. effectively used the DFT methods to explain the outstanding catalytic characteristics of TU-derived anion—thioamidate—in l-LA ROP []. The catalysts containing this anion demonstrated higher activity in comparison with the MeONa initiator, and provided the obtaining of PLA with narrow MWD. In contrast with the mechanism that included the formation of three-component catalytic species TU–amine–ROH, as discussed above, the TU-derived thioamidate acts as a dual hydrogen-bonding activator for both l-LA and MeOH molecules. DFT modeling of the reaction profile was performed at B3LYP-D3/6-31+G(d)/aug-cc-pVTZ [,,,,,]; the impact of the solvent (CH2Cl2) was taken into account with the IEF-cPCM method []. The isopropyl-substituted model was used for calculations instead of TU molecule; Figure 31 presents the structures of the stationary points and transition states with the corresponding free energies. Note that the binding of l-LA molecule to the MeOH-bound thioimidate (INT5) was ~12 kcal/mol more favorable than binding of the open-chain ester (INT3). The mechanism that is presented in Figure 31 is very similar to the donor-acceptor mechanism of TBD-catalyzed ROP of lactones and lactides (Figure 14).
Figure 31. The structures and calculated free energies of the stationary points and transition states for thioamidate-catalyzed ROP of l-LA [].
The thioimidate catalyst demonstrated good catalytic activity in lactide polymerization, but lower activity in TMC ROP and low productivity in polymerization of lactones. The novel simple and promising catalysts, the close structural analogs of TU-derived thioamidate, were studied in the few past years [,,,,]. These catalysts, which represent the metal derivatives of disubstituted ureas, demonstrated high activities in ROP of cyclic esters of all actual structure types—lactide, lactones, TMC, and ethylene phosphates. Moreover, Meng et al. [] found that these catalysts are effective in the polymerization of the “inert in ROP” [,] cyclic ester, five-membered γBL. The problem of the synthesis of HMW poly(γBL) was solved due to high catalytic activity of these catalysts that allowed for running γBL polymerization at very low reaction temperatures. Meng et al. performed a thorough theoretical study of γBL ROP catalyzed by urea anions with different substituents at the nitrogen atoms while using DFT calculations at GGA-PBE [,,] level of theory for an explanation of the experimental results and understanding the reaction mechanism. The results of computational studies (Figure 32 and Figure 33) demonstrated that urea anion acted as a bifunctional catalyst that activated both alcohol molecule and γBL. According to thermodynamic calculation, the more alkaline urea with electron-donating group exhibited lower activation barrier and, therefore, demonstrated higher catalytic efficiency.
Figure 32. The stationary points and transition states of the proposed mechanism for the polymerization of γBL catalyzed by N,N’-diphenylurea/MeONa. Bond distances are exposed in the black dashed line. Reprinted with permission from []. Copyright (2018) American Chemical Society.
Figure 33. The free energy reaction profiles for different ureas and alkoxides. Reprinted with permission from []. Copyright (2018) American Chemical Society.

8. Acid-Catalyzed ROP

The most of organocatalysts have basic functionalities. This has made the polymerization of monomers containing carboxylic, amide, or other acidic functionalities considerably more difficult, if not impossible, while using traditional basic organocatalysts. Acid-catalyzed ROP is highly attractive for such monomers; however, only traditional substrates were experimentally studied in such reactions. The strong Brønsted acids are only effective ROP catalysts for specific substrates, such as 1,3-dioxepan-2-one [], and only several types of Brønsted and Lewis acids demonstrated the high efficiency in ROP of common cyclic esters, such as δVL, εCL, TMC, or LA. The ability of sulfonic acids [,,,,,,,,,,,], carboxylic acids [,,,], acidic phosphates [,,,], ROH/HCl/ether [,,], and Tf2NH [] to catalyze polymerization of the common lactones, cyclic carbonates, and lactides has been established, the mechanisms of the reactions were proposed but not explored in detail. In [], the thermochemistry of cationic ring-opening of TMC and 5-methylene-1,3-dioxan-2-one was estimated by quantum-chemical modeling at the HF/3-21G** level of theory, but the results of such modeling do not correspond to the subject of our review.
Ten years later, Bourissou et al. showed sulfonic acids as effective polymerization catalysts for TMC ROP [] and then proposed two reaction pathways (Scheme 9) to explain the formation of bimodal poly(TMC). The activated chain end (ACE) mechanism and activated monomer (AM) mechanism were discussed. Note that Baśko and Kubisa discussed such two mechanisms earlier for εCL and LA acid-catalyzed ROP []. However, the first theoretical studies of acid-catalyzed ROP were conducted until 2010.
Scheme 9. Two mechanisms for acid-catalyzed ROP of TMC: (a) Activated monomer mechanism (AM); and, (b) Activated chain end mechanism (ACE) [].
Maron et al. computationally studied the catalytic properties of CF3SO3H and MeSO3H in εCL ROP initiated by MeOH [] while using B3PW91 functional [,] and 6-31G(d,p) basis set []; the F atom was treated with a Stuttgart-Dresden pseudopotential []. The nucleophilic addition and ring-opening were analyzed and discussed separately. Three transition states were found for the insertion step, two energetically close bifunctional transition states TS2 and TS3 were found to be preferable by the value of 20 kcal/mol in comparison with monofunctional TS1 (Figure 34). The products of MeOH insertion were tetrahedral intermediates differed by the nature of the hydrogen bonds involved; these intermediates were close by energy (within 3.5 kcal/mol). The ring-opening step involved the cleavage of the endocyclic C–O bond. For this step, two transition states were located, TS4 (without the participance of the acid) and six-membered TS5 (with a alkylsulsonate bridge between the cyclic O atom and hydroxy group, Figure 34). The energy of the ring-opening TS5 was lower than transition states of the initiation step due to the assistance by CF3SO3H, so that the ring-opening was an easy process. The initial nucleophilic attack was the rate-determining step, and the overall free enthalpy change for εCL methanolysis was estimated by the value of −3 kcal/mol.
Figure 34. Free energy reaction profiles for alternative mechanisms of CF3SO3H-catalyzed methanolysis of εCL, and the structures of the key transition states. Reprinted with permission from []. Copyright (2010) American Chemical Society.
In summary, it was clear that sulfonic acid demonstrated a bifunctional behaviour, acting as a “proton shuttle” via its acidic hydrogen atom and basic oxygen atoms. A similar study was performed for MeSO3H; the energy reaction profile of this process was found to be similar to the reaction profile that is presented in Figure 34. The activation barriers that were predicted for MeSO3H and CF3SO3H (22.7 and 16.7 kcal/mol, respectively) were in the same range, despite a significant difference in acidity. The slightly higher values found for MeSO3H over CF3SO3H suggested that ring-opening polymerization of εCL should proceed faster when catalyzed by CF3SO3H, whereas MeSO3H was experimentally found to be more active []. Such difference between theoretical and experimental data was explained later [] (see below).
One year later, Bourissou, Maron et al. [] studied εCL ROP catalyzed by phosphoric acids and phosphoramidic acids (PA and PAA, respectively, Figure 35a). The polymerization experiments in toluene media resulted in obtaining poly(εCL) with narrow ÐM (1.07–1.15); the end-fragments of ROH initiator used (n-C5H11OH) were detected in NMR spectra of polymers; the values of Mn determined by end-group analysis and by SEC were in good agreement; thus, the ROP under these conditions can be considered as a living polymerization. DFT modeling of the reaction mechanisms were performed at B3PW91/6-31G(d,p) [,,] level of theory while using MeOH as an initiator. While taking the bifunctional character of PA and PAA catalysts into consideration, the authors studied and discussed the only bifunctional pathways that were qualitatively different for PA and PAA. Figure 35 also presents the key transition states for the insertion and ring-opening steps. The calculations predicted higher catalytic activity for PA (lower activation barrier), but PAA demonstrated slightly higher productivity in polymerization experiments. Despite this discrepancy, the modeling clearly emphasized the importance of cooperative activation; both systems were acting as proton shuttles.
Figure 35. The structures of phosphoric acids (PA) and phosphoramidic acids (PPA) (a); the geometries and relative energies of the key transition states for εCL methanolysis catalyzed by PA (b) and PAA (c). Reprinted with permission from []. Copyright (2010) American Chemical Society.
Coady et al. conducted a comprehensive computational study of possible reaction pathways comparing the sulfonic and carboxylic acids to gain better understanding of the mechanism of acid-catalyzed ROP of TMC []. Calculations were performed with GAMESS-US38 [] while using M11 density functional theory [] with the 6-311 + G (2d,p) basis set [], followed by single point energy calculations with the aug-cc-pVTZ basis set [,]; the impact of the solvent (CH2Cl2) was taken into account with the SMD (IEF-cPCM) method []. Computation suggested a counterintuitive mechanism: instead of full carbonyl protonation (Scheme 9), a pathway utilizing bifunctional activation was found (Scheme 10).
Scheme 10. Proposed acid-catalyzed mechanism for the ROP of cyclic carbonates [].
The acid molecule plays the role of such bifunctional activator that consists of an acidic group that is capable of hydrogen bond donation (H–A) and a Lewis basic atom capable of accepting a hydrogen bond (B). The first step is a nucleophilic attack (TS-1) with proton transfer to monomer and formation of I-1. At the second step (TS-1), the proton transfer to exocyclic oxygen atom occurs (I-2). The intermediate obtaining forms I-3 by overcoming rotational TS-3. The next step, ring-opening via TS-4, results in the product complex. At the final stage, the molecule of the product forms; this molecule acts as a reagent in the reaction with the next TMC molecule by the same reaction sequence (propagation stage). Figure 36 presents the calculated reaction profiles of the polymerization of TMC initiated by methanol and catalyzed by MeSO3H (MSA), CF3COOH (TFA), and triflic acid CF3SO3H. DFT modeling data were in poor agreement with the experimental results: triflic acid demonstrated the best activity; TFA was inactive as a catalyst.
Figure 36. Calculated reaction energy profiles for ROP of TMC initiated by methanol and catalyzed by MeSO3H (MSA), CF3COOH (TFA), or triflic acid CF3SO3H. Reprinted with permission from []. Copyright (2013) American Chemical Society.
The comparative DFT modeling of the activated chain end (ACE, Scheme 9) mechanism energetics resulted in the free activation energy value of 45 kcal/mol generally considered as too high for a viable mechanistic alternative. The ROP experiments with complete water removal resulted in a decrease of the high MW fraction of the polymer [], which invalidated the ACE mechanism that was proposed earlier [] and confirmed the results of DFT modeling.
The activated momoner mechanism of ROP had been detected not only for sulfonic acids, but also for carboxylic acids. Guo et al. [] studied the polymerization of δVL and εCL, catalyzed by γ-resorcylic acid (RA) or salicylic acid (SA, Figure 37), and performed the DFT modeling of the molecular structures of these acids at B3LYP/6-31++G [,,] level of theory. The calculations performed only allowed for the authors to explain higher acidity of RA by double hydrogen bonding (Figure 37); the probability of bifunctional mechanism was only mentioned but not discussed.
Figure 37. Hydrogen bonding in γ-resorcylic acid (RA) and salicylic acid (SA). Reprinted with permission from []. Copyright (2015) Royal Society of Chemistry.
It is also highly significant that the reaction of the Brønsted acids with lactones—if we do not view the acid catalyst as a bifunctional agent—results in the formation of unsaturated carboxylic acids. Haider et al. studied this issue in detail [] at the GGA-PW91 [] level of theory. The results of simulations demonstrates the linear correlation between the rate constants and the energies of the formation of oxocarbenium ions; these ions undergo nucleophilic attack of the water molecule not on carbonyl carbon, but on ω-carbon of the lactone ring, followed by dehydration (Scheme 11).
Scheme 11. Ring-opening/dehydration of lactones catalyzed by Brønsted acid in water [].

9. Phosphazenes: A Regrettable Lacuna in DFT Modeling

To date, DFT modeling was not applied for theoretical analysis and visualization of the mechanisms of ROP catalyzed by nitrogen–phosphorous hybrid organobases, such as phosphazene bases (PBs, Scheme 12), which possess a remarkably high basicity []. Wade et al. proposed the mechanistic concept of phosphazene-catalyzed polymerization of cyclic esters already in the first publication on the theme []. This mechanism involves the activation of the alcohol for the nucleophilic attack on the carbonyl group of cyclic esters without the activation of cyclic esters (Scheme 12).
Scheme 12. Postulated mechanism of the ROP of lactones using phosphazene BEMP as catalyst [].
A number of publications were devoted to the use of phosphazenes in ROP of cyclic esters [,,,,,,,,,]. For example, phosphazenes demonstrated high catalytic activity in δVL [] εCL [,], l-LA [] polymerization and block copolymerization of εCL with l-LA[], being initiated by alcohols [,] and amides [] as O–H and N–H acids, respectively (Scheme 13). Cyclic phosphazene (CTPB, Scheme 13) catalyzed living ROP of γBL [] and macrocyclic ω-pentadecalactone []. The phosphazene-modified adamantane-biphenylene-based framework was effectively used in δVL and εCL polymerization []. Very recently, bifunctional (thio)urea-phosphazene catalysts were used in ROP of d,l-LA, δVL, and εCL [].
Scheme 13. The examples of phosphazene-catalyzed ROP of cyclic esters.
Note that quantum chemical methods were successfully applied for the design of novel phosphazene catalysts in terms of Brønsted basicity [,]. We believe that DFT modeling could also be usefully extended to phosphazene-catalyzed ROP of different substrates.

10. Concluding Remarks

To date, different organocatalysts have been successfully used in the ROP of cyclic esters (actones, lactides, carbonates) and cyclic phosphates and phosphonates, which was the subject of this review. Starting from basic catalysts, such as DMAP and NHC, through the use of non-nucleophilic bases, such as DBU and phosphazenes, the researchers came to organocatalysts, providing the most gentle and energetically favourable donor-acceptor mechanism, namely, TBD—DBU/TU—thiourea and urea salts. Currently, the last group of the catalysts appears to be highly promising due to versatility [] and knowingly lack of toxicity. The role of DFT modeling in understanding of the ROP mechanisms and in the design of the novel ROP catalysts is crucial, comparative theoretical estimations of the ROP activation barriers for zwitterionic and donor-acceptor mechanisms resulted in the development of efficient catalytic systems that are able to activate both cyclic ester molecule and chain-end fragment.
Different methods have been applied for the DFT modeling of the polymerization of cyclic esters. In Table 1, we summarized the technical aspects of DFT modeling of the organocatalytic ROP. The use of one or other of the alternative optimization methods depends on both research objectives (e.g., comparison of the alternative reaction pathways, understanding of the causes of stereochemical induction, etc.) and presentation purposes (visualization of the reaction complex in context of the proposed reaction mechanism). The use of solvation model in DFT calculations, which is essential for zwitterionic processes, seems to be of less importance for the reactions in non-polar media occurring via hydrogen transfer within the reaction complex. A manifold increase of calculation time, while using actual solvent model, poses a problem for choosing between the completeness of the model and accuracy of geometries and energies of the stationary points and transition states. Hence, the use of the solvation model is a choice of the researchers, particularly in view of the fact that the pre-exponential Arrhenius factor and conformational effects have been totally ignored by most of the researchers, given the staggering complexity of considering these factors in DFT calculations.
Table 1. DFT methods used in the modeling of organocatalytic ROP of cyclic esters.
The main conclusion of the present review was to recognize that proton exchange and hydrogen bonding [] play a determining role in organocatalytic ROP of cyclic esters, especially for ROH-initiated processes. This is the difference between coordination ROP [] and organocatalytic ROP, as discussed in this review, with regard to DFT modeling. Such a difference between coordination and organocatalytic ROP also affects the character of the polymerization: for active coordination catalysts, the polymerization occurs as a living process, the polymerization grade Pn is determined by the monomer/catalyst ratio. In organocatalytic ROP, Pn is determined by the ratio of monomer and ROH initiator; therefore, the amount of the organocatalyst can be substantially minimized, as shown in Figure 1.
Besides, the traditional coordination-insertion ROP mechanism of the “living” polymerization implies the formation of linear polymers with narrow MWD if the activation barriers of ring-opening and transesterification substantially differ in energy. The zwitterionic mechanism becomes possible by using basic organocatalysts in the absence of proton donors, and DFT modeling compellingly demonstrates the relative ease of the formation of macrocyclic oligomers that are apparent polymerization products.
As can be seen, the results of DFT simulations of the reaction profiles are often unrelated with experimental data on catalytic activity in comparing different organocatalysts. Such a discrepancy can be attributed to the imperfections of the methods used in modeling (and usually the authors have been guilty of such an explanation). However, the direct correlations between activation energies and polymerization rates is not entirely correct, because ignoring the pre-exponential Arrhenius factor would not necessarily be equal for different catalysts. The role of the conformations of growing polymer chain, and other factors, facilitating or impeding the formation of catalytic complex, should be taken into account, especially for organocatalytic processes with relatively weak catalyst—monomer—activator interactions. Transesterification processes can become essential, notwithstanding their higher activation barriers if the transport of the substrate molecule is hindered, as was demonstrated in []. In view of the polymer architecture, such shift in the reaction might result in the formation of branched polymers in the case of ethylene phosphates’ ROP, and in broadening of MWD for common cyclic substrates, such as lactones and lactides. Moreover, in the course of ROP the shift from living to immortal polymerization, from chain-growth to step-growth reaction mode, can be occurred.
One of the main purposes for the use of organocatalysis in the synthesis of polyesters was to avoid metal complexes that may be toxic and difficult in the removal from polar polymers. At first glance, organocatalytic ROP allows for solving the problem of the purity of the materials attended for biomedical and microelectronics applications. DFT modeling that help us to understand the reaction mechanisms and, therefore, to outline the ways of the modern catalyst design, is a powerful tool in the development of novel catalysts and prospective materials. However, it should be noted that theoretical estimations of the applicability and effectiveness of one or the other organocatalyst is incomplete without the study of the toxicity of these catalysts [] and, more pertinently, of the possibility of direct covalent bonding between organocatalyst and polyester molecules, and side transformations of cyclic substrate molecules and polyester macromolecules under the action of organocatalysts. Of course, DFT modeling is unable to assist researchers in the study of toxicity, which is the area of biomedical research [], however, DFT methods may assist not only in understanding of the ROP mechanisms, but in excluding of the potentially undesirable reactions during polymer preparation and separation stages.
Polymer degradation and utilization is another prospective field for the application of DFT modeling. To date, only few publications are devoted to this issue [,,].
The clear progress in computer engineering and developing of the methods of quantum chemistry allows for the consideration of the DFT modeling as a common and useful technique for understanding and visualization of polymerization mechanisms. It is fair to say that the potential of the DFT is not being fully utilized in the design of new catalysts, processes, and materials, and we strongly hope that this review will help researchers to expand their scientific tools by the use of this simple and efficient method.

Author Contributions

Conceptualization, P.I.; Methodology, I.N. and P.I.; Writing—Original draft preparation, P.I.; Writing—Review and editing, I.N. and P.I.; Visualization, P.I.; Supervision, I.N.; Project administration, I.N.; Funding acquisition, I.N.

Funding

This research was funded by Russian Science Foundation, grant number 16-13-10344.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Penczek, S.; Cypryk, M.; Duda, A.; Kubisa, P.; Slomkowski, S. Living ring-opening polymerizations of heterocyclic monomers. Prog. Polym. Sci. 2007, 32, 247–282. [Google Scholar] [CrossRef]
  2. Nuyken, O.; Pask, S.D. Ring-Opening Polymerization-An Introductory Review. Polymers 2013, 5, 361–403. [Google Scholar] [CrossRef]
  3. Lecomte, P.; Jérôme, C. Recent Developments in Ring-Opening Polymerization of Lactones. Adv. Polym. Sci. 2012, 245, 173–218. [Google Scholar] [CrossRef]
  4. Guillaume, S.M.; Kirillov, E.; Sarazin, Y.; Carpentier, J.-F. Beyond Stereoselectivity, Switchable Catalysis: Some of the Last Frontier Challenges in Ring-Opening Polymerization of Cyclic Esters. Chem. Eur. J. 2015, 21, 7988–8003. [Google Scholar] [CrossRef]
  5. Arbaoui, A.; Redshaw, C. Metal catalysts for ε-caprolactone polymerisation. Polym. Chem. 2010, 1, 801–826. [Google Scholar] [CrossRef]
  6. Penczek, S.; Pretula, J.B.; Kaluzynski, K.; Lapienis, G. Polymers with Esters of Phosphoric Acid Units: From Synthesis, Models of Biopolymers to Polymer—Inorganic Hybrids. Isr. J. Chem. 2012, 52, 306–319. [Google Scholar] [CrossRef]
  7. Wang, Y.-C.; Yuan, Y.-Y.; Du, J.-Z.; Yang, X.-Z.; Wang, J. Recent progress in polyphosphoesters: From controlled synthesis to biomedical applications. Macromol. Biosci. 2009, 9, 1154–1164. [Google Scholar] [CrossRef]
  8. Yao, K.; Tang, C. Controlled Polymerization of Next-Generation Renewable Monomers and Beyond. Macromolecules 2013, 46, 1689–1712. [Google Scholar] [CrossRef]
  9. Jérôme, C.; Lecomte, P. Recent advances in the synthesis of aliphatic polyesters by ring-opening polymerization. Adv. Drug Deliv. Rev. 2008, 60, 1056–1076. [Google Scholar] [CrossRef]
  10. Ruengdechawiwat, S.; Somsunan, R.; Molloy, R.; Siripitayananon, J.; Franklin, V.J.; Topham, P.D.; Tighe, B.J. Controlled Synthesis and Processing of a Poly(L-lactide-co-ε-caprolactone) Copolymer for Biomedical Use as an Absorbable Monofilament Surgical Suture. Adv. Mater. Res. 2014, 894, 172–176. [Google Scholar] [CrossRef]
  11. Wang, L.; Poirier, V.; Ghiotto, F.; Bochmann, M.; Cannon, R.D.; Carpentier, J.-F.; Sarazin, Y. Kinetic Analysis of the Immortal Ring-Opening Polymerization of Cyclic Esters: A Case Study with Tin (II) Catalysts. Macromolecules 2014, 47, 2574–2584. [Google Scholar] [CrossRef]
  12. 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]
  13. Kowalski, A.; Duda, A.; Penczek, S. Polymerization of l,l-Lactide Initiated by Aluminum Isopropoxide Trimer or Tetramer. Macromolecules 1998, 31, 2114–2122. [Google Scholar] [CrossRef]
  14. Duda, A.; Penczek, S. On the difference of reactivities of various aggregated forms of aluminium triisopropoxide in initiating ring-opening polymerizations. Macromol. Rapid Commun. 1995, 16, 67–76. [Google Scholar] [CrossRef]
  15. Duda, A.; Penczek, S. Polymerization of ε-Caprolactone Initiated by Aluminum Isopropoxide Trimer and/or Tetramer. Macromolecules 1995, 28, 5981–5992. [Google Scholar] [CrossRef]
  16. Dubois, P.; Jacobs, C.; Jérôme, R.; Teyssie, P. Macromolecular engineering of polylactones and polylactides. 4. Mechanism and kinetics of lactide homopolymerization by aluminum isopropoxide. Macromolecules 1991, 24, 2266–2270. [Google Scholar] [CrossRef]
  17. Dechy-Cabaret, O.; Martin-Vaca, B.; Bourissou, D. Controlled Ring-Opening Polymerization of Lactide and Glycolide. Chem. Rev. 2004, 104, 6147–6176. [Google Scholar] [CrossRef]
  18. Sarazin, Y.; Carpentier, J.-F. Discrete Cationic Complexes for Ring-Opening Polymerization Catalysis of Cyclic Esters and Epoxides. Chem. Rev. 2015, 115, 3564–3614. [Google Scholar] [CrossRef]
  19. Copéret, C.; Allouche, F.; Chan, K.W.; Conley, M.P.; Delley, M.F.; Fedorov, A.; Moroz, I.B.; Mougel, V.; Pucino, M.; Searles, K.; et al. Bridging the Gap between Industrial and Well-Defined Supported Catalysts. Angew. Chem. Int. Ed. 2018, 57, 6398–6440. [Google Scholar] [CrossRef]
  20. Tian, H.; Tang, Z.; Zhuang, X.; Chen, X.; Jing, X. Biodegradable synthetic polymers: Preparation, functionalization and biomedical application. Prog. Polym. Sci. 2012, 37, 237–280. [Google Scholar] [CrossRef]
  21. Nachtergael, A.; Coulembier, O.; Dubois, P.; Helvenstein, M.; Duez, P.; Blankert, B.; Mespouille, L. Organocatalysis paradigm revisited: Are metal-free catalysts really harmless? Biomacromolecules 2015, 16, 507–514. [Google Scholar] [CrossRef]
  22. Kamber, N.E.; Lohmeijer, B.G.G.; Hedrick, J.L. Organocatalytic Ring-Opening Polymerization. Chem. Rev. 2007, 107, 5813–5840. [Google Scholar] [CrossRef]
  23. Kiesewetter, M.K.; Shin, E.J.; Hedrick, J.L.; Waymouth, R.M. Organocatalysis: Opportunities and Challenges for Polymer Synthesis. Macromolecules 2010, 43, 2093–2107. [Google Scholar] [CrossRef]
  24. Bossion, A.; Heifferon, K.V.; Meabe, L.; Zivic, N.; Taton, D.; Hedrick, J.L.; Long, T.E.; Sardon, H. Opportunities for organocatalysis in polymer synthesis via step-growth methods. Prog. Polym. Sci. 2019, 90, 164–210. [Google Scholar] [CrossRef]
  25. Fèvre, M.; Vignolle, J.; Gnanou, Y.; Taton, D. 4.06-Organocatalyzed Ring-Opening Polymerizations. Polym. Sci. A Compr. Ref. 2012, 4, 67–115. [Google Scholar] [CrossRef]
  26. Fevre, M.; Pinaud, J.; Gnanou, Y.; Vignolle, J.; Taton, D. N-Heterocyclic carbenes (NHCs) as organocatalysts and structural components in metal-free polymer synthesis. Chem. Soc. Rev. 2013, 42, 2142–2172. [Google Scholar] [CrossRef] [PubMed]
  27. Dove, A.P. Organic Catalysis for Ring-Opening Polymerization. ACS Macro Lett. 2012, 1, 1409–1412. [Google Scholar] [CrossRef]
  28. Naumann, S.; Dove, A.P. N-Heterocyclic carbenes as organocatalysts for polymerizations: Trends and frontiers. Polym. Chem. 2015, 6, 3185–3200. [Google Scholar] [CrossRef]
  29. Cheong, P.H.-Y.; Legault, C.Y.; Um, J.M.; Celebi-Olcum, N.; Houk, K.N. Quantum Mechanical Investigations of Organocatalysis: Mechanisms, Reactivities, and Selectivities. Chem. Rev. 2011, 111, 5042–5137. [Google Scholar] [CrossRef]
  30. Jones, G.O. Contributions of quantum chemistry to the development of ring opening polymerizations and chemical recycling. Tetrahedron 2019, 75, 2047–2055. [Google Scholar] [CrossRef]
  31. Ruipérez, F. Application of quantum chemical methods in polymer chemistry. Int. Rev. Phys. Chem. 2019, 38, 343–403. [Google Scholar] [CrossRef]
  32. Nifant’ev, I.; Ivchenko, P. Coordination Ring-Opening Polymerization of Cyclic Esters: A Critical Overview of DFT Modeling and Visualization of the Reaction Mechanisms. Molecules 2019, 24, 4117. [Google Scholar] [CrossRef]
  33. Olsén, P.; Odelius, K.; Albertsson, A.-C. Thermodynamic Presynthetic Considerations for Ring-Opening Polymerization. Biomacromolecules 2016, 17, 699–709. [Google Scholar] [CrossRef] [PubMed]
  34. Huisgen, R.; Ott, H. Die konfiguration der carbonestergruppe und die sondereigenschaften der lactone. Tetrahedron 1959, 6, 253–267. [Google Scholar] [CrossRef]
  35. Wang, X.; Houk, K.N. Theoretical elucidation of the origin of the anomalously high acidity of Meldrum’s acid. J. Am. Chem. Soc. 1988, 110, 1870–1872. [Google Scholar] [CrossRef]
  36. Wiberg, K.B.; Wong, M.W. Solvent effects. 4. Effect of solvent on the E/Z energy difference for methyl formate and methyl acetate. J. Am. Chem. Soc. 1993, 115, 1078–1084. [Google Scholar] [CrossRef]
  37. Houk, K.N.; Jabbari, A.; Hall, H.K.; Alemán, C. Why δ-Valerolactone Polymerizes and γ-Butyrolactone Does Not. J. Org. Chem. 2008, 73, 2674–2678. [Google Scholar] [CrossRef]
  38. Leitão, M.L.P.; Pilcher, G.; Meng-Yan, Y.; Brown, J.M.; Conn, A.D. Enthalpies of combustion of γ-butyrolactone, γ-valerolactone, and δ-valerolactone. J. Chem. Thermodyn. 1990, 22, 885–891. [Google Scholar] [CrossRef]
  39. 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]
  40. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  41. Hariharan, P.C.; Pople, J.A. The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–222. [Google Scholar] [CrossRef]
  42. Montgomery, J.A., Jr.; Frisch, M.J. A complete basis set model chemistry. VI. Use of density functional geometries and frequencies. J. Chem. Phys. 1999, 110, 2822–2827. [Google Scholar] [CrossRef]
  43. Montgomery, J.A., Jr.; Frisch, M.J.; Ochterski, J.W. A complete basis set model chemistry. VII. Use of the minimum population localization method. J. Chem. Phys. 2000, 112, 6532–6542. [Google Scholar] [CrossRef]
  44. Alemán, C.; Casanovas, J.; Zanuy, D.; Hall, H.K. Systematic Evaluation of the Conformational Properties of Aliphatic ω-Hydroxy Acids. J. Org. Chem. 2005, 70, 2950–2956. [Google Scholar] [CrossRef] [PubMed]
  45. Alemán, C.; Casanovas, J.; Hall, H.K. Systematic Evaluation of the Conformational Properties of Aliphatic ω-Methoxy Methyl Esters. J. Org. Chem. 2005, 70, 7731–7736. [Google Scholar] [CrossRef] [PubMed]
  46. Alemán, C.; Betran, O.; Casanovas, J.; Houk, K.N.; Hall, H.K., Jr. Thermodynamic Control of the Polymerizability of Five-, Six-, and Seven-Membered Lactones. J. Org. Chem. 2009, 74, 6237–6244. [Google Scholar] [CrossRef] [PubMed]
  47. Alemán, C.; Bertran, O.; Houk, K.N.; Padías, A.B.; Hall, H.K., Jr. Thermodynamic and stereochemical aspects of the polymerizability of glycolide and lactide. Theor. Chem. Acc. 2012, 131, 1133. [Google Scholar] [CrossRef]
  48. Nederberg, F.; Connor, E.F.; Möller, M.; Glauser, T.; Hedrick, J.L. New paradigms for organic catalysts: The first organocatalytic living polymerization. Angew. Chem. Int. Ed. 2001, 40, 2712–2715. [Google Scholar] [CrossRef]
  49. Feng, H.; Dong, C.-M. Preparation and characterization of chitosan-graft-poly (ϵ-caprolactone) with an organic catalyst. J. Polym. Sci. Part A Polym. Chem. 2006, 44, 5353–5361. [Google Scholar] [CrossRef]
  50. Bonduelle, C.; Martín-Vaca, B.; Cossío, F.P.; Bourissou, D. Monomer versus Alcohol Activation in the 4-Dimethylaminopyridine-Catalyzed Ring-Opening Polymerization of Lactide and Lactic O-Carboxylic Anhydride. Chem. Eur. J. 2008, 14, 5304–5312. [Google Scholar] [CrossRef]
  51. Rassolov, V.A.; Pople, J.A.; Ratner, M.A.; Windus, T.L. 6-31G * basis set for atoms K through Zn. J. Chem. Phys. 1998, 109, 1223–1229. [Google Scholar] [CrossRef]
  52. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  53. Nogueira, G.; Favrelle, A.; Bria, M.; Prates Ramalho, J.P.; Mendes, P.J.; Valente, A.; Zinck, P. Adenine as an organocatalyst for the ring-opening polymerization of lactide: Scope, mechanism and access to adenine-functionalized polylactide. React. Chem. Eng. 2016, 1, 508–520. [Google Scholar] [CrossRef]
  54. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other fun. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  55. Katiyar, V.; Nanavati, H. Ring-opening polymerization of L-lactide using N-heterocyclic molecules: Mechanistic, kinetics and DFT studies. Polym. Chem. 2010, 1, 1491–1500. [Google Scholar] [CrossRef]
  56. Dove, A.P.; Pratt, R.C.; Lohmeijer, B.G.G.; Culkin, D.A.; Hagberg, E.C.; Nyce, G.W.; Waymouth, R.M.; Hedrick, J.L. N-Heterocyclic carbenes: Effective organic catalysts for living polymerization. Polymer 2006, 47, 4018–4025. [Google Scholar] [CrossRef]
  57. Vogt, M.; Bennett, J.E.; Huang, Y.; Wu, C.; Schneider, W.F.; Brennecke, J.F.; Ashfeld, B.L. Solid-State Covalent Capture of CO2 by Using N-Heterocyclic Carbenes. Chem. Eur. J. 2013, 19, 11134–11138. [Google Scholar] [CrossRef] [PubMed]
  58. Schweizer, J.I.; Sturm, A.G.; Porsch, T.; Berger, M.; Bolte, M.; Auner, N.; Holthausen, M.C. Reactions of Si2Br6 with N-Heterocyclic Carbenes. Z. Anorg. Allg. Chem. 2018, 644, 982–988. [Google Scholar] [CrossRef]
  59. Tukov, A.A.; Normand, A.T.; Nechaev, M.S. N-heterocyclic carbenes bearing two, one and no nitrogen atoms at the ylidenecarbon: Insight from theoretical calculations. Dalton Trans. 2009, 7015–7028. [Google Scholar] [CrossRef]
  60. Hopkinson, M.N.; Richter, C.; Schedler, M.; Glorius, F. An overview of N-heterocyclic carbenes. Nature 2014, 510, 485–496. [Google Scholar] [CrossRef]
  61. Huynh, H.V. Electronic Properties of N-Heterocyclic Carbenes and Their Experimental Determination. Chem. Rev. 2018, 118, 9457–9492. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, N.; Xu, J.; Lee, J.K. The importance of N-heterocyclic carbene basicity in organocatalysis. Org. Biomol. Chem. 2018, 16, 8230–8244. [Google Scholar] [CrossRef] [PubMed]
  63. Connor, E.F.; Nyce, G.W.; Myers, M.; Möck, A.; Hedrick, J.L. First Example of N-Heterocyclic Carbenes as Catalysts for Living Polymerization: Organocatalytic Ring-Opening Polymerization of Cyclic Esters. J. Am. Chem. Soc. 2002, 124, 914–915. [Google Scholar] [CrossRef] [PubMed]
  64. Csihony, S.; Culkin, D.A.; Sentman, A.C.; Dove, A.P.; Waymouth, R.M.; Hedrick, J.L. Single-Component Catalyst/Initiators for the Organocatalytic Ring-Opening Polymerization of Lactide. J. Am. Chem. Soc. 2005, 127, 9079–9084. [Google Scholar] [CrossRef] [PubMed]
  65. Culkin, D.A.; Jeong, W.; Csihony, S.; Gomez, E.D.; Balsara, N.P.; Hedrick, J.L.; Waymouth, R.M. Zwitterionic Polymerization of Lactide to Cyclic Poly(Lactide) by Using N-Heterocyclic Carbene Organocatalysts. Angew. Chem. Int. Ed. 2007, 46, 2627–2630. [Google Scholar] [CrossRef] [PubMed]
  66. Kamber, N.E.; Jeong, W.; Gonzalez, S.; Hedrick, J.L.; Waymouth, R.M. N-Heterocyclic Carbenes for the Organocatalytic Ring-Opening Polymerization of e-Caprolactone. Macromolecules 2009, 42, 1634–1639. [Google Scholar] [CrossRef]
  67. Nyce, G.W.; Glauser, T.; Connor, E.F.; Möck, A.; Waymouth, R.M.; Hedrick, J.L. In situ generation of carbenes: A general and versatile platform for organocatalytic living polymerization. J. Am. Chem. Soc. 2003, 125, 3046–3056. [Google Scholar] [CrossRef]
  68. Coulembier, O.; Lohmeijer, B.G.G.; Dove, A.P.; Pratt, R.C.; Mespouille, L.; Culkin, D.A.; Benight, S.J.; Dubois, P.; Waymouth, R.M.; Hedrick, J.L. Alcohol Adducts of N-Heterocyclic Carbenes: Latent Catalysts for the Thermally-Controlled Living Polymerization of Cyclic Esters. Macromolecules 2006, 39, 5617–5628. [Google Scholar] [CrossRef]
  69. Bhatia, R.; Gaur, J.; Jain, S.; Lal, A.; Tripathi, B.; Attri, P.; Kaushik, N.K. Synthetic Strategies for Free & Stable N-Heterocyclic Carbenes and Their Precursors. Mini-Rev. Org. Chem. 2013, 10, 180–197. [Google Scholar] [CrossRef]
  70. Flanigan, D.M.; Romanov-Michailidis, F.; White, N.A.; Rovis, T. Organocatalytic Reactions Enabled by N-Heterocyclic Carbenes. Chem. Rev. 2015, 115, 9307–9387. [Google Scholar] [CrossRef]
  71. Fèvre, M.; Pinaud, J.; Leteneur, A.; Gnanou, Y.; Vignolle, J.; Taton, D.; Miqueu, K.; Sotiropoulos, J.-M. Imidazol (in) ium Hydrogen Carbonates as a Genuine Source of N-Heterocyclic Carbenes (NHCs): Applications to the Facile Preparation of NHC Metal Complexes and to NHC-Organocatalyzed Molecular and Macromolecular Syntheses. J. Am. Chem. Soc. 2012, 134, 6776–6784. [Google Scholar] [CrossRef]
  72. Kricheldorf, H.R. Cyclic polymers: Synthetic strategies and physical properties. J. Polym. Sci. Part A Polym. Chem. 2010, 48, 251–284. [Google Scholar] [CrossRef]
  73. Hoskins, J.N.; Grayson, S.M. Cyclic polyesters: Synthetic approaches and potential applications. Polym. Chem. 2011, 2, 289–299. [Google Scholar] [CrossRef]
  74. Stukenbroeker, T.S.; Solis-Ibarra, D.; Waymouth, R.M. Synthesis and topological trapping of cyclic poly (alkylene phosphates). Macromolecules 2014, 47, 8224–8230. [Google Scholar] [CrossRef]
  75. Acharya, A.K.; Chang, Y.A.; Jones, G.O.; Rice, J.E.; Hedrick, J.L.; Horn, H.W.; Waymouth, R.M. Experimental and Computational Studies on the Mechanism of Zwitterionic Ring-Opening Polymerization of δ-Valerolactone with N-Heterocyclic Carbenes. J. Phys. Chem. B 2014, 118, 6553–6560. [Google Scholar] [CrossRef]
  76. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: A critical analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef]
  77. 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. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  78. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  79. Jones, G.O.; Chang, Y.A.; Horn, H.W.; Acharya, A.K.; Rice, J.E.; Hedrick, J.L.; Waymouth, R.M. N-Heterocyclic Carbene-Catalyzed Ring Opening Polymerization of ε-Caprolactone with and without Alcohol Initiators: Insights from Theory and Experiment. J. Phys. Chem. B 2015, 119, 5728–5737. [Google Scholar] [CrossRef]
  80. Miertuš, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilizaion of AB initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55, 117–129. [Google Scholar] [CrossRef]
  81. Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar] [CrossRef]
  82. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem. 2003, 24, 669–681. [Google Scholar] [CrossRef] [PubMed]
  83. Falivene, L.; Cavallo, L. Guidelines To Select the N-Heterocyclic Carbene for the Organopolymerization of Monomers with a Polar Group. Macromolecules 2017, 50, 1394–1401. [Google Scholar] [CrossRef]
  84. Aldeco-Perez, E.; Rosenthal, A.J.; Donnadieu, B.; Parameswaran, P.; Frenking, G.; Bertrand, G. Isolation of a C5-Deprotonated Imidazolium, a Crystalline “Abnormal” N-Heterocyclic Carbene. Science 2009, 326, 556–559. [Google Scholar] [CrossRef] [PubMed]
  85. Ung, G.; Bertrand, G. Stability and Electronic Properties of Imidazole-Based Mesoionic Carbenes. Chem. Eur. J. 2011, 17, 8269–8272. [Google Scholar] [CrossRef] [PubMed]
  86. Sen, T.K.; Sau, S.C.; Mukherjee, A.; Modak, A.; Mandal, S.K.; Koley, D. Introduction of abnormal N-heterocyclic carbene as an efficient organocatalyst: Ring opening polymerization of cyclic esters. Chem. Commun. 2011, 47, 11972–11974. [Google Scholar] [CrossRef]
  87. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 39, 3098–3100. [Google Scholar] [CrossRef]
  88. Perdew, J.P. Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 1986, 33, 8822–8824. [Google Scholar] [CrossRef]
  89. Lohmeijer, B.G.G.; Dubois, G.; Leibfarth, F.; Pratt, R.C.; Nederberg, F.; Nelson, A.; Waymouth, R.M.; Wade, C.; Hedrick, J.L. Organocatalytic Living Ring-Opening Polymerization of Cyclic Carbosiloxanes. Org. Lett. 2006, 8, 4683–4686. [Google Scholar] [CrossRef]
  90. Pratt, R.C.; Lohmeijer, B.G.G.; Long, D.A.; Waymouth, R.M.; Hedrick, J.L. Triazabicyclodecene: A Simple Bifunctional Organocatalyst for Acyl Transfer and Ring-Opening Polymerization of Cyclic Esters. J. Am. Chem. Soc. 2006, 128, 4556–4557. [Google Scholar] [CrossRef]
  91. Simón, L.; Goodman, J.M. The mechanism of TBD-catalyzed ring-opening polymerization of cyclic esters. J. Org. Chem. 2007, 72, 9656–9662. [Google Scholar] [CrossRef] [PubMed]
  92. Clark, T.; Chandrasekhar, J.; Spitznagel, G.W.; Von Ragué Schleyer, P. Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21+G basis set for first-row elements, Li–F. J. Comput. Chem. 1983, 4, 294–301. [Google Scholar] [CrossRef]
  93. Gill, P.M.W.; Johnson, B.G.; Pople, J.A.; Frisch, M.J. The performance of the Becke-Lee-Yang-Parr (B-LYP) density functional theory with various basis sets. Chem. Phys. Lett. 1992, 197, 499–505. [Google Scholar] [CrossRef]
  94. Nifant’ev, I.; Shlyakhtin, A.; Bagrov, V.; Lozhkin, B.; Zakirova, G.; Ivchenko, P.; Legon’kova, O. Theoretical and experimental studies of 1,5,7-triazabicyclo[4.4.0]dec-5-ene-catalyzed ring opening/ring closure reaction mechanism for 5-, 6- and 7-membered cyclic esters and carbonates. React. Kinet. Mech. Cat. 2016, 117, 447–476. [Google Scholar] [CrossRef]
  95. Jaffredo, C.G.; Carpentier, J.-F.; Guillaume, S.M. Controlled ROP of β -Butyrolactone Simply Mediated by Amidine, Guanidine, and Phosphazene Organocatalysts. Macromol. Rapid Commun. 2012, 33, 1938–1944. [Google Scholar] [CrossRef]
  96. Chuma, A.; Horn, H.W.; Swope, W.C.; Pratt, R.C.; Zhang, L.; Lohmeijer, B.G.G.; Wade, C.G.; Waymouth, R.M.; Hedrick, J.L.; Rice, J.E. The Reaction Mechanism for the Organocatalytic Ring-Opening Polymerization of L-Lactide Using a Guanidine-Based Catalyst: Hydrogen-Bonded or Covalently Bound? J. Am. Chem. Soc. 2008, 130, 6749–6754. [Google Scholar] [CrossRef]
  97. Lynch, B.J.; Fast, P.L.; Harris, M.; Truhlar, D.G. Adiabatic Connection for Kinetics. J. Phys. Chem. A 2000, 104, 4811–4815. [Google Scholar] [CrossRef]
  98. 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]
  99. Stukenbroeker, T.S.; Bandar, J.S.; Zhang, X.; Lambert, T.H.; Waymouth, R.M. Cyclopropenimine Superbases: Competitive Initiation Processes in Lactide Polymerization. ACS Macro Lett. 2015, 4, 853–856. [Google Scholar] [CrossRef]
  100. Pascual, A.; Sardón, H.; Ruipérez, F.; Gracia, R.; Sudam, P.; Veloso, A.; Mecerreyes, D. Experimental and computational studies of ring-opening polymerization of ethylene brassylate macrolactone and copolymerization with ε-caprolactone and TBD-guanidine organic catalyst. J. Polym. Chem. A Polym. Sci. 2015, 53, 552–561. [Google Scholar] [CrossRef]
  101. Gregory, G.L.; Jenisch, L.M.; Charles, B.; Kociok-Köhn, G.; Buchard, A. Polymers from Sugars and CO2: Synthesis and Polymerization of a d-Mannose-Based Cyclic Carbonate. Macromolecules 2016, 49, 7165–7169. [Google Scholar] [CrossRef]
  102. Chai, J.-D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom—Atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef] [PubMed]
  103. Gregory, G.L.; Kociok-Köhn, G.; Buchard, A. Polymers from sugars and CO2: Ring-opening polymerisation and copolymerisation of cyclic carbonates derived from 2-deoxy-d-ribose. Polym. Chem. 2017, 8, 2093–2104. [Google Scholar] [CrossRef]
  104. Steinbach, T.; Wurm, F.R. Poly(phosphoester)s: A new platform for degradable polymers. Angew. Chem. Int. Ed. 2015, 54, 6098–6108. [Google Scholar] [CrossRef] [PubMed]
  105. Yilmaz, Z.E.; Jeérôme, C. Polyphosphoesters: New trends in synthesis and drug delivery applications. Macromol. Biosci. 2016, 16, 1745–1761. [Google Scholar] [CrossRef] [PubMed]
  106. Bauer, K.N.; Tee, H.T.; Velencoso, M.M.; Wurm, F.R. Main-chain poly(phosphoester)s: History, syntheses, degradation, bio-and flame-retardant applications. Prog. Polym. Sci. 2017, 73, 61–122. [Google Scholar] [CrossRef]
  107. Ashkenazi, N.; Zade, S.S.; Segall, Y.; Karton, Y.; Bendikov, M. Selective site controlled nucleophilic attacks in 5-membered ring phosphate esters: Unusual C–O vs. common P–O bond cleavage. Chem. Commun. 2005, 5879–5881. [Google Scholar] [CrossRef]
  108. Xia, F.; Zhu, H. Alkaline hydrolysis of ethylene phosphate: An ab initio study by supermolecule model and polarizable continuum approach. J. Comput. Chem. 2011, 32, 2545–2554. [Google Scholar] [CrossRef]
  109. Nifant’ev, I.E.; Shlyakhtin, A.V.; Tavtorkin, A.N.; Kosarev, M.A.; Gavrilov, D.E.; Komarov, P.D.; Ilyin, S.O.; Karchevsky, S.G.; Ivchenko, P.V. Mechanistic study of transesterification in TBD-catalyzed ring-opening polymerization of methyl ethylene phosphate. Eur. Polym. J. 2019, 118, 393–403. [Google Scholar] [CrossRef]
  110. Nifant’ev, I.E.; Shlyakhtin, A.V.; Kosarev, M.A.; Komarov, P.D.; Karchevsky, S.G.; Ivchenko, P.V. Data for theoretical study of the mechanisms of ring-opening polymerization of methyl ethylene phosphate. Data Brief 2019, 26, 104431. [Google Scholar] [CrossRef]
  111. Sosa, C.; Andzelm, J.; Elkin, B.C.; Wimmer, E.; Dobbs, K.D.; Dixon, D.A. A local density functional study of the structure and vibrational frequencies of molecular transition-metal compounds. J. Phys. Chem. 1992, 96, 6630–6636. [Google Scholar] [CrossRef]
  112. Godbout, N.; Salahub, D.R.; Andzelm, J.; Wimmer, E. Optimization of Gaussian-type basis sets for local spin density functional calculations. Part I. Boron through neon, optimization technique and validation. Can. J. Chem. 1992, 70, 560–571. [Google Scholar] [CrossRef]
  113. Clément, B.; Grignard, B.; Koole, L.; Jérome, C.; Lecomte, P. Metal-free strategies for the synthesis of functional and well-defined polyphosphoesters. Macromolecules 2012, 45, 4476–4486. [Google Scholar] [CrossRef]
  114. Steinbach, T.; Schröder, R.; Ritz, S.; Wurm, F.R. Microstructure analysis of biocompatible phosphoester copolymers. Polym. Chem. 2013, 4, 4469–4479. [Google Scholar] [CrossRef]
  115. Nifant’ev, I.E.; Shlyakhtin, A.V.; Bagrov, V.V.; Komarov, P.D.; Kosarev, M.A.; Tavtorkin, A.N.; Minyaev, M.E.; Roznyatovsky, V.A.; Ivchenko, P.V. Controlled ring-opening polymerisation of cyclic phosphates, phosphonates and phosphoramidates catalysed by hereroleptic BHT-alkoxy magnesium complexes. Polym. Chem. 2017, 8, 6806–6816. [Google Scholar] [CrossRef]
  116. Zhang, L.; Pratt, R.C.; Nederberg, F.; Horn, H.W.; Rice, J.E.; Waymouth, R.M.; Wade, C.G.; Hedrick, J.L. Acyclic Guanidines as Organic Catalysts for Living Polymerization of Lactide. Macromolecules 2010, 43, 1660–1664. [Google Scholar] [CrossRef]
  117. Eisenreich, F.; Viehmann, P.; Müller, F.; Hecht, S. Electronic Activity Tuning of Acyclic Guanidines for Lactide Polymerization. Macromolecules 2015, 48, 8729–8732. [Google Scholar] [CrossRef]
  118. Nederberg, F.; Lohmeijer, B.G.G.; Leibfarth, F.; Pratt, R.C.; Choi, J.; Dove, A.P.; Waymouth, R.M.; Hedrick, J.L. Organocatalytic Ring Opening Polymerization of Trimethylene Carbonate. Biomacromolecules 2007, 8, 153–160. [Google Scholar] [CrossRef]
  119. Lohmeijer, B.G.G.; Pratt, R.C.; Leibfarth, F.; Logan, J.W.; Long, D.A.; Dove, A.P.; Nederberg, F.; Choi, J.; Wade, C.; Waymouth, R.M.; et al. Guanidine and Amidine Organocatalysts for Ring-Opening Polymerization of Cyclic Esters. Macromolecules 2006, 39, 8574–8583. [Google Scholar] [CrossRef]
  120. Iwasaki, Y.; Yamaguchi, E. Synthesis of well-defined thermoresponsive polyphosphoester macroinitiators using organocatalysts. Macromolecules 2010, 43, 2664–2666. [Google Scholar] [CrossRef]
  121. Zhai, X.; Huang, W.; Liu, J.; Pang, Y.; Zhu, X.; Zhou, Y.; Yan, D. Micelles from amphiphilic block copolyphosphates for drug delivery. Macromol. Biosci. 2011, 11, 1603–1610. [Google Scholar] [CrossRef] [PubMed]
  122. Zhang, S.; Li, A.; Zou, J.; Lin, L.Y.; Wooley, K.L. Facile synthesis of clickable, water-soluble, and degradable polyphosphoesters. ACS Macro Lett. 2012, 1, 328–333. [Google Scholar] [CrossRef] [PubMed]
  123. Brown, H.A.; De Crisci, A.G.; Hedrick, J.L.; Waymouth, R.M. Amidine-Mediated Zwitterionic Polymerization of Lactide. ACS Macro Lett. 2012, 1, 1113–1115. [Google Scholar] [CrossRef]
  124. Zhao, Y.; Truhlar, D.G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157–167. [Google Scholar] [CrossRef] [PubMed]
  125. Coady, D.J.; Fukushima, K.; Horn, H.W.; Rice, J.E.; Hedrick, J.L. Catalytic insights into acid/base conjugates: Highly selective bifunctional catalysts for the ring-opening polymerization of lactide. Chem. Commun. 2011, 47, 3105–3107. [Google Scholar] [CrossRef] [PubMed]
  126. Schäfer, A.; Horn, H.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets for atoms Li to Kr. J. Chem. Phys. 1992, 97, 2571–2577. [Google Scholar] [CrossRef]
  127. Woon, D.E.; Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. V. Core-valence basis sets for boron through neon. J. Chem. Phys. 1995, 103, 4572–4585. [Google Scholar] [CrossRef]
  128. Dharmaratne, N.U.; Pothupitiya, J.U.; Kiesewetter, M.K. The mechanistic duality of (thio)urea organocatalysts for ring-opening polymerization. Org. Biomol. Chem. 2019, 17, 3305–3313. [Google Scholar] [CrossRef]
  129. Dove, A.P.; Pratt, R.C.; Lohmeijer, B.G.G.; Waymouth, R.M.; Hedrick, J.L. Thiourea-Based Bifunctional Organocatalysis: Supramolecular Recognition for Living Polymerization. J. Am. Chem. Soc. 2005, 127, 13798–13799. [Google Scholar] [CrossRef]
  130. Zhu, R.-X.; Wang, R.-X.; Zhang, D.-J.; Liu, C.-B. A Density Functional Theory Study on the Ring-Opening Polymerization of d-Lactide Catalyzed by a Bifunctional-Thiourea Catalyst. Aust. J. Chem. 2009, 62, 157–164. [Google Scholar] [CrossRef]
  131. Frisch, M.J.; Pople, J.A. Self-consistent molecular orbital methods 25. Supplementary functions for Gaussian basis sets. J. Chem. Phys. 1984, 80, 3265–3269. [Google Scholar] [CrossRef]
  132. Coady, D.J.; Engler, A.C.; Horn, H.W.; Bajjuri, K.M.; Fukushima, K.; Jones, G.O.; Nelson, A.; Rice, J.E.; Hedrick, J.L. Catalyst Chelation Effects in Organocatalyzed Ring-Opening Polymerization of Lactide. ACS Macro Lett. 2012, 1, 19–22. [Google Scholar] [CrossRef]
  133. Kazakov, O.I.; Kiesewetter, M.K. Cocatalyst Binding Effects in Organocatalytic Ring-Opening Polymerization of l-Lactide. Macromolecules 2015, 48, 6121–6126. [Google Scholar] [CrossRef]
  134. Zhang, X.; Jones, G.O.; Hedrick, J.L.; Waymouth, R.M. Fast and selective ring-opening polymerizations by alkoxides and thioureas. Nat. Chem. 2016, 8, 1047–1053. [Google Scholar] [CrossRef]
  135. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  136. Kendall, R.A.; Dunning, T.H., Jr. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef]
  137. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  138. Lin, B.; Waymouth, R.M. Urea Anions: Simple, Fast, and Selective Catalysts for Ring-Opening Polymerizations. J. Am. Chem. Soc. 2017, 139, 1645–1652. [Google Scholar] [CrossRef]
  139. Shen, Y.; Zhao, Z.; Li, Y.; Liu, S.; Liu, F.; Li, Z. A facile method to prepare high molecular weight bio-renewable poly(γ-butyrolactone) using a strong base/urea binary synergistic catalytic system. Polym. Chem. 2019, 10, 1231–1237. [Google Scholar] [CrossRef]
  140. Jiang, Z.; Zhao, J.; Zhang, G. Ionic Organocatalyst with a Urea Anion and Tetra-n-butyl Ammonium Cation for Rapid, Selective, and Versatile Ring-Opening Polymerization of Lactide. ACS Macro Lett. 2019, 8, 759–765. [Google Scholar] [CrossRef]
  141. Kan, Z.; Luo, W.; Shi, T.; Wei, C.; Han, B.; Zheng, D.; Liu, S. Facile Preparation of Stereoblock PLA From Ring-Opening Polymerization of rac-Lactide by a Synergetic Binary Catalytic System Containing Ureas and Alkoxides. Front. Chem. 2018, 6, 547. [Google Scholar] [CrossRef] [PubMed]
  142. Lin, B.; Hedrick, J.L.; Park, N.H.; Waymouth, R.M. Programmable High-Throughput Platform for the Rapid and Scalable Synthesis of Polyester and Polycarbonate Libraries. J. Am. Chem. Soc. 2019, 141, 8921–8927. [Google Scholar] [CrossRef] [PubMed]
  143. Lin, L.; Han, D.; Qin, J.; Wang, S.; Xiao, M.; Sun, L.; Meng, Y. Nonstrained γ-Butyrolactone to High-Molecular-Weight Poly(γ-butyrolactone): Facile Bulk Polymerization Using Economical Ureas/Alkoxides. Macromolecules 2018, 51, 9317–9322. [Google Scholar] [CrossRef]
  144. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  145. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef]
  146. Lee, I.-H.; Martin, R.M. Applications of the generalized-gradient approximation to atoms, clusters, and solids. Phys. Rev. B 1997, 56, 7197–7205. [Google Scholar] [CrossRef]
  147. Shibasaki, Y.; Sanda, F.; Endo, T. Cationic Ring-Opening Polymerization of Seven-Membered Cyclic Carbonate with Water–Hydrogen Chloride through Activated Monomer Process. Macromolecules 2000, 33, 3590–3593. [Google Scholar] [CrossRef]
  148. Nakano, S. Polycarbonate-modified acrylic polymers for coating materials. Prog. Org. Coat. 1999, 35, 141–151. [Google Scholar] [CrossRef]
  149. Rokicki, G. Aliphatic cyclic carbonates and spiroorthocarbonates as monomers. Prog. Polym. Sci. 2000, 25, 259–342. [Google Scholar] [CrossRef]
  150. Sanda, F.; Fueki, T.; Endo, T. Cationic Ring-Opening Polymerization of an Exomethylene Group Carrying Cyclic Carbonate. Pseudo-Living Polymerization of 5-Methylene-1, 3-dioxan-2-one by the Assistance of the Exomethylene Group. Macromolecules 1999, 32, 4220–4224. [Google Scholar] [CrossRef]
  151. Kricheldorf, H.R.; Jonté, J.M.; Dunsing, R. Polylactones, 7. The mechanism of cationic polymerization of β-propiolactone and ϵ-caprolactone. Makromol. Chem. 1986, 187, 771–785. [Google Scholar] [CrossRef]
  152. Kricheldorf, H.R.; Dunsing, R. Polylactones, 8. Mechanism of the cationic polymerization of L,L-dilactide. Makromol. Chem. 1986, 187, 1611–1625. [Google Scholar] [CrossRef]
  153. Kricheldorf, H.R.; Kreiser, I. Polylactones, 11. Cationic copolymerization of glycolide with L,L-dilactide. Makromol. Chem. 1987, 188, 1861–1873. [Google Scholar] [CrossRef]
  154. Kricheldorf, H.R.; Dunsing, R.; Serra, A. Polylactones. 10. Cationic polymerization of δ-valerolactone by means of alkylating reagents. Macromolecules 1987, 20, 2050–2057. [Google Scholar] [CrossRef]
  155. Baśko, M.; Kubisa, P. Cationic copolymerization of ε-caprolactone and L,L-lactide by an activated monomer mechanism. J. Polym. Sci. Part A Polym. Chem. 2006, 44, 7071–7081. [Google Scholar] [CrossRef]
  156. Baśko, M.; Kubisa, P. Polyester oligodiols by cationic AM copolymerization of L,L-lactide and ε-caprolactone initiated by diols. J. Polym. Sci. Part A Polym. Chem. 2007, 45, 3090–3097. [Google Scholar] [CrossRef]
  157. Gazeau-Bureau, S.; Delcroix, D.; Martín-Vaca, B.; Bourissou, D.; Navarro, C.; Magnet, S. Organo-Catalyzed ROP of ϵ-Caprolactone: Methanesulfonic Acid Competes with Trifluoromethanesulfonic Acid. Macromolecules 2008, 41, 3782–3784. [Google Scholar] [CrossRef]
  158. Couffin, A.; Delcroix, D.; Martín-Vaca, B.; Bourissou, D.; Navarro, C. Mild and Efficient Preparation of Block and Gradient Copolymers by Methanesulfonic Acid Catalyzed Ring-Opening Polymerization of Caprolactone and Trimethylene Carbonate. Macromolecules 2013, 46, 4354–4360. [Google Scholar] [CrossRef]
  159. Wang, H.; Wu, W.; Li, Z.; Zhi, X.; Chen, C.; Zhao, C.; Li, X.; Zhang, Q.; Guo, K. 2,4-Dinitrobenzenesulfonic acid in an efficient Brønsted acid-catalyzed controlled/living ring-opening polymerization of ε-caprolactone. RSC Adv. 2014, 4, 55716–55722. [Google Scholar] [CrossRef]
  160. Sanda, F.; Sanada, H.; Shibasaki, Y.; Endo, T. Star Polymer Synthesis from ε-Caprolactone Utilizing Polyol/Protonic Acid Initiator. Macromolecules 2002, 35, 680–683. [Google Scholar] [CrossRef]
  161. Persson, P.V.; Schröder, J.; Wickholm, K.; Hedenström, E.; Iversen, T. Selective Organocatalytic Ring-Opening Polymerization: A Versatile Route to Carbohydrate-Functionalized Poly(ε-caprolactones). Macromolecules 2004, 37, 5889–5893. [Google Scholar] [CrossRef]
  162. Persson, P.V.; Casas, J.; Iversen, T.; Córdova, A. Direct Organocatalytic Chemoselective Synthesis of a Dendrimer-like Star Polyester. Macromolecules 2006, 39, 2819–2822. [Google Scholar] [CrossRef]
  163. Casas, J.; Persson, P.V.; Iversen, T.; Córdova, A. Direct Organocatalytic Ring-Opening Polymerizations of Lactones. Adv. Synth. Catal. 2004, 346, 1087–1089. [Google Scholar] [CrossRef]
  164. Makiguchi, K.; Satoh, T.; Kakuchi, T. Diphenyl Phosphate as an Efficient Cationic Organocatalyst for Controlled/Living Ring-Opening Polymerization of δ-Valerolactone and ε-Caprolactone. Macromolecules 2011, 44, 1999–2005. [Google Scholar] [CrossRef]
  165. Makiguchi, K.; Ogasawara, Y.; Kikuchi, S.; Satoh, T.; Kakuchi, T. Diphenyl Phosphate as an Efficient Acidic Organocatalyst for Controlled/Living Ring-Opening Polymerization of Trimethylene Carbonates Leading to Block, End-Functionalized, and Macrocyclic Polycarbonates. Macromolecules 2013, 46, 1772–1782. [Google Scholar] [CrossRef]
  166. Delcroix, D.; Couffin, A.; Susperregui, N.; Navarro, C.; Maron, L.; Martin-Vaca, B.; Bourissou, D. Phosphoric and phosphoramidic acids as bifunctional catalysts for the ring-opening polymerization of ε-caprolactone: A combined experimental and theoretical study. Polym. Chem. 2011, 2, 2249–2256. [Google Scholar] [CrossRef]
  167. Zhou, X.; Hong, L. Controlled ring-opening polymerization of cyclic esters with phosphoric acid as catalysts. Colloid Polym. Sci. 2013, 291, 2155–2162. [Google Scholar] [CrossRef]
  168. Shibasaki, Y.; Sanada, H.; Yokoi, M.; Sanda, F.; Endo, T. Activated Monomer Cationic Polymerization of Lactones and the Application to Well-Defined Block Copolymer Synthesis with Seven-Membered Cyclic Carbonate. Macromolecules 2000, 33, 4316–4320. [Google Scholar] [CrossRef]
  169. Shibasaki, Y.; Sanda, F.; Endo, T. Activated monomer cationic polymerization of 1, 3-dioxepan-2-one initiated by water-hydrogen chloride. Macromol. Rapid Commun. 1999, 20, 532–535. [Google Scholar] [CrossRef]
  170. Lou, X.; Detrembleur, C.; Jérôme, R. Living Cationic Polymerization of δ-Valerolactone and Synthesis of High Molecular Weight Homopolymer and Asymmetric Telechelic and Block Copolymer. Macromolecules 2002, 35, 1190–1195. [Google Scholar] [CrossRef]
  171. Kakuchi, R.; Tsuji, Y.; Chiba, K.; Fuchise, K.; Sakai, R.; Satoh, T.; Kakuchi, T. Controlled/Living Ring-Opening Polymerization of δ-Valerolactone Using Triflylimide as an Efficient Cationic Organocatalyst. Macromolecules 2010, 43, 7090–7094. [Google Scholar] [CrossRef]
  172. Delcroix, D.; Martín-Vaca, B.; Bourissou, D.; Navarro, C. Ring-Opening Polymerization of Trimethylene Carbonate Catalyzed by Methanesulfonic Acid: Activated Monomer versus Active Chain End Mechanisms. Macromolecules 2010, 43, 8828–8835. [Google Scholar] [CrossRef]
  173. Susperregui, N.; Delcroix, D.; Martin-Vaca, B.; Bourissou, D.; Maron, L. Ring-Opening Polymerization of ε-Caprolactone Catalyzed by Sulfonic Acids: Computational Evidence for Bifunctional Activation. J. Org. Chem. 2010, 75, 6581–6587. [Google Scholar] [CrossRef] [PubMed]
  174. Bergner, A.; Dolg, M.; Küchle, W.; Stoll, H.; Preuß, H. Ab initio energy-adjusted pseudopotentials for elements of groups 13–17. Mol. Phys. 1993, 80, 1431–1441. [Google Scholar] [CrossRef]
  175. Coady, D.J.; Horn, H.W.; Jones, G.O.; Sardon, H.; Engler, A.C.; Waymouth, R.M.; Rice, J.E.; Yang, Y.Y.; Hedrick, J.L. Polymerizing Base Sensitive Cyclic Carbonates Using Acid Catalysis. ACS Macro Lett. 2013, 2, 306–312. [Google Scholar] [CrossRef]
  176. GAMESS. Available online: https://www.msg.chem.iastate.edu/gamess/ (accessed on 21 November 2019).
  177. Peverati, R.; Truhlar, D.G. Improving the Accuracy of Hybrid Meta-GGA Density Functionals by Range Separation. J. Phys. Chem. Lett. 2011, 2, 2810–2817. [Google Scholar] [CrossRef]
  178. Xu, J.; Liu, J.; Li, Z.; Li, X.; Chen, C.; Zhao, C.; Xu, S.; Pan, X.; Liu, J.; Guo, K. Three is company: Dual intramolecular hydrogen-bond enabled carboxylic acid active in ring-opening polymerization. Polym. Chem. 2016, 7, 1111–1120. [Google Scholar] [CrossRef]
  179. Gupta, S.; Arora, R.; Sinha, N.; Alam, M.I.; Ali Haider, M. Mechanistic insights into the ring-opening of biomass derived lactones. RSC Adv. 2016, 6, 12932–12942. [Google Scholar] [CrossRef]
  180. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  181. Boileau, S.; Illy, N. Activation in anionic polymerization: Why phosphazene bases are very exciting promoters. Progr. Polym. Sci. 2011, 36, 1132–1151. [Google Scholar] [CrossRef]
  182. Zhang, L.; Nederberg, F.; Pratt, R.C.; Waymouth, R.M.; Hedrick, J.L.; Wade, C.G. Phosphazene Bases: A New Category of Organocatalysts for the Living Ring-Opening Polymerization of Cyclic Esters. Macromolecules 2007, 40, 4154–4158. [Google Scholar] [CrossRef]
  183. Liu, S.; Ren, C.; Zhao, N.; Shen, Y.; Li, Z. Phosphazene Bases as Organocatalysts for Ring-Opening Polymerization of Cyclic Esters. Macromol. Rapid. Commun. 2018, 39, 1800485. [Google Scholar] [CrossRef] [PubMed]
  184. Zhao, N.; Ren, C.; Li, H.; Li, Y.; Liu, S.; Li, Z. Selective Ring-Opening Polymerization of Non-Strained γ-Butyrolactone Catalyzed by A Cyclic Trimeric Phosphazene Base. Angew. Chem. Int. Ed. 2017, 56, 12987–12990. [Google Scholar] [CrossRef] [PubMed]
  185. Coumes, F.; Darcos, V.; Domurado, D.; Li, S.; Coudane, J. Synthesis and ring-opening polymerisation of a new alkyne-functionalised glycolide towards biocompatible amphiphilic graft copolymers. Polym. Chem. 2013, 4, 3705–3713. [Google Scholar] [CrossRef]
  186. Zhang, L.; Nederberg, F.; Messman, J.M.; Pratt, R.C.; Hedrick, J.L.; Wade, C.G. Organocatalytic Stereoselective Ring-Opening Polymerization of Lactide with Dimeric Phosphazene Bases. J. Am. Chem. Soc. 2007, 129, 12610–12611. [Google Scholar] [CrossRef]
  187. Ladelta, V.; Bilalis, P.; Gnanoub, Y.; Hadjichristidis, N. Ring-opening polymerization of ω-pentadecalactone catalyzed by phosphazene superbases. Polym. Chem. 2017, 8, 511–515. [Google Scholar] [CrossRef]
  188. Zhao, N.; Ren, C.; Shen, Y.; Liu, S.; Li, Z. Facile Synthesis of Aliphatic ω-Pentadecalactone Containing Diblock Copolyesters via Sequential ROP with l-Lactide, ε-Caprolactone, and δ-Valerolactone Catalyzed by Cyclic Trimeric Phosphazene Base with Inherent Tribasic Characteristics. Macromolecules 2019, 52, 1083–1091. [Google Scholar] [CrossRef]
  189. Lin, B.; Waymouth, R.M. Organic Ring-Opening Polymerization Catalysts: Reactivity Control by Balancing Acidity. Macromolecules 2018, 51, 2932–2938. [Google Scholar] [CrossRef]
  190. Pothupitiya, J.U.; Dharmaratne, N.U.; Jouaneh, T.M.M.; Fastnacht, K.V.; Coderre, D.N.; Kiesewetter, M.K. H-Bonding Organocatalysts for the Living, Solvent-Free Ring-Opening Polymerization of Lactones: Toward an All-Lactones, All-Conditions Approach. Macromolecules 2017, 50, 8948–8954. [Google Scholar] [CrossRef]
  191. Liras, M.; Verde-Sesto, E.; Iglesias, M.; Sánchez, F. Synthesis of polyesters by an efficient heterogeneous phosphazene (P1)-Porous Polymeric Aromatic Framework catalyzed-Ring Opening Polymerization of lactones. Eur. Polym. J. 2017, 95, 775–784. [Google Scholar] [CrossRef]
  192. Yuan, R.; Xu, G.; Lv, C.; Zhou, L.; Yang, R.; Wang, Q. Bifunctional phosphazene-thiourea/urea catalyzed ring-opening polymerization of cyclic esters. Mater. Today Commun. In Press. [CrossRef]
  193. Khademloo, E.; Saeidian, H.; Mirjafary, Z.; Aliabad, J.M. Design of Robust Organosuperbases and Anion Receptors by Combination of Azine Heterocycle Skeleton and Phosphazene Motif. Chem. Sel. 2019, 4, 3762–3767. [Google Scholar] [CrossRef]
  194. Shariatinia, Z.; Moghadam, E.J.; Maghsoudi, N.; Mousavi, H.S.M.; Dusek, M.; Eigner, V. Synthesis, Spectroscopy, X-ray Crystallography, and DFT Computations of Nanosized Phosphazenes. Z. Anorg. Allg. Chem. 2015, 641, 967–978. [Google Scholar] [CrossRef]
  195. Thomas, C.; Bibal, B. Hydrogen-bonding organocatalysts for ring-opening polymerization. Green Chem. 2014, 16, 1687–1699. [Google Scholar] [CrossRef]
  196. Horn, H.W.; Jones, G.O.; Wei, D.S.; Fukushima, K.; Lecuyer, J.M.; Coady, D.J.; Hedrick, J.L.; Rice, J.E. Mechanisms of Organocatalytic Amidation and Trans-Esterification of Aromatic Esters As a Model for the Depolymerization of Poly(ethylene) Terephthalate. J. Phys. Chem. A 2012, 116, 12389–12398. [Google Scholar] [CrossRef]
  197. Jones, G.O.; Yuen, A.; Wojtecki, R.J.; Hedrick, J.L.; García, J.M. Computational and experimental investigations of one-step conversion of poly(carbonate)s into value-added poly(aryl ether sulfone)s. Proc. Natl. Acad. Sci. USA 2016, 113, 7722–7726. [Google Scholar] [CrossRef]

Article Metrics

Citations

Article Access Statistics

Multiple requests from the same IP address are counted as one view.