Next Article in Journal
A New Azide-Bridged Polymeric Manganese (III) Schiff Base Complex with an Allylamine-Derived Ligand: Structural Characterization and Activity Spectra
Previous Article in Journal
Progress, Applications, and Challenges of Amorphous Alloys: A Critical Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

DFT Modeling of Coordination Polymerization of Polar Olefin Monomers by Molecular Metal Complexes

1
State Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences, Shanghai 200032, China
2
State Key Laboratory of Fine Chemicals, School of Chemical Engineering, Dalian University of Technology, Dalian 116024, China
3
National Elite Institute of Engineering, China National Petroleum Corporation (CNPC), Beijing 100096, China
4
PetroChina Petrochemical Research Institute, Beijing 102206, China
*
Authors to whom correspondence should be addressed.
Inorganics 2024, 12(9), 233; https://doi.org/10.3390/inorganics12090233
Submission received: 16 July 2024 / Revised: 22 August 2024 / Accepted: 26 August 2024 / Published: 28 August 2024

Abstract

:
Introducing polar functional groups into polyolefin chains through polar olefin monomer coordination (co)polymerization can directly and significantly improve the surface properties of polymer materials and expand their application range. Therefore, the related research has always received considerable attention from both academia and industry. Many experimental studies have been reported in this field, and molecular metal complexes have shown high catalytic activity and selectivity in polar olefin monomer polymerizations. Although considerable DFT calculations have also been conducted for better understanding of the (co)polymerization performance, the factors governing the activity, selectivity, and molecular weight of resulting polymers are still ambiguous. This review mainly focuses on the DFT studies of polar olefin monomer coordination (co)polymerization catalyzed by molecular metal complexes in recent years, discussing the chain initiation and propagation, the origin of polymerization activity and selectivity, and the specific role of additives in the (co)polymerization reactions.

1. Introduction

More than half a century ago, Karl Ziegler and Giulio Natta won the Nobel Prize for their great achievements in olefin coordination polymerization, marking the beginning of rapid development within the realm of polyolefin materials. So far, tremendous success in olefin polymerization has been achieved in both academic and industrial fields [1,2,3,4,5,6]. Polyolefin is currently the most widely used among polymer materials, and the global demand for polyolefin accounts for more than half of the total demand in the plastic market [7,8,9,10,11,12,13]. However, the non-polarity of conventional polyolefins limits their performance improvement and the expansion of their application range. In order to further improve their properties of printing, dyeing, adhesion, compatibility and blending, polyolefins can be functionalized by introducing highly reactive polar functional groups into the polymer chain (Scheme 1) [14,15].
There are four common methods for preparing functionalized polyolefins [16,17]: (a) post-functionalization method. This method is currently commercially used. However, due to the high inertness of the C-H bond in the polymer chain, the reaction conditions are relatively harsh (high temperature and pressure), and side reactions often occur. (b) Free radical polymerization. Through free radical polymerization, copolymerization of olefins with a small number of polar olefin monomers can be achieved, and the prepared functionalized polyolefins have been commercialized. However, the disadvantage of this method is the harsh polymerization reaction conditions (high temperature and high pressure), high product branching, poor microstructure controllability, and high content of polar olefin monomers. (c) Ring-opening metathesis polymerization. This method can provide functionalized polyolefins with precise and controllable chain structures, but the monomers required for this method are not easily obtainable and are relatively expensive. (d) Coordination (co)polymerization catalyzed by transition metals. Compared to the other three methods, the coordination (co)polymerization catalyzed by transition metal complexes has the advantages of mild conditions, controllability (including sequence distribution and structure, stereo and region regularity, geometry structure, and branching index), etc. [18,19,20]. Therefore, it is a more efficient and economical method.
At present, the system of polar monomer coordination polymerization catalyzed by transition metal complexes is under development [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39]. Experimentally, the polymerization activity and the microstructure of the product can be regulated by changing the catalyst ligand, metal center, monomer substituent, and additive. However, the extreme sensitivity of most active species to air and water and the inability to capture key transition states (TS) in the reaction hinder the in-depth study of a series of basic but important chemical issues regarding the factors governing regio- and stereo-selectivity as well as activity. With the gradual development of quantum chemistry and the continuous improvement of computer hardware, computational chemistry has become an effective way to study polymerization reaction mechanisms at the molecular level. Up to now, many research groups have widely applied computational chemistry to the study of the reaction mechanism of olefin polymerization catalyzed by transition metal complexes, which has promoted the design and development of catalysts [40,41,42,43,44,45]. Therefore, this review introduces the important theoretical progress of copolymerization of olefins with polar olefin monomers catalyzed by molecular metal complexes, in order to help experimental and theoretical chemists further comprehensively understand the mechanism of polar olefin monomer coordination copolymerization. According to different types of monomers, this review is divided into four parts, namely the DFT theoretical calculations related to the polymerizations of conjugated polar olefin monomer, non-conjugated polar olefin monomer, heteroatom functionalized styrene and vinylpyridine monomer (Scheme 2). In addition to early transition metal complexes and late-transition metal complexes, rare earth metal complexes (IIIB) are also included in the scope of this review.

2. Coordination Polymerization of Conjugated Vinyl Polar Monomer

Substantial progress has been made in the coordination polymerization of conjugated polar olefin monomers by early and late metal catalysts over the past two decades, particularly in terms of enhancing polymerization activity and stereochemical regulation. A large number of publications have been published on this important topic. The experimental and computational works on the polymerization of conjugated polar monomer have been reviewed in 2009 by Eugene Y.-X. Chen [41]. This article provides a detailed and comprehensive description of the important experimental and theoretical progress in the coordination polymerization of conjugated vinyl polar monomers by transition metal and rare-earth metal complexes. In addition, some reviews have covered the important experimental findings on the polymerization of conjugated vinyl polar monomer in the past fifteen years [40,41,42,43,44,45]. However, there is currently no comprehensive review on related theoretical calculations. Therefore, to add better understanding to this topic, it is necessary to comprehensively review the main progress of theoretical calculation in the field of coordination polymerization of conjugated polar olefin monomer by transition metal and rare-earth metal complexes (Scheme 3).
In 2010, Luigi Cavallo and Eugene Y.-X. Chen et al. conducted a combination of synthesis, kinetics, and theoretical calculations to explore several unique features of the polymerization of conjugated vinyl polar monomer catalyzed by Cs-ligated cationic ansa-metallocene ester enolate complexes (Figure 1) [46]. DFT calculations have revealed that installing the bulky Ph2C < bridge can decrease the Flu-C-Cp angle, effectively bringing the η3-bound Flu ligand in closer proximity to both the growing chain and the monomer. This could increase energy difference ∆Estereo between the competing TS for the addition of an MMA molecule to the opposite (correct and wrong) enantiofaces of the growing enolate chain; thus catalyst 6 (Ph2C < bridged) exhibits higher stereoselectivity than 2 (the Me2C < bridge) or 11 (the Me2Si < bridge) (Figure 2). Moreover, computational results also offer theoretical insights into the monomer-assisted catalyst-site epimerization, occurring after an enantiofacial error, leading to a transition to a thermodynamically more stable resting state (Figure 3). The dissociation energy of the C=O group in the last inserted unit within the three representative systems 2, 3, and 6 exhibits a decreasing trend that qualitatively aligns with the observed experimental activity. Moreover, the enhanced activity observed in these closely related catalysts can be reasonably attributed to the metallocene ligand’s capacity to hinder the tight binding of the bulky counterion.
In the same year, Luigi Cavallo and Stefan Mecking et al. also offered detailed insights into the pathways of multiple consecutive acrylate insertions through meticulous DFT calculations (Figure 4) [47]. The theoretical results indicate that the π-coordination of acrylate to the products resulting from multiple acrylate insertions is linked to a relatively modest energy gain. On a free energy basis, this process is likely endergonic, owing to the relatively favorable chelating coordination. Acrylate needs to overcome a 100 kJ mol−1 energy barrier to insert into 3-meso, 3-rac. Additionally, the authors also meticulously compared the relative energies (in kJ mol−1) of various isomers resulting from the coordination of methyl acrylate or ethylene to 3-meso.
In 2011, Bas de Bruin et al. explored the significance of β-H elimination as a possible mechanism to induce chain termination/transfer and/or the formation of stereodefects in the Rh(diene)-mediated oligomerization and polymerization of carbenes [48]. They found through theoretical calculations that the energy barrier of chain propagation is considerably higher than that of β-H elimination and complete dissociation of the elimination product at the beginning of the reaction, thereby elucidating the experimental formation of atactic oligomers at the beginning of the reaction. However, under the catalytic conditions, the Rh(diene) pre-catalyst undergoes modifications to generate new active species. This process appears to decelerate β-H elimination, resulting in nonstereospecific oligomerization activity, and ultimately, it may cease entirely, leading to stereoselective polymerization activity (Figure 5).
Laurent Maron and Peter W. Roesky et al. investigated the ability of rare earth metal complex [{CH(PPh2NSiMe3)2}La(BH4)2(THF)] and [{CH(PPh2NSiMe3)2}Ln(BH4)2] (Ln = Y, Lu) to polymerize methyl methacrylate (MMA) both experimentally and computationally [49]. DFT investigations into the insertion of the first two MMA molecules unveiled that the arrival of the initial MMA molecule was the pivotal step, irrespective of the metal center’s nature. The nucleophilic attack of MMA triggers the formation of the initial adduct B, succeeded by the remarkable trapping of the liberated BH3 group by the nitrogen of the phosphiniminomethanide ligand, yielding the active enolate species C. The distinctive and crucial function of the phosphiniminomethanide ligand has been distinctly uncovered and supported computationally (Figure 6). In addition, the chain propagation energy barrier of MMA polymerization catalyzed by yttrium catalyst was significantly higher (19.5 kcal mol−1) than that of lanthanum (steric considerations), indicating a lower reaction activity.
Luigi Cavallo and Eugene Y.-X. Chen et al. proposed using chiral C2-symmetric zirconocene catalysts to catalyze coordination polymerization of renewable α-methylene-γ-(methyl)butyrolactones [50]. Through the analysis of the key transition state, it can be concluded that the stereoselectivity of the product is mainly governed by steric interactions involving the monomer, the chain, and the catalyst ligand (Figure 7).
Luigi Cavallo and Stefan Mecking et al. explored the steric effects of the phosphinesulfonato ligand framework (Figure 8) on monomer (ethylene and MA) coordination and insertion through DFT methods [51]. According to the calculation results, monomers tend to coordinate at the cis position and then insert at the trans position in this polymerization system (Figure 9). In addition, the calculation results also show that the steric shielding leads to a pronounced increase in the molecular weight of resulting copolymer. They also discovered that the behavior of the catalysts cannot be comprehensively explained solely by considering the energetic profiles of the main steps of the insertion reaction. Additional factors such as deactivation and compound stability must also be taken into account; for instance, the deactivation of the compounds Ph1.
The stereoselectivity regulation mechanism of β-methyl-α-methylene-γ-butyrolactone (βMMBL) polymerization catalyzed by half-sandwich rare earth metal complexes has also been systematically explored by Luigi Cavallo et al. [52]. The formation of an isotactic polymer primarily results from interactions involving the methyl groups on the chiral β-C atom of the five-membered ring in both the monomer and the last inserted βMMBL unit of the chain. Contrary to the stereo-control exhibited by the current half-sandwich complexes, the auxiliary ligand on the metal minimally contributes to this process (Figure 10).
A new strategy proposed by Luigi Cavallo and Eugene Y.-X. Chen et al. can achieve complete chemoselective and highly syndiospecific coordination polymerization of divinyl polar monomers (AMA and VMA), facilitating the synthesis of highly syndiotactic polar vinyl polymers containing a pendant reactive C=C bond on every repeat unit [53]. At the same time, the stereoselectivity of these reactions was studied in detail by DFT. The detailed analysis of the energy and geometric structure of the key transition state shows that the main source of stereoselectivity is the steric interaction between monomer and metallocene skeleton (Figure 11).
The origin of different catalytic performance of diphosphazanemonoxide and phosphine-sulfonate palladium complexes toward copolymerization of ethylene and vinyl polar monomers (methyl acrylate MA or methyl methacrylate MMA) has been disclosed through DFT calculations by Luo et al. [54] It has been discovered that in the five-membered cationic system based on diphosphazanemonoxide, the coordination of monomers preferably occurs at the trans site (monomer trans to the P atom), while their insertions occur at the cis site (monomer cis to the P atom) (Figure 12). The geometric deformation caused by monomer insertion is the main reason for the occurrence of 2,1-insertion in the diphosphazane-monoxide palladium system. Furthermore, in the 2,1-insertion TS, the stronger MeO···H interaction between the ancillary ligand and the Me group of the inserting MMA moiety makes the TS more stable. The rate-determining step is the ethylene insertion into the MA pre-enchained species in the copolymerization reaction, rather than the MA insertion itself. Moreover, β-H elimination reaction is more likely to occur to result in the MMA chain-end, owing to the steric repulsion between the methyl group of preinserted MMA and the subsequent inserting ethylene moiety.
Bimal Pudasaini has provided an in-depth computational mechanistic analysis of the Y-catalyzed precision polymerization of vinyl phosphonates [55]. First of all, two different initiation mechanisms of dimethylvinyl-phosphonate (DMVP) catalyzed by rare earth metal alkyl complexes (with –CH3 ligand) have been compared, and the calculation results show that the methyl addition mechanism is more favorable than the deprotonation mechanism (Figure 13). Then, a series of various ligands were compared by DFT calculation, and the results showed that most of the other ligands studied did not show preference for the deprotonation mechanism. It is worth noting that catalysts featuring a dimethylamido (–NMe2) ligand exhibit the lowest initiation barriers and may serve as a promising experimental option. In addition, when the catalyst with ligand –CH2TMS is calculated, barrier tunneling needs to be considered, because transmission coefficients (κ) of –CH2TMS increase very quickly as the temperature decreases.
In 2020, the polymerization mechanism of methyl methacrylate (MMA) catalyzed by rare-earth/phosphorus (RE/P) Lewis pairs has been comprehensively examined through DFT calculations by Luo et al. [56]. They found that the polymerization of MMA mediated by intermolecular RE/P (RE = Sc and La) Lewis pairs mainly follow the bimetallic mechanism (Figure 14), while the possibility of the monometallic pathway cannot be disregarded in the case of La analogue. By calculating and comparing several different Lewis bases, it can be found that the enhanced activity of cyclohexyl phosphorus in MMA polymerization may be attributed to its electron-donating capabilities, facilitating increased electron transfer in the addition reaction.
In order to obtain an ideal catalyst with high selectivity, high activity, and high polar monomer incorporation rate, Barbara Milani and colleagues altered the conventional Brookhart’s catalyst by incorporating a hemilabile, potentially bidentate ligand, such as a thiophenimine (N–S), into the fourth coordination site of palladium [57]. In addition to verifying the activity and selectivity of the new catalyst through various experimental methods, the author also used DFT theoretical calculations to explore the ethylene/MA copolymerization mechanism of the new catalyst 1SPh (Figure 15). The calculation results indicate that the N–S ligand maintains proximity to the palladium ion, promoting the formation of an open-chain intermediate over the anticipated six-membered metallacycle. This open-chain intermediate plays a crucial role in capturing the polar monomer into the main chain.
In 2022, a new type of α-sulfonato-β-diimine nickel catalyst has been proposed by Gao Haiyang et al., demonstrating notable thermal stability and control, along with exceptional tolerance towards polar alkyl acrylates [58]. To better understand the mechanism of α-sulfonato-β-diimine nickel-catalyzed copolymerization of ethylene and MA, DFT study was also carried out (Figure 16). The calculation results reveal that the 2,1-insertion of MA is favored, which effectively accounts for the formation of EMA copolymer with a terminal MA unit.

3. Coordination Polymerization of Non-Conjugated Polar Olefin Monomer

Similar to the conjugated olefin polar monomers mentioned above, non-conjugated olefin polar monomers can also directly introduce polar functional groups into the polymer chain, improving the performance of polymers and making them more widely applicable. At present, many transition-metal complexes have been calculated in detail to catalyze the polymerization of non-conjugated polar monomers [59,60,61,62,63,64,65,66]. It is worth noting that Eugene Y.-X. Chen and Hou et al. have successively summarized the experimental and computational work on coordination polymerization of polar olefin monomers by single site metal catalysts [41] and half-sandwich rare-earth catalysts before 2015 [60]. Based on the published work, this minireview mainly provides a comprehensive summary of the theoretical calculations of transition metal and rare-earth metal complexes catalyzed non-conjugated polar monomer polymerization systems that have been published since 2015 (Scheme 4).
In 2016, Lucia Caporaso and Stefan Mecking et al. proposed to use PdII-catalyzed copolymerization of ethylene and 2-vinylfuran (VF) generates α,ω-di-furan telechelic polyethylene [67]. Meanwhile, systematic theoretical calculations on the copolymerization process of VF and ethylene have been performed. According to mechanism calculations, following a VF insertion into the chain, a subsequent chain transfer reaction predominantly involves the VF monomer as well. This process leads to the termination of the chain with an unsaturated VF-based unit, observed experimentally, while concurrently initiating a new chain with VF (Figure 17).
The key findings on the heteroatom-assisted olefin polymerization (HOP) strategy of heteroatom (O, S, Se, N, and P)-functionalized polyolefins polymerization, catalyzed by half-sandwich rare-earth catalysts, were made by Hou and Luo [68]. The computational investigations in this study have unveiled that a suitable interaction between the heteroatom in an α-olefin and the catalyst metal center is paramount in attaining olefin polymerization activity and stereoselectivity (Figure 18). The mechanism of heteroatom assisted olefin insertion has been proposed for the first time and seems to be applicable to other polymerization systems, if the suitable interaction between metal and heteroatoms is achievable.
In 2019, density functional theory (DFT) methods were used to investigate the mechanism of organoscandium-catalyzed ethylene and amino olefin (AO, N-(1-butenyl)nPr2 and N-(1-octenyl)nPr2) copolymerization by T. Marks et al. [69]. The computational results demonstrate that the short-chain N-(1-butenyl)nPr2 undergoes enchainment through a self-assisted insertion pathway (energy barrier of 6.0 kcal/mol), whereas the long-chain N-(1-octenyl)nPr2 enchains through unassisted 1,2-insertion with exogenous s amine coordination (energy barrier of 7.2 kcal/mol) (Figure 19).
The nickel and palladium complexes bearing new diarylamido-based unsymmetrical [NNNox] and [SNNox] pincer ligands were synthesized and tested for polymerization of norbornene by Luo and Shi [70]. The sidearm effects, which involve quinolino ([NNNox] ligands) and phenylthio ([SNNox] ligands) moieties, have also demonstrated a noticeable impact on the catalytic performance in the polymerization of norbornene. The study revealed that the increased steric hindrance resulting from the coordinating P-sidearm in the [PNNox]-ligated complex could explain its lower catalytic activity compared to the [NNNox] ligated complex (Figure 20).
The electron-rich nature of vinyl ethers (CH2=CHOR) are highly attractive monomers for the synthesis of many polymers and copolymers. Chen et al. developed a system that integrates the palladium-catalyzed dimerization of vinyl ethers to produce β, γ-unsaturated acetals and palladium-catalyzed ethylene copolymerization with the dimerization product to generate polar functionalized branched polyolefins [71]. The mechanism and catalyst optimization for the dimerization reaction were also thoroughly investigated (Figure 21). The calculated results indicate that the energy barrier for the third insertion is 4.7 kcal/mol higher than that of the second insertion. Moreover, the energy barrier for the β-OMe elimination reaction is significantly lower than that of the third vinyl ether insertion (17.8 vs. 25.3 kcal/mol). This finding provides a comprehensive explanation for the experimentally observed high selectivity in the dimerization process.
In 2020, the unconventional highly 1,2-selective polymerization of polar and nonpolar alkylallene compounds has been achieved by Laurent Maron and Dongmei Cui et al. [72]. They conducted detailed mechanism research on this system using DFT theory calculations, and they discovered that the reaction sequence is governed by the readily occurring 1,2-insertion step rather than the coordination of the monomer to the metal center (Figure 22). There is a certain steric hindrance between the growing polymer chain and the allene, which is the main reason for preventing the formation of 1,2 adducts.
Ethylene and vinyl halide (VX, X = F or Cl) copolymerization mechanism in the presence of several neutral palladium phosphine sulfonate catalysts A ((POOMe,OMe)PdMe, POOMe,OMe = {2(2-MeOC6H4)(2-SO3-5-MeC6H3)P}) and A’ ((POBp,OMe)PdMe, POBp,OMe = {(2-MeOC6H4)(2-{2,6-(MeO)2C6H3}C6H4)(2-SO3-5-MeC6H3)P}) have been comparatively studied via density functional theory (DFT) calculations by Yi Luo et al.; the origin of different chain-end microstructures produced by A and A’ and the discrepancy due to VX insertion into Pd–X (X = F and Cl) species are disclosed [73]. For catalyst A’, more positive charge of metal Pd assisting the β-F elimination and stronger H···OMe and C–H···π interactions are the main reasons for obtaining the chain-end of –CH2–CHF2. On the contrary, the strong Pd-Cl bond and the reversible β-Cl elimination are the main reasons why the chain-end unit –CH2–CHCl2 cannot be obtained (Figure 23). Moreover, the mechanism of ethylene with vinyl ether (VE, CH2=CHOEt) copolymerization catalyzed by phosphine-sulfonate palladium complex was also investigated by Yi Luo et al. [74]; the calculation results show that due to the steric hindrance between the coming VE and growing chain, the energy barrier for repeated insertion of VE is too high (29.1 kcal/mol); as a result, the copolymer exclusively incorporates OEt in both the polymer chain and chain end, without the repetitive insertion units of VE (Figure 24).
A scandium dication active species [(IPr)Sc(μ-CH2SiMe3)(μ-CH2CHMe2)AliBu2]2+[B(C6F5)4]2− (IPr = (2,6-C6H3iPr2NCH)2C) in situ was synthesized by Li Xiaofang et al. and exhibited unprecedentedly high activity and syndiotactic selectivity in the polymerization of o-methoxystyrene (oMOS) and its silyloxy- or fluorine-substituted derivatives [75]. The polymerization mechanism of oMOS by this Sc dication active species has also been theoretically studied (Figure 25). Computational DFT studies suggest that the increased oxyphilic nature of scandium, which features two vacant coordination sites, enables synergistic coordination of both the heteroatom group and the double bond of oMOS to the scandium center in a σ−π coordination mode. This σ–π-coordination mode can effectively reduce the energy barrier of C=C insertion, and the primary factor contributing to the formation of syndiotactic polymerization products is the coordination of the methoxy groups from the last two inserted oMOS units in the polymer chain with the metal center simultaneously.
Usually, copolymerizing ethylene with industrially significant short-chain alkenoic acids remains a significant challenge. Luo and Tang et al. reported the efficient direct copolymerization of ethylene with vinyl acetic acid by tetranuclear nickel complexes, and the protic monomer can be extended to acrylic acid, allylacetic acid, ω-alkenoic acid, allyl alcohol, and homoallyl alcohol [76]. In addition, they also explored in detail the mechanism of ethylene and polar monomer copolymerization in this catalytic system. According to the calculation results, a proposed rationale involves the enchainment of VA facilitated by unique Ni···Ni synergistic effects (Figure 26). Motivated by the mechanistic insights, a binuclear pre-catalyst with a shorter Ni···Ni distance is synthesized and proves to be significantly more efficient for the copolymerization of ethylene with VA and AA compared to tetranuclear complexes.
In 2022, Luo et al. investigated the detrimental impact of various polar monomers on Brookhart-type catalysts using a combination of DFT calculations and multiple linear regression analyses [77]. They established a connection between the structure and the poisoning effect (ΔΔE(π-σ)) by integrating the heteroatom coordination complex structure with the descriptors of the polar monomer (Figure 27). The results indicate that the descriptors of monNMRCβ and B2BondNi/Pd-X are the main factors influencing the poisoning effect. It is important to highlight that the non-covalent interaction between polar monomers and the catalyst could be the primary reason for the deviation of the predicted value by the regression model.
Following this, Luo et al. presented an in-depth computational study on the reaction mechanism and reactivity involved in the polymerization of polar monomers catalyzed by early transition metals, where methylene spacers are present between the vinyl and functional groups [78]. They demonstrated that early transition metal catalysts with strong oxygen affinity, such as titanium(III) and zirconium(III) metallocenes, may not be susceptible to poisoning during the polymerization of certain polar monomers, allowing for successful heteroatom-chelating olefin (Figure 28). This finding offers a novel pathway for catalytic polymerization of polar monomers using early transition metal complexes.
To further address the issues related to the copolymerization system of polar and non-polar monomers, Cui et al. proposed the copolymerization of ethylene with 6-phenoxy-1-hexene (POH) by scandium precatalysts with different steric hindrances [79]. The DFT calculations show that both the polar group and the olefin moiety coordinate to the scandium ion simultaneously, serving as a crucial factor in enhancing the copolymerization process and boosting the incorporation of polar olefins through a self-assisted polymerization mechanism (Figure 29). This strategy can effectively utilize the coordination of heteroatom substituents on polar monomers to improve the polymerization activity and insertion rate of polar monomers, providing an effective approach for the polymerization of other similar polar monomers. Additionally, Luo et al. utilized DFT calculations to investigate the mechanistic aspects of the Sc-catalyzed co-syndiospecific alternating copolymerization of anisylpropylene (AP) and styrene in a model reaction [80]. Their findings highlight the critical role of the delicate balance between electronic and steric factors in monomer insertions, and they introduced a novel amino dissociated mechanism for AP insertion during chain initiation. Meanwhile, they used the buried volume of the metal center of the preinserted species to support that the alternating monomer-sequences were mainly determined by the “steric matching” principle (Figure 30).

4. Coordination Polymerization of 2-Vinylpyridine

Poly(2-vinylpyridine) (P2VP) is widely utilized in various applications such as thin films, molecular detection, photochemistry, ion exchange, and self-assembly [81,82]. Since 1960 [83], various experimental systems for 2VP polymerization have been continuously proposed, for example, radical polymerization, anionic polymerization, Lewis’s acid–base pair polymerization and coordination polymerization. Among them, coordination polymerization initiated by rare-earth metal catalysts is the most efficient method, and this strategy has also developed rapidly in recent years [84]. With the continuous enrichment of the experimental system, theoretical calculations on the polymerization mechanism of 2VP are constantly being conducted (Scheme 5).
In 2016, Xu et al. proposed novel yttrium bis(phenolate) ether catalysts to mediate 2VP polymerization with highly isotactic (Pm = 0.97) and high molecular weight [85]. Simultaneously, they conducted detailed calculations on the mechanism of 2VP polymerization (Figure 31). The calculation results indicated that coordination insertion of 2-VP on the re-face was more favored than that on si-face in the chain initiation stage, and the chain initiation stage was the rate-determining step in the polymerization process. During the chain propagation stage, 2VP are mainly inserted on the si-face, thus the isotactic polymerization products would be obtained. This work provides a detailed demonstration of the polymerization mechanism of 2VP, providing important ideas for further regulating the polymerization system of 2VP.
In 2017, the chain initiations of β-butyrolactone (BBL) and 2-vinylpyridine (2VP) polymerizations mediated by yttrium complexes were computationally investigated, respectively, by Luo et al. [86]. Through theoretical calculations, they found that the orbital matching degree between the initial alkyl group of a catalytically active species and vinyl group of 2VP is the main factor affecting the chain initiation efficiency (Figure 32). When the orbital matching is good, the initiation efficiency of active species is high, and vice versa. These theoretical findings offer enhanced insights into the distinct activities involved in the polymerizations of BBL and 2VP.
Moreover, the catalytic activity of 2-aminoalkoxybis (phenolate)yttrium and lutetium complexes for polar monomer polymerization was experimentally and theoretically compared by Rieger et al. [87]. Through mechanism calculations and key structural comparisons, they proposed that 2VP is a strong electron donor; when a catalyst with electron-donating ligands was used, an overload of electron density to the metal center is provoked. The presence of an electron-donating nitrogen atom and reduced electron density on the activated CH2 group of collidine, caused by charge delocalization, did not lead to effective nucleophilic transfer to 2VP. Thus, initiation of catalyst with collidine was lower than that with alkyl (Figure 33). In comparison, catalyst 1b demonstrates versatility as a suitable option for DMAA polymerizations, mainly because DMAA can be added to catalyst 1b to form a stable eight-membered ring intermediate.
In 2018, excellent isoselectivity (mmmm > 99%) and high activity (TOF > 4000 h−1) were achieved for the first time in the polymerization of the polar 2-vinylpyridine using simple lutetium-based catalysts by Xu et al. [88]. Significantly, they identified a novel approach to produce isotactic-atactic stereomultiblock poly(2-vinylpyridine)s through the addition and removal of Lewis bases during polymerization. DFT calculations were utilized to deduce the stereoselective control mechanism during the polymerization process. The calculation results indicate that the introduction of a Lewis base reduces the energy gap between the transition states of re-face and si-face, resulting in a decrease in the tacticity of the polymer (Figure 34).

5. Stereoselective Polymerization of 4-Vinylpyridine

Similar to 2-vinylpyridine, the metal-catalyzed polymerization of 4-vinylpyridine (4VP) presents a substantial challenge due to the restricted geometry of the initiating group and metal center during catalysis [89]. So far, there have been few works on the calculation of the polymerization mechanism of 4-ethylpyridine. In 2021, Xu and colleagues documented the successful polymerization of 4-vinylpyridine (4VP) through the utilization of Lewis pairs consisting of homoleptic rare-earth (RE) aryloxides and ylide-functionalized phosphines (YPhos). Notably, the computational investigation also yielded essential insights into the comparison between YPhos and PEt3 as the Lewis base component in the polymerization process. DFT calculations conducted by Luo and co-workers revealed that the reduced stability of TS1PE compared to TS1P2, attributed to increased steric repulsion and weaker interactions between 4VP and the PEt3 group, led to lower polymerization activity than that achieved with YPhos (Figure 35) [90].

6. Stereoselective Polymerization of Heteroatom-Functionalized Styrene

Polystyrenes with enhanced surface properties due to functionalization are in high demand for various industrial applications [91]. Compared to post-modification processes, the direct (co)polymerization of heteroatom-functionalized styrene using transition metal complexes is considered one of the most convenient and controllable methods to generate functionalized polystyrenes (Scheme 6) [63,84].
In 2015, the stereoselective coordination/insertion polymerization of the polar ortho-methoxystyrene (oMOS) was successfully accomplished for the first time using the cationic β-diketiminato rare earth-metal species by Cui et al. [92]. Based on the experimental results, they calculated the polymerization mechanism through DFT theory and found that the coordination between ortho-methoxy group and metal center can reduce the coordination energy and insertion energy barrier of the C=C bond of o-MOS, promoting the polymerization reaction (Figure 36). This approach offers a novel pathway for synthesizing homopolymers and copolymers with polar groups by utilizing carefully selected catalysts and polar monomers.
Subsequently, Cui et al. systematically investigated copolymerization of unmasked polar methoxystyrene monomers with nonpolar styrene by using a cationic CGC rare-earth yttrium catalyst [93]. Comparative calculations reveal that the energy barrier for styrene insertion into the active species with a polar styrene unit attached is lower compared to catalysts without such a group. In addition, the electron-donating nitrogen atom from pyridine in the CGC catalyst weakens the interaction between polar styrene and the active species. Meanwhile, due to the different chelation modes between methoxy groups and metal centers, the position of methoxy groups on MOS can affect the sequence distribution of copolymers (Figure 37). In the same year, Cui’s group also illustrated the novel coordination-insertion polymerization of a polar electron-withdrawing fluorine-substituted styrene and its copolymerization with styrene (St) employing rare-earth-metal-based catalysts [94]. The calculation results indicate that the coordination between the electron-donating pyridine side-arm and the metal center could weaken the Lewis acidity of the metal center, thereby reducing the interaction energy between the F atom and the metal center (Figure 38b). Therefore, the coordination energy of the F atom is similar to that of the C=C bond. This can increase the efficiency of the polymerization reaction. By contrast, when the catalyst bearing the unlinked half-sandwich ligand has been used, the metal center tends to coordinate with the F atoms of two monomers at the same time, forming more stable intermediates, resulting in relatively difficult coordination and insertion of C=C bonds (Figure 38a). The above calculation results provide an important theoretical basis for improving the reactivity of the homopolymerization and copolymerization reaction systems of methoxy styrene.
In 2018, Luo and co-authors conducted a comparative analysis of the polymerization mechanism of styrenes substituted with polar functional groups catalyzed by cationic rare-earth metal alkyl complexes using DFT calculations [95]. The underlying reasons for different polymerization mechanisms, namely step growth involving C-H activation and chain growth through vinyl insertion, have been elucidated. They observed that reduced steric hindrance and higher average charge distribution in the four-center transition state are advantageous for C-H activation associated with step-growth polymerization. Additionally, the coordination of the heteroatom from the functional group may promote the insertion of the adjacent vinyl group, leading to a somewhat negative impact on C-H activation (Figure 39). Moreover, due to the steric hindrance and electron donor effects caused by the catalyst side arms, the C-H activation is inhibited. This work mainly conducted a systematic study on the conditions for the occurrence of step-growth polymerization, providing important information for the subsequent development of more efficient step-growth polymerization systems.
In 2019, Cui et al. studied the coordination polymerization of the unmasked polar alkoxystyrenes (AOS) by using the cationic quinolyl anilido yttrium alkyl species that was inert to styrene polymerization, and the mechanism was elucidated by the DFT calculations (Figure 40) [96]. These computational investigations have highlighted the importance of a robust interaction between the polar o-alkoxy group of AOS and the active yttrium center in enhancing the activity and stereoselectivity of AOS polymerization. Importantly, modifying the ortho-oxygen substituent from methyl to benzyl can alter the stereoregularity of the resulting polymer from highly syndiotactic to atactic by weakening the Y-σ-O coordination bond.
By using DFT calculations, Luo et al. demonstrated that the stereoselectivity of Sc-catalyzed polymerization of halogenated styrenes can be co-regulated by C–H···π interactions between a coordinated THF molecule and the phenyl ring of a monomer [97]. The theoretical calculation results suggest that the selectivity of halogenated styrene polymerization products can be adjusted from syndiotactic to random by changing the number and sites of halogen element substitutions on halogenated styrene, which is achieved by changing the interaction between monomers and THF (Figure 41). The strategy of regulating stereoselectivity through a noncovalent interaction can provide a new approach for selective polymerization of olefins.
Afterwards, Cui et al. proposed copolymerizations of ethylene (E) and ortho-/meta-/para-fluorostyrenes by using quinolyl methylene fluorenyl scandium complex (Flu-CH2-Qu)Sc(CH2SiMe3)2 [98]. According to theoretical calculations, the Lewis acidity of the metal center and the local steric hindrance are influenced by the bulky and electron-withdrawing quinolyl side arm, resulting in a tendency towards continuous insertion of ethylene during copolymerization (Figure 42). Furthermore, it was also found that the insertion of fluorostyrenes is a thermodynamic product.
The chemoselective polymerization of divinyl monomers is a well-recognized approach for producing functional polymers through post-polymerization; however, this method may lead to issues such as polymer cross-linking. Cui and Zhao et al. reported the coordination polymerization of polar divinyl styrenyl monomers 1-(allyloxy)-4-vinylbenzene (AOS), 1-(but-3-en-1-yloxy)-4-vinylbenzene (BOS), 1-(pent-4-en-1-yloxy)-4-vinylbenzene (POS), and 1-(hex-5-en-1-yloxy)-4-vinylbenzene (HOS) by employing rare-earth metal complexes [99]. At the same time, the factors influencing the (co)polymerization behaviors were explored through DFT calculations. The spacer length between the olefinic C=C bond and the oxygen atom of divinyl monomers is the main factor affecting the microstructure of polymerization products. For monomers containing a short spacer, there exists the concurrent Y-μ1-O: η2-C=C coordination, resulting in polymer chain back-skipping and equal energy barriers for re-insertion and si-insertion, leading to the formation of atactic polymers (Figure 43). In the case of a monomer with a long spacer, the hindrance of polymer chain back-skipping is increased, promoting the preferred alternating re- and si-insertion of the styrenyl C=C bond to yield syndiotactic polymers.
Recently, based on the “self-assisted” theory previously proposed by the Cui’s group, they also presented new racemic ansa-bis(benz[e]indenyl) rare-earth metal complex for isospecific polymerization of polar styrenes, and the isospecific mechanism has been revealed through DFT [100]. According to mechanism calculations, the bulky C2-symmetric [BIndCMe2BInd]2–ligand facilitates pMOS si-face coordination, leading to ideal isoselectivity through the enantiomorphic site-control mechanism (Figure 44).
Besides, Luo’s research group systematically investigated the factors contributing to the atactic and syndiotactic selectivity in the polymerization of amino-functionalized styrenes catalyzed by rare-earth-metal catalysts [101]. The computational findings revealed that the stronger the electronic influence of the aromatic rings (-NMe2 > -OH > -F), the stronger the C-H...π interaction between the monomer and complex. Consequently, the greater the energy barrier difference ΔΔG, the higher the syndioselectivity (Figure 45). Furthermore, when the steric hindrance of functionalized styrene and ligand substituents are both small, energy barriers of re-face and si-face insertion are similar, resulting in random polymerization products. In 2023, Luo et al. also investigated the effect of additives on the stereoselectivity of styrene polymerization catalyzed by rare-earth metal complexes [102]; their findings indicated that the primary factor influencing stereoselectivity was the noncovalent interaction between the additive alloxybenzene and the polymer chain/monomer (Figure 46).
In addition, the ring opening polymerization (ROP) of polar monomers such as cyclic esters, which is out of the scope of this review, has also received widespread attention in the polymer community [103,104,105,106,107,108,109]. Significantly, transition metal complex-catalyzed ROP has always been at the forefront of the development of ROP catalysts for epoxides, cyclic esters, and carbonates, and is even more commonly used than organic catalysis [110,111,112,113]. In experiments, a considerable number of researchers have studied cyclic esters such as ε-Caprolactone (CL), lactide (LA), and other related monomers. There are a certain number of review articles [114,115,116,117,118,119,120] and textbooks [121,122] that have covered the development of ROP of cyclic esters catalyzed by transition metal complexes, documenting the rapid development of this field. Meanwhile, the ROP mechanisms have been comprehensively explored by DFT calculations. In 2019, the DFT calculation results for different ROP mechanisms of various cyclic substrates were systematically integrated by Ilya Nifant’ev and Pavel Ivchenko [117]. In a recent review, Brian K. Long and colleagues discussed the early developments, key recent advancements, and future prospects in the area of redox-switchable ring-opening polymerizations of cyclic esters and cyclic ethers [120]. These two reviews combine theoretical and experimental works to provide a detailed and systematic understanding in this field.

7. Conclusions and Perspectives

In recent years, an increasing number of experimental and theoretical researchers are interested in polar monomer coordination polymerization catalyzed by molecular metal complexes. This is mainly because, unlike radical copolymerization, it can achieve precise control of the microstructure of the polar polymer under mild conditions. At the same time, with the rapid development of computational methods and computer hardware, DFT modeling has gradually become a commonly useful tool in the studies of catalysis, including catalytic polymerization. Theoretical calculations can model various species that may form during the catalytic process, as well as possible non-covalent interactions, which often play an important role in the activity and microstructure regulations.
The technology of polar monomer coordination copolymerization is currently not industrially available yet. Therefore, catalyst development certainly requires fundamental research in this field although it is challenging in increasing activity, molecular weight of resulting copolymer, and the content of polar monomer in the polymer chain. In this sense, there should be considerable opportunities for theoretical calculation studies to design more efficient polymerization catalysts. It is noteworthy that the combination of DFT calculations and data analysis (such as machine learning, multiple linear regression, and principal component analysis) is a more promising approach for rational design of catalysts. It is expected that the combination of experiment and theory speeds up the development of copolymerization systems of polar monomers and nonpolar olefins.

Funding

This work was supported by the NSFC (No. 22071015), Science and Technology Commission of Shanghai Municipality (No. 23JC1404500), and supported by the Strategic Priority Research Program of the Chinese Academy of Sciences, Grant No. XDB0610000.

Conflicts of Interest

Zhenli Zhang was employed by the National Elite Institute of Engineering, China National Petroleum Corporation (CNPC). The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Reetz, M.T. One Hundred Years of the Max-Planck-Institut für Kohlenforschung. Angew. Chem. Int. Ed. 2014, 53, 8562–8586. [Google Scholar] [CrossRef] [PubMed]
  2. Stürzel, M.; Mihan, S.; Mülhaupt, R. From Multisite Polymerization Catalysis to Sustainable Materials and All-Polyolefin Composites. Chem. Rev. 2016, 116, 1398–1433. [Google Scholar] [CrossRef]
  3. Kaminsky, W. (Ed.) Polyolefins 50 Years after Ziegler and Natta I Polyethylene and Polypropylene; Springer: New York, NY, USA, 2013; Volume 257. [Google Scholar]
  4. Chadwick, J.C. Polyolefins—Catalyst and Process Innovations and their Impact on Polymer Properties. Macromol. React. Eng. 2009, 3, 428–432. [Google Scholar] [CrossRef]
  5. Busico, V. Ziegler-Natta catalysis: Forever young. MRS Bull. 2013, 38, 224–228. [Google Scholar] [CrossRef]
  6. Ziegler, K.; Natta, G. Paving the way to polythene. Chem. World 2013, 10, 50–53. [Google Scholar]
  7. McDaniel, M.P. A Review of the Phillips Supported Chromium Catalyst and Its Commercial Use for Ethylene Polymerization. Angew. Chem. Int. Ed. 2003, 42, 5010–5030. [Google Scholar]
  8. Baier, M.C.; Zuideveld, M.A.; Mecking, S. Post-Metallocenes in the Industrial Production of Polyolefins. Angew. Chem. Int. Ed. 2014, 53, 9722–9744. [Google Scholar] [CrossRef]
  9. Mülhaupt, R. Green Polymer Chemistry and Bio-based Plastics: Dreams and Reality. Macromol. Chem. Phys. 2013, 214, 159–174. [Google Scholar] [CrossRef]
  10. Dubé, M.; Salehpour, S. Applying the Principles of Green Chemistry to Polymer Production Technology. Macromol. React. Eng. 2014, 8, 7–28. [Google Scholar] [CrossRef]
  11. Gahleitner, M.; Resconi, L.; Doshev, P. Heterogeneous Ziegler-Natta, Metallocene, and Post-Metallocene Catalysis: Successes and Challenges in Industrial Application. MRS Bull. 2013, 38, 229–233. [Google Scholar]
  12. Qiao, J.; Guo, M.; Wang, L.; Liu, D.; Zhang, X.; Yu, L.; Song, W.; Liu, Y. Recent Advances in Polyolefin Technology. Polym. Chem. 2011, 2, 1611–1623. [Google Scholar] [CrossRef]
  13. Chum, P.; Swogger, K.W. Olefin Polymer Technologies—History and Recent Progress at the Dow Chemical Company. Prog. Polym. Sci. 2008, 33, 797–819. [Google Scholar] [CrossRef]
  14. Jian, Z. Synthesis of Functionalized Polyolefins: Design from Catalysts to Polar Monomers. Acta Polym. Sin. 2018, 11, 1359–1371. [Google Scholar]
  15. Dong, J.; Hu, Y. Design and Synthesis of Structurally Well-Defined Functional Polyolefins via Transition Metal-Mediated Olefin Polymerization Chemistry. Coord Chem. Rev. 2006, 250, 47–65. [Google Scholar] [CrossRef]
  16. Boaen, N.K.; Hillmyer, M.A. Post-Polymerization Functionalization of Polyolefins. Chem. Soc. Rev. 2005, 34, 267–275. [Google Scholar] [CrossRef] [PubMed]
  17. Franssen, N.M.G.; Reeka, J.N.H.; Bruin, B. Synthesis of Functional ‘Polyolefins’: State of the Art and Remaining Challenges. Chem. Soc. Rev. 2013, 42, 5809–5832. [Google Scholar] [CrossRef] [PubMed]
  18. Chen, C. Designing Catalysts for olefin Polymerization and Copolymerization: Beyond Electronic and Steric Tuning. Nat. Rev. Chem. 2018, 2, 6–14. [Google Scholar] [CrossRef]
  19. Gladysz, J.A.; Ball, Z.T.; Bertrand, G.; Blum, S.A.; Dong, V.M.; Dorta, R.; Hahn, F.E.; Humphrey, M.G.; Jones, W.D.; Klosin, J.; et al. Organometallics Roundtable 2011. Organometallics 2012, 31, 1–18. [Google Scholar] [CrossRef]
  20. Chen, M.; Chen, C. Polar and Functionalized Polyolefins: New Catalysts, New Modulation Strategies and New Materials. Acta Polym. Sin. 2018, 11, 1372–1384. [Google Scholar]
  21. Nakamura, A.; Anselment, T.; Claverie, J.; Goodall, B.; Jordan, R.; Mecking, S.; Rieger, B.; Sen, A.; Leeuwen, P.; Nozaki, K. Ortho-Phosphonebenzenesulfonate: A Superb Ligand for Palladium-Catalyzed Coordination-Insertion Copolymerization of Polar Vinyl Monomers. Acc. Chem. Res. 2013, 46, 1438–1449. [Google Scholar] [CrossRef]
  22. Kesti, M.R.; Coates, G.W.; Waymouth, R.M. Homogeneous Ziegler-Natta Polymerization of Functionalized Monomers Catalyzed by Cationic Group IV Metallocenes. J. Am. Chem. Soc. 1992, 114, 9679. [Google Scholar] [CrossRef]
  23. Kim, K.H.; Jo, W.H.; Kwak, S.; Kim, K.U.; Hwang, S.S.; Kim, J. Synthesis and Characterization of Poly(styrene-co-4-[(tert-butyldimethylsilyl)oxy]styrene) as a Precursor of Hydroxyl-Functionalized Syndiotactic Polystyrene. Macromolecules 1999, 32, 8703. [Google Scholar] [CrossRef]
  24. Xu, G.; Chung, T.C. Synthesis of Syndiotactic Polystyrene Derivatives Containing Amino Groups. Macromolecules 2000, 33, 5803. [Google Scholar] [CrossRef]
  25. Kawabe, M.; Murata, M. Syndiospecific Living Polymerization of Silyl-Protected Hydroxystyrene Derivatives and Preparation of Syndiotactic Poly(4-hydroxystyrene) with Narrow Molecular Weight Distribution. Macromol. Chem. Phys. 2001, 202, 3157–3164. [Google Scholar]
  26. Saito, Y.; Nakata, N.; Ishii, A. Copolymerization of Ethylene withiPr3Si-Protected 5-Hexen-1-ol with an [OSSO]-Type Bis(phenolato) Dichloro Zirconium(IV) Complex. Bull. Chem. Soc. Jpn. 2016, 89, 666. [Google Scholar] [CrossRef]
  27. Saito, Y.; Nakata, N.; Ishii, A. Highly Isospecifi c Polymerization of Silyl-Protected ω-Alkenols Using an [OSSO]-Type Bis(phenolato) Dichloro Zirconium(IV) Precatalyst. Macromol. Rapid Commun. 2016, 37, 969. [Google Scholar] [CrossRef]
  28. Zhu, Y.; Li, S.; Liang, H.; Xie, X.; Zhu, F. Titanium Complex with an [OSSO]-Type Bis(Phenolate) Ligand for Ethylene Copolymerization with Vinyl Polar Monomer Based on Group Protection. RSC Adv. 2019, 9, 26582–26587. [Google Scholar] [CrossRef]
  29. Luckham, S.; Nozaki, K. Toward the Copolymerization of Propylene with Polar Comonomers. Acc. Chem. Res. 2021, 54, 344–355. [Google Scholar] [CrossRef]
  30. Chen, Z.; Brookhart, M. Exploring Ethylene/Polar Vinyl Monomer Copolymerizations Using Ni and Pd α-Diimine Catalysts. Acc. Chem. Res. 2018, 51, 1831–1839. [Google Scholar] [CrossRef]
  31. Tan, C.; Chen, C. Emerging Palladium and Nickel Catalysts for Copolymerization of Olefins with Polar Monomers. Angew. Chem. Int. Ed. 2019, 58, 7192–7200. [Google Scholar] [CrossRef]
  32. Ito, S. Palladium-Catalyzed Homo- and Copolymerization of Polar Monomers: Synthesis of Aliphatic and Aromatic Polymers. Bull. Chem. Soc. Jpn. 2018, 91, 251–261. [Google Scholar] [CrossRef]
  33. Nakamura, A.; Ito, S.; Nozaki, K. Coordination-Insertion Copolymerization of Fundamental Polar Monomers. Chem. Rev. 2009, 109, 5215–5244. [Google Scholar] [CrossRef]
  34. Mu, H.; Pan, L.; Song, D.; Li, Y. Neutral Nickel Catalysts for Olefin Homo- and Copolymerization: Relationships between Catalyst Structures and Catalytic Properties. Chem. Rev. 2015, 115, 12091–12137. [Google Scholar] [CrossRef] [PubMed]
  35. Delferro, M.; Marks, T. Multinuclear Olefin Polymerization Catalysts. Chem. Rev. 2011, 111, 2450–2485. [Google Scholar] [CrossRef]
  36. Mu, H.; Zhou, G.; Hu, X.; Jian, Z. Recent Advances in Nickel Mediated Copolymerization of Olefin with Polar Monomers. Coord. Chem. Rev. 2021, 435, 213802–213819. [Google Scholar] [CrossRef]
  37. Birajdar, R.S.; Chikkali, S. Insertion Copolymerization of Functional Olefins: Quo Vadis? Eur. Polym. J. 2021, 143, 110183–110221. [Google Scholar] [CrossRef]
  38. Tan, C.; Zou, C.; Chen, C. Material Properties of Functional Polyethylenes from Transition-Metal-Catalyzed Ethylene−Polar Monomer Copolymerization. Macromolecules 2022, 55, 1910–1922. [Google Scholar] [CrossRef]
  39. Mua, H.; Jian, Z. Stereoselective Copolymerization of Olefin with Polar Monomers to Access Stereoregular Functionalized Polyolefin. Org. Mater. 2022, 4, 178–189. [Google Scholar] [CrossRef]
  40. Boffa, L.; Novak, B. Copolymerization of Polar Monomers with Olefins Using Transition-Metal Complexes. Chem. Rev. 2000, 100, 1479–1493. [Google Scholar] [CrossRef]
  41. Chen, E. Coordination Polymerization of Polar Vinyl Monomers by Single-Site Metal Catalysts. Chem. Rev. 2009, 109, 5157–5214. [Google Scholar] [CrossRef]
  42. Soller, B.S.; Salzinger, S.; Rieger, B. Rare Earth Metal-Mediated Precision Polymerization of Vinylphosphonates and Conjugated Nitrogen-Containing Vinyl Monomers. Chem. Rev. 2016, 116, 1993–2022. [Google Scholar] [CrossRef]
  43. Xu, T. Catalysts for Stereoselective Polymerization of Polar Vinyl Monomer. Prog. Chem. 2017, 29, 285–292. [Google Scholar]
  44. Adams, F.; Pahl, P.; Rieger, B. Metal-Catalyzed Group-Transfer Polymerization: A Versatile Tool for Tailor-Made Functional (Co)Polymers. Chem. Eur. J. 2018, 24, 509–518. [Google Scholar] [CrossRef] [PubMed]
  45. Teator, A.J.; Varner, T.P.; Knutson, P.C.; Sorensen, C.C.; Leibfarth, F.A. 100th Anniversary of Macromolecular Science Viewpoint: The Past, Present, and Future of Stereocontrolled Vinyl Polymerization. ACS Macro Lett. 2020, 9, 1638–1654. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, Y.; Ning, Y.; Caporaso, L.; Cavallo, L.; Chen, E. Catalyst-Site-Controlled Coordination Polymerization of Polar Vinyl Monomers to Highly Syndiotactic Polymers. J. Am. Chem. Soc. 2010, 132, 2695–2709. [Google Scholar] [CrossRef]
  47. Guironnet, D.; Caporaso, L.; Neuwald, B.; Göttker-Schnetmann, I.; Cavallo, L.; Mecking, S.J. Mechanistic Insights on Acrylate Insertion Polymerization. Am. Chem. Soc. 2010, 132, 4418–4426. [Google Scholar] [CrossRef]
  48. Finger, M.; Reek, J.; Bruin, B. Mechanistic Insights on Acrylate Insertion Polymerization. Organometallics 2011, 30, 1094–1101. [Google Scholar] [CrossRef]
  49. Guillaume, S.; Brignou, P.; Susperregui, N.; Maron, L.; Kuzdrowskac, M.; Roesky, P. Bis(Phosphinimino)Methanide Borohydride Complexes of the Rare-Earth Elements as Initiators for the Polymerization of Methyl Methacrylate: Combined Experimental and Computational Investigations. Polym. Chem. 2011, 2, 1728–1736. [Google Scholar] [CrossRef]
  50. Chen, X.; Caporaso, L.; Cavallo, L.; Chen, E. Stereoselectivity in Metallocene-Catalyzed Coordination Polymerization of Renewable Methylene Butyrolactones: From Stereo-Random to Stereo-perfect Polymers. J. Am. Chem. Soc. 2012, 134, 7278–7281. [Google Scholar] [CrossRef]
  51. Neuwald, B.; Falivene, L.; Caporaso, L.; Cavallo, L.; Mecking, S. Exploring Electronic and Steric Effects on the Insertion and Polymerization Reactivity of Phosphinesulfonato PdII Catalysts. Chem. Eur. J. 2013, 19, 17773–17788. [Google Scholar] [CrossRef]
  52. Hu, Y.; Wang, X.; Chen, Y.; Caporaso, L.; Cavallo, L.; Chen, E. Rare-Earth Half-Sandwich Dialkyl and Homoleptic Trialkyl Complexes for Rapid and Stereoselective Polymerization of a Conjugated Polar Olefin. Organometallics 2013, 32, 1459–1465. [Google Scholar] [CrossRef]
  53. Vidal, F.; Falivene, L.; Caporaso, L.; Cavallo, L.; Chen, E. Robust Cross-Linked Stereocomplexes and C60 Inclusion Complexes of Vinyl-Functionalized Stereoregular Polymers Derived from Chemo/Stereoselective Coordination Polymerization. J. Am. Chem. Soc. 2016, 138, 9533–9547. [Google Scholar] [CrossRef]
  54. Sun, J.; Chen, M.; Luo, G.; Chen, C.; Luo, Y. Diphosphazane-monoxide and Phosphine-sulfonate Palladium Catalyzed Ethylene Copolymerization with Polar Monomers: A Computational Study. Organometallics 2019, 38, 638–646. [Google Scholar] [CrossRef]
  55. Pudasaini, B. Yttrium Catalyzed Dialkyl Vinyl Phosphonate Polymerization: Mechanistic Insights on the Precision Polymerization from DFT. Organometallics 2019, 38, 1091–1098. [Google Scholar] [CrossRef]
  56. Zhao, Y.; Luo, G.; Xu, X.; Hou, Z.; Luo, Y. A Computational Study of the Reactivity of Rare-Earth/Phosphorus Lewis Pairs Toward Polymerization of Conjugated Polar Alkenes. Inorg. Chem. Front. 2020, 7, 4600–4610. [Google Scholar] [CrossRef]
  57. Alberoni, C.; D’Alterio, M.; Balducci, G.; Immirzi, B.; Polentarutti, M.; Pellecchia, C.; Milani, B. Tunable “In-Chain” and “At the End of the Branches” Methyl Acrylate Incorporation in the Polyolefin Skeleton through Pd(II) Catalysis. ACS Catal. 2022, 12, 3430–3443. [Google Scholar] [CrossRef]
  58. Du, W.; Zheng, H.; Li, Y.; Cheung, C.; Li, D.; Gao, H.; Deng, H.; Gao, H. Neutral Tridentate α-Sulfonato-β-diimine Nickel Catalyst for (Co)polymerizations of Ethylene and Acrylates. Macromolecules 2022, 55, 3096–3105. [Google Scholar] [CrossRef]
  59. Michalak, A.; Ziegler, T. Late Transition Metals as Homo- and Co-Polymerization Catalysts. Top. Organomet. Chem. 2005, 12, 145–186. [Google Scholar]
  60. Nishiura, M.; Guo, F.; Hou, Z. Half-Sandwich Rare-Earth-Catalyzed Olefin Polymerization, Carbometalation, and Hydroarylation. Acc. Chem. Res. 2015, 48, 2209–2220. [Google Scholar] [CrossRef]
  61. Su, B.; Jia, P.; Wang, Y.; Li, Y.; He, H.; Li, Q. Copolymerization of Ethylene/Polar Monomer Catalyzed by Phosphinoarenesulfonate (PO) Metal Catalysts and the Catalytic Mechanism. Chin. J. Org. Chem. 2016, 36, 2344–2352. [Google Scholar] [CrossRef]
  62. Keyes, A.; Alhan, H.; Ordonez, E.; Ha, U.; Beezer, D.; Dau, H.; Liu, Y.; Tsogtgerel, E.; Jones, G.; Harth, E. Olefins and Vinyl Polar Monomers: Bridging the Gap for Next Generation Materials. Angew. Chem. Int. Ed. 2019, 58, 12370–12391. [Google Scholar] [CrossRef]
  63. Cui, D. Studies on Homo- and Co-polymerizations of Polar and Non-polar Monomers Using Rare-earth Metal Catalysts. Acta Polym. Sin. 2020, 51, 12–29. [Google Scholar]
  64. Chen, J.; Gao, Y.; Marks, T. Early Transition Metal Catalysis for Olefin–Polar Monomer Copolymerization. Angew. Chem. Int. Ed. 2020, 59, 14726–14735. [Google Scholar] [CrossRef]
  65. Zhou, G.; Cui, L.; Mu, H.; Jian, Z. Custom-Made Polar Monomers Utilized in Nickel and Palladium Promoted Olefin Copolymerization. Polym. Chem. 2021, 12, 3878–3892. [Google Scholar] [CrossRef]
  66. Li, Y.; Wang, F.; Cao, Y. Late Transition Metal Complexes for Olefin Copolymerization with Polar Monomers. Chin. J. Org. Chem. 2021, 41, 1396–1433. [Google Scholar] [CrossRef]
  67. Jian, Z.; Falivene, L.; Boffa, G.; Sánchez, S.; Caporaso, L.; Grassi, A.; Mecking, S. Direct Synthesis of Telechelic Polyethylene by Selective Insertion Polymerization. Angew. Chem. Int. Ed. 2016, 55, 14378–14383. [Google Scholar] [CrossRef]
  68. Wang, C.; Luo, G.; Nishiura, M.; Song, G.; Yamamoto, A.; Luo, Y.; Hou, Z. Heteroatom-Assisted Olefin Polymerization by Rare-Earth Metal Catalysts. Sci. Adv. 2017, 3, e1701011. [Google Scholar] [CrossRef] [PubMed]
  69. Chen, J.; Motta, A.; Zhang, J.; Gao, Y.; Marks, T. Mechanism of Organoscandium-Catalyzed Ethylene Copolymerization with Amino-Olefins: A Quantum Chemical Analysis. ACS Catal. 2019, 9, 8810–8818. [Google Scholar] [CrossRef]
  70. You, F.; Liu, H.; Luo, G.; Shi, X. Tridentate Diarylamido-based Pincer Complexes of Nickel and Palladium: Sidearm Effects in Polymerization of Norbornene. Dalton Trans. 2019, 48, 12219–12227. [Google Scholar] [CrossRef]
  71. Hong, C.; Wang, X.; Chen, C. Palladium-Catalyzed Dimerization of Vinyl Ethers: Mechanism, Catalyst Optimization, and Polymerization Applications. Macromolecules 2019, 52, 7123–7129. [Google Scholar]
  72. Zhong, Y.; Douair, I.; Wang, T.; Wu, C.; Maron, L.; Cui, D. Access to Hydroxy-Functionalized Polypropylene through Coordination Polymerization. Angew. Chem. Int. Ed. 2020, 59, 4947–4952. [Google Scholar] [CrossRef]
  73. Mehmood, A.; Xu, X.; Kang, X.; Luo, Y. Origin of Different Chain-End Microstructures in Ethylene/Vinyl Halide Copolymerization Catalysed by Phosphine–Sulfonate Palladium Complexes. New J. Chem. 2020, 44, 16941–16947. [Google Scholar] [CrossRef]
  74. Mehmood, A.; Xu, X.; Raza, W.; Kim, K.; Luo, Y. Mechanistic Studies for Palladium Catalyzed Copolymerization of Ethylene with Vinyl Ethers. Polymers 2020, 12, 2401–2416. [Google Scholar] [CrossRef]
  75. Song, C.; Chen, J.; Fu, Z.; Yan, L.; Gao, F.; Cao, Q.; Li, H.; Yan, X.; Chen, S.; Zhang, S.; et al. Syndiospecific Polymerization of o-Methoxystyrene and Its Silyloxy or Fluorine-Substituted Derivatives by HNC-Ligated Scandium Catalysts: Synthesis of Ultrahigh-Molecular-Weight Functionalized Polymers. Macromolecules 2021, 54, 10838–10849. [Google Scholar] [CrossRef]
  76. Ji, G.; Chen, Z.; Wang, X.; Ning, X.; Xu, C.; Zhang, X.; Tao, W.; Li, J.; Gao, Y.; Shen, Q.; et al. Direct Copolymerization of Ethylene With Protic Comonomers Enabled By Multinuclear Ni Catalyst. Nat. Commun. 2021, 12, 6283–6292. [Google Scholar] [CrossRef]
  77. Zhao, W.; Liu, Z.; Zhao, Y.; Luo, Y.; He, S. Multivariate Linear Regression Models to Predict Monomer Poisoning Effect in Ethylene/Polar Monomer Copolymerization Catalyzed by Late Transition Metals. Inorganics 2022, 10, 26–36. [Google Scholar] [CrossRef]
  78. Zhao, Y.; Li, H.; Xin, S.; Li, H.; Luo, Y.; He, S. DFT Studies on the Early-Transition-Metal-Catalyzed Polymerization of Polar Monomers with a Methylene Spacer between Vinyl and Functional Groups. Organometallics 2022, 41, 3514–3521. [Google Scholar] [CrossRef]
  79. Jiang, Y.; Zhang, Z.; Jiang, H.; Wang, Q.; Li, S.; Cui, D. Polar Group-Promoted Copolymerization of Ethylene with Polar Olefins Macromolecules 2023, 56, 1547−1553.
  80. Zhang, N.; Xue, Z.; Shi, L.; Luo, G. Unveiling the Detailed Mechanism and Origins of Chemo-, Regio-, and Stereoselectivity of Rare-Earth Catalyzed Alternating Copolymerization of Polar and Nonpolar Olefins. Inorg. Chem. 2024, 63, 3544–3559. [Google Scholar] [CrossRef] [PubMed]
  81. Klinger, D.; Wang, C.; Connal, L.; Audus, D.; Jang, S.; Kraemer, S.; Killops, K.; Fredrickson, G.; Kramer, E.; Hawker, C. A Facile Synthesis of Dynamic, Shape-Changing Polymer Particles. Angew. Chem. Int. Ed. 2014, 53, 7018–7022. [Google Scholar] [CrossRef]
  82. Jang, S.; Audus, D.; Klinger, D.; Krogstad, D.; Kim, B.; Cameron, A.; Kim, S.; Delaney, K.; Hur, S.; Killops, K.; et al. Striped, Ellipsoidal Particles by Controlled Assembly of Diblock Copolymers. J. Am. Chem. Soc. 2013, 135, 6649–6657. [Google Scholar] [CrossRef]
  83. Natta, G.; Mazzanti, G.; Dall’Asta, G.; Long, P. Crystalline 2-Vinyl-Pyridine Polymers. Makromol. Chem. 1960, 37, 160–162. [Google Scholar] [CrossRef]
  84. Li, X.; Xu, T. Stereoselective Polymerization of Aromatic Vinyl Polar Monomers. Eur. J. Inorg. Chem. 2023, 26, e202200699. [Google Scholar] [CrossRef]
  85. Xu, T.; Yang, G.; Lu, X. Highly Isotactic and High-Molecular-Weight Poly(2-vinylpyridine) by Coordination Polymerization with Yttrium Bis(phenolate) Ether Catalysts. ACS Catal. 2016, 6, 4907–4913. [Google Scholar] [CrossRef]
  86. Yang, J.; Yu, Y.; Qu, J.; Luo, Y. Effects of Nucleophilic Ligands on the Chain Initiation Efficiency of Polar Monomer Polymerizations Catalyzed by 2-Methoxyethylaminobis(Phenolate)Yttrium Complexes: A DFT Study. Dalton Trans. 2017, 46, 16993–16999. [Google Scholar] [CrossRef]
  87. Adams, F.; Machat, M.; Altenbuchner, P.; Ehrmaier, J.; Pöthig, A.; Karsili, T.; Rieger, B. Toolbox of Nonmetallocene Lanthanides: Multifunctional Catalysts in Group-Transfer Polymerization. Inorg. Chem. 2017, 56, 9754–9764. [Google Scholar] [CrossRef] [PubMed]
  88. Yan, C.; Xu, T.; Lu, X. From Stereochemically Tunable Homopolymers to Stereomultiblock Copolymers: Lewis Base Regulates Stereochemistry in the Coordination Polymerization of 2-Vinylpyridine. Macromolecules 2018, 51, 2240–2246. [Google Scholar] [CrossRef]
  89. Weger, M.; Giuman, M.; Knaus, M.; Ackermann, M.; Drees, M.; Hornung, J.; Altmann, P.; Fischer, R.; Rieger, B. Single-Site, Organometallic Aluminum Catalysts for the Precise Group Transfer Polymerization of Michael-Type Monomers. Chem. Eur. J. 2018, 24, 14950–14957. [Google Scholar] [CrossRef]
  90. Su, Y.; Zhao, Y.; Zhang, H.; Luo, Y.; Xu, X. Rare-Earth Aryloxide/Ylide-Functionalized Phosphine Frustrated Lewis Pairs for the Polymerization of 4-Vinylpyridine and Its Derivatives. Macromolecules 2021, 54, 7724–7731. [Google Scholar] [CrossRef]
  91. Jaymand, M. Recent Progress in the Chemical Modification of Syndiotactic Polystyrene. Polym. Chem. 2014, 5, 2663–2690. [Google Scholar] [CrossRef]
  92. Liu, D.; Yao, C.; Wang, R.; Wang, M.; Wang, Z.; Wu, C.; Lin, F.; Li, S.; Wan, X.; Cui, D. Highly Isoselective Coordination Polymerization of ortho-Methoxystyrene with β-Diketiminato Rare-Earth-Metal Precursors. Angew. Chem. Int. Ed. 2015, 54, 5205–5209. [Google Scholar] [CrossRef] [PubMed]
  93. Liu, D.; Wang, M.; Wang, Z.; Wu, C.; Pan, Y.; Cui, D. Stereoselective Copolymerization of Unprotected Polar and Nonpolar Styrenes by anYttrium Precursor: Control of Polar-Group Distribution and Mechanism. Angew. Chem. Int. Ed. 2017, 56, 2714–2719. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, Z.; Wang, M.; Liu, J.; Liu, D.; Cui, D. Rapid Syndiospecific (Co)Polymerization of Fluorostyrene with High Monomer Conversion. Chem. Eur. J. 2017, 23, 18151–18155. [Google Scholar] [CrossRef]
  95. Zhao, Y.; Luo, G.; Wang, X.; Kang, X.; Cui, D.; Hou, Z.; Luo, Y. DFT Studies on the Polymerization of Functionalized Styrenes Catalyzed by Rare-Earth-Metal Complexes: Factors Affecting C−H Activation Relevant to Step-Growth Polymerization. Organometallics 2018, 37, 3210–3218. [Google Scholar] [CrossRef]
  96. Liu, D.; Wang, M.; Chai, Y.; Wan, X.; Cui, D. Self-Activated Coordination Polymerization of Alkoxystyrenes by a Yttrium Precursor: Stereocontrol and Mechanism. ACS Catal. 2019, 9, 2618–2625. [Google Scholar] [CrossRef]
  97. Zhao, Y.; Luo, G.; Kang, X.; Guo, F.; Zhu, X.; Zheng, R.; Hou, Z.; Luo, Y. “C–H⋅⋅⋅π Interaction” Regulates the Stereoselectivity in Olefin Polymerization. Chem. Commun. 2019, 55, 6689–6692. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, T.; Wu, C.; Ji, X.; Cui, D. Direct Synthesis of Functional Thermoplastic Elastomer with Excellent Mechanical Properties by Scandium-Catalyzed Copolymerization of Ethylene and Fluorostyrenes. Angew. Chem. Int. Ed. 2021, 60, 25735–25740. [Google Scholar] [CrossRef]
  99. Liu, Z.; Liu, B.; Zhao, Z.; Cui, D. Chemo- and Stereoselective Polymerization of Polar Divinyl Monomers by Rare-Earth Complexes. Macromolecules 2021, 54, 3181–3190. [Google Scholar] [CrossRef]
  100. Jiang, Y.; Zhang, Z.; Li, S.; Cui, D. Isospecific (Co)polymerization of Unmasked Polar Styrenes by Neutral Rare-Earth Metal Catalysts. Angew. Chem. Int. Ed. 2022, 61, e202112966. [Google Scholar] [CrossRef]
  101. Wen, X.; Xu, X.; Li, H.; Zhao, Y.; Luo, Y. Theoretical Mechanistic Studies of the Polymerization of Functionalized Styrenes Catalyzed by Rare-Earth-Metal Complexes: Stereoselectivity Regulation. Organometallics 2022, 41, 2466–2473. [Google Scholar]
  102. Wen, X.; Xing, X.; Zhang, W.; Lu, H.; Zhu, L.; Xu, X.; Zhao, Y.; Cui, D.; Luo, Y. DFT Studies on the Effect of Additives on Stereoselectivity in the Polymerization of Styrene Catalyzed by Rare-Earth Metal Complexes. New J. Chem. 2023, 47, 19516–19522. [Google Scholar] [CrossRef]
  103. Leborgne, A.; Vincens, V.; Joulgard, M.; Spassky, N. RingOpening Oligomerization Reactions Using Aluminium Complexes of Schiff’s Bases as Initiators. Makromol. Chem. Macromol. Symp. 1993, 73, 37–46. [Google Scholar] [CrossRef]
  104. Wisniewski, M.; Le Borgne, A.; Spassky, N. Synthesis and Properties of (D)- and (L)-Lactide Stereocopolymers Using the System Achiral Schiff’s Base/Aluminium Methoxide as Initiator. Macromol. Chem. Phys. 1997, 198, 1227–1238. [Google Scholar] [CrossRef]
  105. Ovitt, T.M.; Coates, G.W. Stereoselective Ring-Opening Polymerization of mesoLactide: Synthesis of Syndiotactic Poly(lactic acid). J. Am. Chem. Soc. 1999, 121, 4072–4073. [Google Scholar] [CrossRef]
  106. Cheng, M.; Attygalle, A.B.; Lobkovsky, E.B.; Coates, G.W. Single-Site Catalysts for Ring-Opening Polymerization: Synthesis of Heterotactic Poly(lactic acid) from rac-Lactide. J. Am. Chem. Soc. 1999, 121, 11583–11584. [Google Scholar] [CrossRef]
  107. Chamberlain, B.M.; Cheng, M.; Moore, D.R.; Ovitt, T.M.; Lobkovsky, E.B.; Coates, G.W. Polymerization of Lactide with Zinc and Magnesium β-Diiminate Complexes: Stereocontrol and Mechanism. J. Am. Chem. Soc. 2001, 123, 3229–3238. [Google Scholar] [CrossRef] [PubMed]
  108. Ovitt, T.M.; Coates, G.W. Stereochemistry of Lactide Polymerization with Chiral Catalysts: New Opportunities for Stereocontrol Using Polymer Exchange Mechanisms. J. Am. Chem. Soc. 2002, 124, 1316–1326. [Google Scholar] [CrossRef] [PubMed]
  109. 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] [PubMed]
  110. Matsumura, S. Enzymatic Synthesis of Polyesters via Ring-Opening Polymerization. Adv. Polym. Sci. 2006, 194, 95–132. [Google Scholar]
  111. Kamber, N.E.; Jeong, W.; Waymouth, R.M.; Pratt, R.C.; Lohmeijer, B.; Hedrick, J.L. Organocatalytic Ring-Opening Polymerization. Chem. Rev. 2007, 107, 5813–5840. [Google Scholar] [CrossRef]
  112. 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]
  113. Dove, A.P. Organic Catalysis for Ring-Opening Polymerization. ACS Macro Lett. 2012, 1, 1409–1412. [Google Scholar] [CrossRef]
  114. Sarazin, Y.; Carpentier, J. Discrete Cationic Complexes for Ring-Opening Polymerization Catalysis of Cyclic Esters and Epoxides. Chem. Rev. 2015, 115, 3564–3614. [Google Scholar] [CrossRef]
  115. Carpentier, J. Rare-Earth Complexes Supported by Tripodal Tetradentate Bis(phenolate) Ligands: A Privileged Class of Catalysts for Ring-Opening Polymerization of Cyclic Esters. Organometallics 2015, 34, 4175–4189. [Google Scholar] [CrossRef]
  116. Redshaw, C. Metallocalixarene catalysts: α-olefin polymerization and ROP of cyclic esters. Dalton Trans. 2016, 45, 9018–9030. [Google Scholar] [CrossRef]
  117. 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, 4117–4167. [Google Scholar] [CrossRef]
  118. Wei, J.; Diaconescu, P. Redox-Switchable Ring-Opening Polymerization with Ferrocene Derivatives. Acc. Chem. Res. 2019, 52, 415–424. [Google Scholar] [CrossRef] [PubMed]
  119. Shen, T.; Ni, X.; Ling, J. Recent Progress in Ring-opening Polymerizations Catalyzed by Rare Earth Catalysts. Acta Polym. Sin. 2021, 52, 445–455. [Google Scholar]
  120. Shawver, N.; Doerr, A.M.; Long, B. A perspective on redox-switchable ring-opening polymerization. J. Polym. Sci. 2023, 61, 361–371. [Google Scholar] [CrossRef]
  121. Guillaume, S.; Maron, L.; Roesky, P. Handbook on the Physics and Chemistry of Rare Earths; ScienceDirect: Amsterdam, The Netherlands, 2014; Volume 44, pp. 1–86. [Google Scholar]
  122. Sutar, A.K.; Maharana, T.; Routaray, A.; Nath, N. Ring-Opening Polymerization of Lactide. In Advanced Catalytic Materials; John Wiley & Sons: Hoboken, NJ, USA, 2015; pp. 193–224. [Google Scholar]
Scheme 1. Structures of polyolefins and functionalized polyolefins.
Scheme 1. Structures of polyolefins and functionalized polyolefins.
Inorganics 12 00233 sch001
Scheme 2. The polymerization mechanism of four different types of polar monomers catalyzed by transition metal complexes.
Scheme 2. The polymerization mechanism of four different types of polar monomers catalyzed by transition metal complexes.
Inorganics 12 00233 sch002
Scheme 3. Several transition metal complexes used for catalytic calculation of conjugated polar monomer polymerization [46,47,48,49,50,51,52,53,54,55,56,57,58]. * Represents the corresponding author.
Scheme 3. Several transition metal complexes used for catalytic calculation of conjugated polar monomer polymerization [46,47,48,49,50,51,52,53,54,55,56,57,58]. * Represents the corresponding author.
Inorganics 12 00233 sch003
Figure 1. Ansa-Zirconocene complexes incorporating Cs-symmetric [R2E(Cp)(Flu)]-Based ligands and their variants investigated. Reprinted from ref. [46]. Study Copyright (2010) American Chemical Society.
Figure 1. Ansa-Zirconocene complexes incorporating Cs-symmetric [R2E(Cp)(Flu)]-Based ligands and their variants investigated. Reprinted from ref. [46]. Study Copyright (2010) American Chemical Society.
Inorganics 12 00233 g001
Figure 2. Transition states generating a stereomistake in MMA addition to the enolate growing chain with 2 (a), 6 (b), and 11 (c). Distances in Å and ∆Estereo in kcal mol−1. Reprinted from ref. [46]. Study Copyright (2010) American Chemical Society.
Figure 2. Transition states generating a stereomistake in MMA addition to the enolate growing chain with 2 (a), 6 (b), and 11 (c). Distances in Å and ∆Estereo in kcal mol−1. Reprinted from ref. [46]. Study Copyright (2010) American Chemical Society.
Inorganics 12 00233 g002
Figure 3. Energy profiles corresponding to the epimerization reaction with systems 2, 3, and 6 in CH2Cl2. Reprinted from ref. [46]. Study Copyright (2010) American Chemical Society.
Figure 3. Energy profiles corresponding to the epimerization reaction with systems 2, 3, and 6 in CH2Cl2. Reprinted from ref. [46]. Study Copyright (2010) American Chemical Society.
Inorganics 12 00233 g003
Figure 4. Methyl acrylate and ethylene insertion from 3-meso. Relative energies of the various species are given in parentheses (in kJ mol−1; black: intermediates; red: transition states). Reprinted from ref. [47]. Study Copyright (2010) American Chemical Society.
Figure 4. Methyl acrylate and ethylene insertion from 3-meso. Relative energies of the various species are given in parentheses (in kJ mol−1; black: intermediates; red: transition states). Reprinted from ref. [47]. Study Copyright (2010) American Chemical Society.
Inorganics 12 00233 g004
Figure 5. Pathways that follow β-H elimination, leading to isomerization of the polymer chain. Reprinted from ref. [48]. Study Copyright (2010) American Chemical Society.
Figure 5. Pathways that follow β-H elimination, leading to isomerization of the polymer chain. Reprinted from ref. [48]. Study Copyright (2010) American Chemical Society.
Inorganics 12 00233 g005
Figure 6. Calculated free energy profiles for the reaction of MMA with [{CH(PPh2NSiMe3)2}La(BH4)2(THF)]. Reprinted from ref. [49]. Study Copyright (2011) Royal Society of Chemistry.
Figure 6. Calculated free energy profiles for the reaction of MMA with [{CH(PPh2NSiMe3)2}La(BH4)2(THF)]. Reprinted from ref. [49]. Study Copyright (2011) Royal Society of Chemistry.
Inorganics 12 00233 g006
Figure 7. TS geometries for the competitive βMMBL addition on an R chain (ac) and on an S chain (df). In all cases the monomer and the chain are located on the same side (cis) of the equatorial belt of the (R,R)-(EBI)Zr catalyst. The energies (kcal/mol) are relative to the most stable TSs Chain(S)_Mon(S)_si TS structure (d). Reprinted from ref. [50]. Study Copyright (2012) Royal Society of Chemistry.
Figure 7. TS geometries for the competitive βMMBL addition on an R chain (ac) and on an S chain (df). In all cases the monomer and the chain are located on the same side (cis) of the equatorial belt of the (R,R)-(EBI)Zr catalyst. The energies (kcal/mol) are relative to the most stable TSs Chain(S)_Mon(S)_si TS structure (d). Reprinted from ref. [50]. Study Copyright (2012) Royal Society of Chemistry.
Inorganics 12 00233 g007
Figure 8. Phosphinesulfonato substitution patterns studied. Reprinted from ref. [51]. Study Copyright (2013) Wiley.
Figure 8. Phosphinesulfonato substitution patterns studied. Reprinted from ref. [51]. Study Copyright (2013) Wiley.
Inorganics 12 00233 g008
Figure 9. Energy scheme and nomenclature for ethylene insertion. Reprinted from ref. [51]. Study Copyright (2013) Wiley.
Figure 9. Energy scheme and nomenclature for ethylene insertion. Reprinted from ref. [51]. Study Copyright (2013) Wiley.
Inorganics 12 00233 g009
Figure 10. Transition state geometries for the competitive βMMBL addition on S-l-Y. Reprinted from ref. [52]. Study Copyright (2013) American Chemical Society.
Figure 10. Transition state geometries for the competitive βMMBL addition on S-l-Y. Reprinted from ref. [52]. Study Copyright (2013) American Chemical Society.
Inorganics 12 00233 g010
Figure 11. (I) Transition state geometries for the competitive addition at the (a,c) si and (b,d) re faces of the growing chain for (a,b) AMA and (c,d) VMA with the (S, S)–EBI−Zr system. (II) Transition state geometries of the competitive addition at the (a,c) si and (b,d) re faces of the growing chain for (a,b) AMA and (c,d) VMA with the (S)–Ph2C(Cp)Flu–Zr system. The free energies (in kcal/mol, in DCM) are relative to the TS involving the si face of the growing chain (a,c). The dashed lines indicate the emerging C−C bonds. Reprinted from ref. [53]. Study Copyright (2016) American Chemical Society.
Figure 11. (I) Transition state geometries for the competitive addition at the (a,c) si and (b,d) re faces of the growing chain for (a,b) AMA and (c,d) VMA with the (S, S)–EBI−Zr system. (II) Transition state geometries of the competitive addition at the (a,c) si and (b,d) re faces of the growing chain for (a,b) AMA and (c,d) VMA with the (S)–Ph2C(Cp)Flu–Zr system. The free energies (in kcal/mol, in DCM) are relative to the TS involving the si face of the growing chain (a,c). The dashed lines indicate the emerging C−C bonds. Reprinted from ref. [53]. Study Copyright (2016) American Chemical Society.
Inorganics 12 00233 g011
Figure 12. Energy profiles for the copolymerization of MA and ethylene mediated by (A) and (B), respectively. The energies are relative to corresponding catalysts and monomers. Reprinted from ref. [54]. Study Copyright (2019) American Chemical Society.
Figure 12. Energy profiles for the copolymerization of MA and ethylene mediated by (A) and (B), respectively. The energies are relative to corresponding catalysts and monomers. Reprinted from ref. [54]. Study Copyright (2019) American Chemical Society.
Inorganics 12 00233 g012
Figure 13. (a) Gibbs energy profile, in kJ mol−1, of the catalytic resting state (A4) following the initiation by methyl addition. (b) Gibbs energy profile, in kJ mol−1, of the catalytic resting state (D4) following the initiation by deprotonation. Calculated energies are in parentheses. Reprinted from ref. [55]. Study Copyright (2019) American Chemical Society.
Figure 13. (a) Gibbs energy profile, in kJ mol−1, of the catalytic resting state (A4) following the initiation by methyl addition. (b) Gibbs energy profile, in kJ mol−1, of the catalytic resting state (D4) following the initiation by deprotonation. Calculated energies are in parentheses. Reprinted from ref. [55]. Study Copyright (2019) American Chemical Society.
Inorganics 12 00233 g013
Figure 14. Computed energy profiles for 3/PEt3 mediated (intermolecular RE/P systems) bimetallic pathway for MMA enchainment. Free energies (kcal mol−1) are relative to isolated reactants. Reprinted from ref. [56]. Study Copyright (2020) Royal Society of Chemistry.
Figure 14. Computed energy profiles for 3/PEt3 mediated (intermolecular RE/P systems) bimetallic pathway for MMA enchainment. Free energies (kcal mol−1) are relative to isolated reactants. Reprinted from ref. [56]. Study Copyright (2020) Royal Society of Chemistry.
Inorganics 12 00233 g014
Figure 15. MEPs of the ethylene/MA copolymerization mechanism: MEP for 1SPh (right) and MEP for 1 (left). Reprinted from ref. [57]. Study Copyright (2022) American Chemical Society.
Figure 15. MEPs of the ethylene/MA copolymerization mechanism: MEP for 1SPh (right) and MEP for 1 (left). Reprinted from ref. [57]. Study Copyright (2022) American Chemical Society.
Inorganics 12 00233 g015
Figure 16. (A) Key monomer-coordinated nickel intermediates. (B) Proposed reaction pathway and free energies (kcal mol−1) of nickel-catalyzed E/MA copolymerization. Reprinted from ref. [58]. Study Copyright (2022) American Chemical Society.
Figure 16. (A) Key monomer-coordinated nickel intermediates. (B) Proposed reaction pathway and free energies (kcal mol−1) of nickel-catalyzed E/MA copolymerization. Reprinted from ref. [58]. Study Copyright (2022) American Chemical Society.
Inorganics 12 00233 g016
Scheme 4. Several transition metal complexes used for catalytic calculation of non-conjugated polar monomer polymerization [67,68,69,70,71,72,73,74,75,76,77,78,79,80].
Scheme 4. Several transition metal complexes used for catalytic calculation of non-conjugated polar monomer polymerization [67,68,69,70,71,72,73,74,75,76,77,78,79,80].
Inorganics 12 00233 sch004
Figure 17. Reaction mechanisms of the competitive pathways for ethylene insertion, the VF insertion, and the chain transfer reaction to ethylene and VF from L1Pd-CH(CH2-CH3) (C4H3O) species (free energies in toluene in kcal mol−1). Reprinted from ref. [67]. Study Copyright (2013) Wiley.
Figure 17. Reaction mechanisms of the competitive pathways for ethylene insertion, the VF insertion, and the chain transfer reaction to ethylene and VF from L1Pd-CH(CH2-CH3) (C4H3O) species (free energies in toluene in kcal mol−1). Reprinted from ref. [67]. Study Copyright (2013) Wiley.
Inorganics 12 00233 g017
Figure 18. Computational analysis of the polymerization of 1i by the cationic species Cat. (A) DFT-calculated energy profile of the polymerization of 1i by Cat. (B) Structures of the stationary points shown in the energy profile. The less favored pathway is pale-colored. R = CH2C6H4NMe2-o. Reprinted from ref. [68]. Study Copyright (2017) Science.
Figure 18. Computational analysis of the polymerization of 1i by the cationic species Cat. (A) DFT-calculated energy profile of the polymerization of 1i by Cat. (B) Structures of the stationary points shown in the energy profile. The less favored pathway is pale-colored. R = CH2C6H4NMe2-o. Reprinted from ref. [68]. Study Copyright (2017) Science.
Inorganics 12 00233 g018
Figure 19. (a) Gibbs free energy profiles (kcal/mol) for the N−(1−butenyl)nPr2 incorporation (1,2−insertion) step evaluated along three pathways. (b) Gibbs free energy profiles (kcal/mol) for the long AO incorporation (1,2 insertion) step evaluated along three pathways. Pathway A, no amine coordination; pathway B, self−assisted with amine coordination; pathway C, unassisted with amine coordination. Reprinted from ref. [69]. Study Copyright (2019) American Chemical Society.
Figure 19. (a) Gibbs free energy profiles (kcal/mol) for the N−(1−butenyl)nPr2 incorporation (1,2−insertion) step evaluated along three pathways. (b) Gibbs free energy profiles (kcal/mol) for the long AO incorporation (1,2 insertion) step evaluated along three pathways. Pathway A, no amine coordination; pathway B, self−assisted with amine coordination; pathway C, unassisted with amine coordination. Reprinted from ref. [69]. Study Copyright (2019) American Chemical Society.
Inorganics 12 00233 g019
Figure 20. Topographic steric maps of the active species for the [PNNox] (APd-P), [SNNox] (APd-P), and [NNNox] (APd-P) palladium complexes. %V_Bur denotes the percent of buried volume of Pd atom. Reprinted from ref. [70]. Study Copyright (2019) Royal Society of Chemistry.
Figure 20. Topographic steric maps of the active species for the [PNNox] (APd-P), [SNNox] (APd-P), and [NNNox] (APd-P) palladium complexes. %V_Bur denotes the percent of buried volume of Pd atom. Reprinted from ref. [70]. Study Copyright (2019) Royal Society of Chemistry.
Inorganics 12 00233 g020
Figure 21. Computed Gibbs energy profiles (free energy in kcal/mol) for the dimerization of CH2=CHOCH3 catalyzed by Pd−OMe+. C1−C and A−O+C contain the same palladium complex, and their energy differences are due to the presence of two vinyl ether molecules for C1−C and the presence of dimer product for A−O+C. Reprinted from ref. [71]. Study Copyright (2019) American Chemical Society.
Figure 21. Computed Gibbs energy profiles (free energy in kcal/mol) for the dimerization of CH2=CHOCH3 catalyzed by Pd−OMe+. C1−C and A−O+C contain the same palladium complex, and their energy differences are due to the presence of two vinyl ether molecules for C1−C and the presence of dimer product for A−O+C. Reprinted from ref. [71]. Study Copyright (2019) American Chemical Society.
Inorganics 12 00233 g021
Figure 22. Proposed mechanism for the polymerization of alkylallene monomers A, B, C, and G by catalyst 3 [computed enthalpy of monomer B is provided (kcal mol−1)]. Reprinted from ref. [72]. Study Copyright (2020) Wiley.
Figure 22. Proposed mechanism for the polymerization of alkylallene monomers A, B, C, and G by catalyst 3 [computed enthalpy of monomer B is provided (kcal mol−1)]. Reprinted from ref. [72]. Study Copyright (2020) Wiley.
Inorganics 12 00233 g022
Figure 23. Calculated energy profiles for the generation of units –CH2–CHX2 mediated by A’. Free energies are relative to the energy sum of catalyst (A’) and monomers (E/VX). Reprinted from ref. [73]. Study Copyright (2020) Royal Society of Chemistry.
Figure 23. Calculated energy profiles for the generation of units –CH2–CHX2 mediated by A’. Free energies are relative to the energy sum of catalyst (A’) and monomers (E/VX). Reprinted from ref. [73]. Study Copyright (2020) Royal Society of Chemistry.
Inorganics 12 00233 g023
Figure 24. Energy profiles for second monomer insertion into VE pre-inserted intermediate. Reprinted from ref. [74]. Study Copyright (2020) MDPI.
Figure 24. Energy profiles for second monomer insertion into VE pre-inserted intermediate. Reprinted from ref. [74]. Study Copyright (2020) MDPI.
Inorganics 12 00233 g024
Figure 25. Energy profiles for 2,1-insertion of oMOS and generation of stereoregularity. Reprinted from ref. [75]. Copyright (2021) American Chemical Society.
Figure 25. Energy profiles for 2,1-insertion of oMOS and generation of stereoregularity. Reprinted from ref. [75]. Copyright (2021) American Chemical Society.
Inorganics 12 00233 g025
Figure 26. Computed energy profiles for Ni-mediated copolymerization of ethylene with VA anion (the effect of large aluminum counterion was not considered). The energies are relative to intermediate 18. Reprinted from ref. [76]. Copyright (2021) Nature.
Figure 26. Computed energy profiles for Ni-mediated copolymerization of ethylene with VA anion (the effect of large aluminum counterion was not considered). The energies are relative to intermediate 18. Reprinted from ref. [76]. Copyright (2021) Nature.
Inorganics 12 00233 g026
Figure 27. Plot of computed vs. predicted ΔΔE(π-σ) (kcal mol−1) for complexes IINi (a) and IIPd (b) using the multivariate linear regression models. Reprinted from ref. [77]. Copyright (2022) MDPI.
Figure 27. Plot of computed vs. predicted ΔΔE(π-σ) (kcal mol−1) for complexes IINi (a) and IIPd (b) using the multivariate linear regression models. Reprinted from ref. [77]. Copyright (2022) MDPI.
Inorganics 12 00233 g027
Figure 28. Computed energy profiles for B+-mediated vinyl-coordination insertion (gray curve), heteroatom-chelating vinyl-coordination insertion (black curve), and heteroatom coordination (pink curve) of 1. Free energies are relative to isolated reactants. Reprinted from ref. [78]. Copyright (2022) American Chemical Society.
Figure 28. Computed energy profiles for B+-mediated vinyl-coordination insertion (gray curve), heteroatom-chelating vinyl-coordination insertion (black curve), and heteroatom coordination (pink curve) of 1. Free energies are relative to isolated reactants. Reprinted from ref. [78]. Copyright (2022) American Chemical Society.
Inorganics 12 00233 g028
Figure 29. Energy profiles for the POH incorporation (2,1-insertion) step using precatalysts 3 and 4. Reprinted from ref. [79]. Copyright (2023) American Chemical Society.
Figure 29. Energy profiles for the POH incorporation (2,1-insertion) step using precatalysts 3 and 4. Reprinted from ref. [79]. Copyright (2023) American Chemical Society.
Inorganics 12 00233 g029
Figure 30. Origin of monomer insertion on the chain repeating units. Bond distances and energies are given in Å and kcal/mol, respectively. * represents chirality. Reprinted from ref. [80]. Copyright (2024) American Chemical Society.
Figure 30. Origin of monomer insertion on the chain repeating units. Bond distances and energies are given in Å and kcal/mol, respectively. * represents chirality. Reprinted from ref. [80]. Copyright (2024) American Chemical Society.
Inorganics 12 00233 g030
Scheme 5. Several rare-earth metal complexes used for calculation of 2-vinylpyridine polymerization [85,86,87,88]. * represents the corresponding author.
Scheme 5. Several rare-earth metal complexes used for calculation of 2-vinylpyridine polymerization [85,86,87,88]. * represents the corresponding author.
Inorganics 12 00233 sch005
Figure 31. Proposed initiation and propagation in 2-VP polymerization catalyzed by Y bis(ether phenolate) complex. Reprinted from ref. [85]. Copyright (2016) American Chemical Society.
Figure 31. Proposed initiation and propagation in 2-VP polymerization catalyzed by Y bis(ether phenolate) complex. Reprinted from ref. [85]. Copyright (2016) American Chemical Society.
Inorganics 12 00233 g031
Figure 32. (a) Energy profile for the GTP of 2VP initiated by complexes 1a, 1b, and 1c. (b) Frontier molecular orbitals of the fragments in complexes 7a, 7b and 7c (all hydrogen atoms are omitted for clarity). Reprinted from ref. [86]. Copyright (2017) Royal Society of Chemistry.
Figure 32. (a) Energy profile for the GTP of 2VP initiated by complexes 1a, 1b, and 1c. (b) Frontier molecular orbitals of the fragments in complexes 7a, 7b and 7c (all hydrogen atoms are omitted for clarity). Reprinted from ref. [86]. Copyright (2017) Royal Society of Chemistry.
Inorganics 12 00233 g032
Figure 33. Reaction path for the initiation process: catalyst 1a (black) and 1b (red) with 2VP and DMAA. Reprinted from ref. [87]. Copyright (2017) American Chemical Society.
Figure 33. Reaction path for the initiation process: catalyst 1a (black) and 1b (red) with 2VP and DMAA. Reprinted from ref. [87]. Copyright (2017) American Chemical Society.
Inorganics 12 00233 g033
Figure 34. Energy profiles for the propagation process. Reprinted from ref. [88]. Copyright (2018) American Chemical Society.
Figure 34. Energy profiles for the propagation process. Reprinted from ref. [88]. Copyright (2018) American Chemical Society.
Inorganics 12 00233 g034
Figure 35. Distortion/interaction analyses (kcal mol−1) for TS1P2 (a) and TS1PEt (b) and Mulliken charge in C1, P2, and PEt3. The values in black denote bond length. Reprinted from ref. [90]. Copyright (2021) American Chemical Society.
Figure 35. Distortion/interaction analyses (kcal mol−1) for TS1P2 (a) and TS1PEt (b) and Mulliken charge in C1, P2, and PEt3. The values in black denote bond length. Reprinted from ref. [90]. Copyright (2021) American Chemical Society.
Inorganics 12 00233 g035
Scheme 6. Transition metal complexes studied in DFT modeling of polymerization of heteroatom-functionalized styrene [92,93,94,95,96,97,98,99,100,101,102,103].
Scheme 6. Transition metal complexes studied in DFT modeling of polymerization of heteroatom-functionalized styrene [92,93,94,95,96,97,98,99,100,101,102,103].
Inorganics 12 00233 sch006
Figure 36. Energy profiles for 2,1–insertion of oMOS and generation of stereoregularity. R = 2,6-dimethylphenyl. Reprinted from ref. [92]. Copyright (2015) Wiley.
Figure 36. Energy profiles for 2,1–insertion of oMOS and generation of stereoregularity. R = 2,6-dimethylphenyl. Reprinted from ref. [92]. Copyright (2015) Wiley.
Inorganics 12 00233 g036
Figure 37. (a) Energy profiles for the insertion of oMOS and St at 298 K. (b) Energy profiles for the insertion of pMOS and St at 298 K. Reprinted from ref. [93]. Copyright (2017) Wiley.
Figure 37. (a) Energy profiles for the insertion of oMOS and St at 298 K. (b) Energy profiles for the insertion of pMOS and St at 298 K. Reprinted from ref. [93]. Copyright (2017) Wiley.
Inorganics 12 00233 g037
Figure 38. The energy profiles for coordination-insertion polymerization of pFS with complex 4 (a) and 1 (b). Reprinted from ref. [94]. Study Copyright (2017) Wiley.
Figure 38. The energy profiles for coordination-insertion polymerization of pFS with complex 4 (a) and 1 (b). Reprinted from ref. [94]. Study Copyright (2017) Wiley.
Inorganics 12 00233 g038
Figure 39. Computed energy profiles for C–H activation (black curve) and C=C insertion (red curve) polymerization of the first and second molecules of oMOS by cationic species A+. Free energies are relative to the energy sum of species A+ and corresponding monomers. Reprinted from ref. [95]. Copyright (2018) American Chemical Society.
Figure 39. Computed energy profiles for C–H activation (black curve) and C=C insertion (red curve) polymerization of the first and second molecules of oMOS by cationic species A+. Free energies are relative to the energy sum of species A+ and corresponding monomers. Reprinted from ref. [95]. Copyright (2018) American Chemical Society.
Inorganics 12 00233 g039
Figure 40. Energy profiles for insertion of oMOS and generation of stereoregularity. Reprinted from ref. [96]. Copyright (2019) American Chemical Society.
Figure 40. Energy profiles for insertion of oMOS and generation of stereoregularity. Reprinted from ref. [96]. Copyright (2019) American Chemical Society.
Inorganics 12 00233 g040
Figure 41. Energy gap (ΔΔG (iso–syn), kcal mol−1) between the corresponding iso- and syn-enchainment transition states for various monomer polymerizations catalyzed by a and NPA charges (data in blue) on the phenyl rings of halogenated styrene in TS2asyn. Reprinted from ref. [97]. Copyright (2019) Royal Society of Chemistry.
Figure 41. Energy gap (ΔΔG (iso–syn), kcal mol−1) between the corresponding iso- and syn-enchainment transition states for various monomer polymerizations catalyzed by a and NPA charges (data in blue) on the phenyl rings of halogenated styrene in TS2asyn. Reprinted from ref. [97]. Copyright (2019) Royal Society of Chemistry.
Inorganics 12 00233 g041
Figure 42. Energy profiles for the copolymerization of E and pFS using complex 3. Reprinted from ref. [98]. Copyright (2021) Wiley.
Figure 42. Energy profiles for the copolymerization of E and pFS using complex 3. Reprinted from ref. [98]. Copyright (2021) Wiley.
Inorganics 12 00233 g042
Figure 43. Computed energy profiles (free energy in kcal mol−1) for AOS polymerization under the conditions of two monomers coordinating to the center metal. Reprinted from ref. [99]. Copyright (2021) American Chemical Society.
Figure 43. Computed energy profiles (free energy in kcal mol−1) for AOS polymerization under the conditions of two monomers coordinating to the center metal. Reprinted from ref. [99]. Copyright (2021) American Chemical Society.
Inorganics 12 00233 g043
Figure 44. Energy profiles for the coordination and insertion of pMOS and generation of stereoregularity. Reprinted from ref. [100]. Copyright (2022) Wiley.
Figure 44. Energy profiles for the coordination and insertion of pMOS and generation of stereoregularity. Reprinted from ref. [100]. Copyright (2022) Wiley.
Inorganics 12 00233 g044
Figure 45. Geometric structures (distances in Å). The values in blue denote atomic NPA charges of Sc, C1, and C2; green denotes the total NPA charges of C6 and H. Reprinted from ref. [101]. Copyright (2022) American Chemical Society.
Figure 45. Geometric structures (distances in Å). The values in blue denote atomic NPA charges of Sc, C1, and C2; green denotes the total NPA charges of C6 and H. Reprinted from ref. [101]. Copyright (2022) American Chemical Society.
Inorganics 12 00233 g045
Figure 46. (a) Interaction analyses (energies in kcal mol−1) for aTSiso, aTSsyn. (b) Interaction analyses (energies in kcal mol−1) for bTSiso and bTSsyn. The green cloud denotes a weak interaction. Reprinted from ref. [102]. Copyright (2023) Royal Society of Chemistry.
Figure 46. (a) Interaction analyses (energies in kcal mol−1) for aTSiso, aTSsyn. (b) Interaction analyses (energies in kcal mol−1) for bTSiso and bTSsyn. The green cloud denotes a weak interaction. Reprinted from ref. [102]. Copyright (2023) Royal Society of Chemistry.
Inorganics 12 00233 g046
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, Y.; Zhang, Z.; Luo, Y. DFT Modeling of Coordination Polymerization of Polar Olefin Monomers by Molecular Metal Complexes. Inorganics 2024, 12, 233. https://doi.org/10.3390/inorganics12090233

AMA Style

Zhao Y, Zhang Z, Luo Y. DFT Modeling of Coordination Polymerization of Polar Olefin Monomers by Molecular Metal Complexes. Inorganics. 2024; 12(9):233. https://doi.org/10.3390/inorganics12090233

Chicago/Turabian Style

Zhao, Yanan, Zhenli Zhang, and Yi Luo. 2024. "DFT Modeling of Coordination Polymerization of Polar Olefin Monomers by Molecular Metal Complexes" Inorganics 12, no. 9: 233. https://doi.org/10.3390/inorganics12090233

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