Next Article in Journal
The Zr Modified γ-Al2O3 Catalysts for Stable Hydrolytic Decomposition of CF4 at Low Temperature
Next Article in Special Issue
Catalysis in Plastics for the 21st Century
Previous Article in Journal
Dry Reforming of Methane on NiCu and NiPd Model Systems: Optimization of Carbon Chemistry
Previous Article in Special Issue
Synthesis of Semicrystalline Long Chain Aliphatic Polyesters by ADMET Copolymerization of Dianhydro-D-glucityl bis(undec-10-enoate) with 1,9-Decadiene and Tandem Hydrogenation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transition Metal-Catalyzed and MAO-Assisted Olefin Polymerization; Cyclic Isomers of Sinn’s Dimer Are Excellent Ligands in Iron Complexes and Great Methylating Reagents

1
Department of Chemistry, University of Missouri, Columbia, MO 65211, USA
2
Department of Chemistry, Missouri University of Science & Technology, Rolla, MO 65409, USA
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(3), 312; https://doi.org/10.3390/catal12030312
Submission received: 11 February 2022 / Revised: 2 March 2022 / Accepted: 6 March 2022 / Published: 9 March 2022
(This article belongs to the Special Issue Catalysis in Plastics for the 21st Century)

Abstract

:
Methylaluminoxane (MAO) is the most commonly used co-catalyst for transition metal-catalyzed olefin polymerization, but the structures of MAO species and their catalytic functions remain topics of intensive study. We are interested in MAO-assisted polymerization with catalysts L(R2)FeCl2 (L = tridentate pyridine-2,6-diyldimethanimine; imine-R = Me, Ph). It is our hypothesis that the MAO species is not merely enabling Fe–Me bond formation but functions as an integral part of the active catalyst, a MAO adduct of the Fe-precatalyst [L(R2)FeCl]+. In this paper, we explored the possible structures of acyclic and cyclic MAO species and their complexation with pre-catalysts [L(R2)FeCl]+ using quantum chemical approaches (MP2 and DFT). We report absolute and relative oxophilicities associated with the Fe ← O(MAO) adduct formation and provide compelling evidence that oxygen of an acyclic MAO species (i.e., O(AlMe2)2, 4) cannot compete with the O-donor in cyclic MAO species (i.e., (MeAlO)2, 7; MeAl(OAlMe2)2, cyclic 5). Significantly, our work demonstrates that intramolecular O → Al dative bonding results in cyclic isomers of MAO species (i.e., cyclic 5) with high oxophilicities. The stabilities of the [L(R2)FeClax(MAO)eq]+ species demonstrate that 5 provides for the ligating benefits of the cyclic MAO species 4 without the thermodynamically costly elimination of TMA. Mechanistic implications are discussed for the involvement of such Fe–O–Al bridged catalyst in olefin polymerization.

Graphical Abstract

1. Introduction

Ziegler-Natta polymerization [1,2] is the dominant method for the production of poly-olefin [3,4]. Ziegler-Natta catalysts usually contain transition metal complexes (i.e., TiCl4) and partially hydrolyzed aluminum alkyls as co-catalyst, among which methylaluminoxane [5,6,7] (MAO) is most commonly used today [8,9,10]. It was a major advance in the field when Brookhart discovered that Fe(II) complexes with di-nitrogen or tri-nitrogen ligands can serve as pre-catalyst with satisfying catalytic activity [11,12,13]. More recently, Sun and co-workers explored structurally similar bidentate bis(imino)pyridyl Fe(II) complexes [14], and tridentate 2,8-bis(imino)quinoline Fe(II) complexes [15], and related systems with nickel [16,17,18], cobalt [19,20], and titanium [21,22] also have been studied. However, these kinds of pre-catalysts require the presence of a very large excess of MAO as co-catalyst (>2000:1) and the structure and the function of MAO have not been clarified at all [23]. The experimental studies of aluminoxanes by Barron [24,25], Pasynkiewicz [26], and Sinn [27] suggested a plethora of chain, ladder, and cage structures. Theoretical studies by Ziegler et al. [28], by Hall and co-workers [29], and by Linnolahti et al. [30,31] explored possible species with ladder and cage structures derived from dimethylaluminum hydroxide (DMAH).
Most of the MAO studies have described very large MAO structures and it was believed that these very large MAO structures also are the active co-catalysts [32,33]. We hold a different view and hypothesize that initially formed MAO species of moderate size are the active co-catalysts and that the large excess required merely reflects the fact that small MAO species tend to aggregate into larger and catalytically less or non-active species. We have published a theoretical study [34] of the formation of varies MAO species (Scheme 1) by partial hydrolysis of trimethylaluminum (TMA) 1. Here we are exploring the possible function of small MAO species as co-catalysts.
In Scheme 2, three mechanisms of catalysis are outlined for a generic LFeCl2 pre-catalyst, where L signifies a tri-nitrogen ligand. All of them share the feature that at least one chloride is abstracted to generate an overall positively charged Fe(II) complex. The abstracted “first” chloride might aggregate with a MAO species and form a large “A” anion which associated with the iron complex, and this scenario is shown in the top row. In this commonly considered mechanism [35], the next step consists in the replacement of the second chloride by a methyl anion (from MAO) forming an Fe–Me bond. After this alkylation step and on entry of an alkene into the vacant coordination site, a nucleophilic addition of the methyl anion to the coordinated alkene results in the replacement of the methyl anion ligand by a propyl anion ligand, and so on. Alternatively, one may consider mechanisms in which the MAO species does not alkylate the transition metal and instead coordinates to the transition metal (Scheme 2, center and bottom) and create a Fe–O–Al bridged catalyst. After their coordination to the transition metal, the MAO species would transfer a methyl group directly from Al to the olefin.
An Al-based alkylation mechanism with Fe–O(MAO) bonding should discriminate strongly between acyclic and cyclic aluminoxanes (Scheme 3). The bridging oxygen in an acyclic aluminoxane may engage both of its p-lone pairs to engage in O → Al dative bonding. This is illustrated by resonance forms 4-IIa, 4-IIb, and 4-III in Scheme 3 for Me2Al–O–AlMe2 (R = Me). The strength of this O → Al dative bonding may be reduced by coordination of other donors to Al, but it will not be eliminated. Therefore, any O → Fe dative bonding in a complex of an acyclic species will always have to compete with two options for O → Al dative bonding. Cyclodialuminoxane 7 exemplifying cyclic MAO species in Scheme 2. Each oxygen in a cyclic aluminoxane can only engage its out-of-plane p-lone pair poop for O → Al dative bonding and, thus, the one in-plane σ-lone pair σip will be available for O → Fe dative bonding. We will substantiate this argument below and it is for that reason that the mechanisms shown in Scheme 1 involve cyclic MAO species. The consideration of cyclodialuminoxane as the active co-catalyst warrants the study of two possible issues. First, the formation of cyclodialuminoxane 7 from the Sinn dimer 5a by TMA elimination might not be able to compete with oligomerization of the Sinn dimer 5a to the Sinn trimer 6. Second, the Fe-coordinated cyclodialuminoxane 7 is only a weak methyl donor.
Here we report on the thermochemistry of the formation of the Sinn dimer 5 from the Sinn monomer 4. Detailed exploration of the potential energy surface of 5 reveals that the acyclic Sinn dimer structure 5a can rearrange easily into cyclic isomers 5b and 5c, which effectively are TMA-complexed cyclodialuminoxane, 7•TMA. We then discuss the ligand binding capacities of acyclic 4, cyclic 7, and cyclic 5 with Fe2+ complexes of type [L(R2)FeCl]+. The organic ligand L is the tridentate pyridine-2,6-diyldimethanimine and we studied the N,N-diphenyl ligand L(Ph)2 and the N,N-dimethyl ligand L(Me)2. The structures of the [L(R2)FeCl(MAO)]+ species show that the 5 is a much better Me-donor than 4, and [L(R2)FeCl(5)]+ illustrates well the mechanism shown in the bottom row of Scheme 2. The stabilities of the [L(R2)FeCl(MAO)]+ species demonstrate that 5 provides for the ligating benefits of the cyclic MAO species 4 without the thermodynamically costly elimination of TMA.

2. Results and Discussion

Isomer preference energies, activation, and reaction energies are discussed, and we report relative energies ΔE, enthalpies, ΔH0 = Δ(E + VZPE) and ΔH298 = Δ(E + TE), and free enthalpies ΔG298 = Δ(E + TE − 298.153·S). Results for the MAO species are collected in Table 1 and all relevant energy of complexation chemistry are collected in Table 2. Computated energies of MAO species and the iron pre-catalyst are shown in Tables S1 and S2. Structures of optimized MAO species are shown in Figures S1–S3. Cartesian coordinates of all the stationary structures are shown in the supporting information.

2.1. Formation of Sinn Dimer 5 and Its Structural Isomerization

The most likely path for the formation of trialuminoxane 5 is outlined in Scheme 1, stationary structures of 5 are shown in Figure 1, and relevant energies are collected in Table 1. Acyclic trialuminoxane 5a can be formed from 4 by adding DMAH and losing methane. The reaction energy ΔG298 = −45.4 kcal/mol for this intermolecular CH4 elimination between DMAH and 4 to form 5a is very similar to the respective value of ΔG298 = −41.2 kcal/mol for the reaction of DMAH and TMA.
The acyclic structure 5a features two near-linear Al–O–Al moieties indicative of a strong O → Al π-dative bonding. Structure 5a can stabilize itself by intramolecular O → Al σ-dative bond formation and resulting in minimum 5b (Figure 1). The cyclization energy of 5 is just about compensated by the loss in cyclization entropy, and 5b is only 0.6 kcal/mol more stable than 5a. The additional O–Al contact (1.91 Å) is only slightly longer than a normal O–Al bond (1.8 Å) and the exocyclic AlMe2 moiety assumes a position to serve as the recipient of methyl-bridging by one CH3 group of the endocyclic AlMe2 group: d(Alendo–CH3) = 2.091 Å, d(Alexo–CH3) = 2.223 Å. Interesting is the fact that there exists a second cyclic minimum 5c that is essentially isoenergetic with 5b. Structure 5c is Cs-symmetric, it contains two bridging methyl groups, the methyl bridges are asymmetric in the opposite direction compared to 5b, that is, d(Alendo–CH3) = 2.256 Å > d(Alexo–CH3) = 2.089 Å.
The stationary structures of 5a, 5b, and 5c are showed in the Figure 1, together with the transition state structure TSRF(5a,5b) for ring formation 5a5b, TSMT(5b,5c) for isomerization 5b5c, and ATS(5c) for automerization of 5c. All the isomerization processes are facile even at low temperature. Hence, we conclude that cyclic 5b is thermodynamically competitive and kinetically accessible.
Structures 5b and 5c can be viewed as the adducts of 7 and TMA. We previously discussed the formation of 7 via dimerization of 3 and the difficulty of generating 3 [32]. In this context, the formation of cyclic structures of 5 can be viewed as an “Umleitung” (German, detour) to a TMA adduct of 7 without going through 7.

2.2. Iron-Complexes: Dichlorides and Monochlorides

Here we report on the complexation of iron(II) complexes derived from complexes of the type L(R2)FeCl2 (Figure 2). The ligand L(R)2 signifies a tridentate organic ligand and pyridine-2,6-diyldimethanimine (R = H) was considered as the parent organic tridentate ligand. Many of the organic ligands employed in Fe(II)/MAO catalyzed olefin polymerizations are N-aryl imines, and we studied the N,N-diphenyl ligand L(Ph)2 and the N,N-dimethyl ligand L(Me)2.
The structures of dichloride complexes are shown in the first column of Figure 2. For L(Me2)FeCl2 there is one C2v symmetric minimum. Two minima exist for complex L(Ph2)FeCl2 (Figure 2) depending as to whether the phenyl twists occur in the same (conrot., C2) or in opposite (disrot., Cs) directions, and these complexes are essentially isoenergetic (Table 2).
It is thought that the dichloride complex is a pre-catalyst and that the dissociation of one chloride generates the cationic complex [L(R2)FeCl]+. The cationic complexes [L(R2)FeCl]+ (Figure 2) essentially retain the coordination geometry between iron and the organic ligand L(R2) as in the structure of the neutral dichloride complexes. There is one Cs-structure [L(Me2)FeCl]+ and two perspectives are shown in columns two and three of Figure 2. Isomers are possible for [L(Ph2)FeCl]+ depending on the conformation of the N–Ph bonds. In the monochloride, there are now two ways to obtain the minima with disrotatory N–Ph twists, I and II, in addition to the structure III resulting from conrotatory N–Ph twists. Structure I is slightly preferred over II and III (Table 2, top).
All structures of [L(R)2FeCl]+ feature chloride in pseudo-axial position and there are no minima in which chlorine would be placed in the best plane of the organic ligand. We found this observation significant and wanted to explore the possible origin. In Figure 3 are displayed the electrostatic potential maps of [L(Me)2FeCl]+ and the red and blue regions coincide with the location of the chlorine and iron, respectively. The iron center is the most electrophilic site by far (dark blue region in Figure 3) and the perfect attractor for olefin monomers. The placement of the chloride ligand in the pseudo-axial position concentrates the iron electrophilicity with the additional advantage of minimizing steric interaction with the incoming olefin.

2.3. Iron-Complexation by Sinn Monomer 4, Cyclic Aluminoxane 7, and Sinn Dimer 5

The iron dichloride complexes contain both chlorides in pseudo-axial positions, that is, the Fe–Cl directions are almost perpendicular to the LFe plane. Dissociation of one chloride leaves the other one in its pseudo-axial position in the [L(R)2FeCl]+ complexes (Figure 2). One might expect that the entry of a MAO ligand would occupy the pseudo-axial vacancy left by the removed chloride. Yet, as will be shown by the optimized structures below, this is not the case in any of [L(R)2FeCl(MAO)]+ complexes. Instead, the MAO species will occupy a pseudo-equatorial position so that it may act as a bidentate ligand engaging in O → Fe bonding and also in Cl → Al bonding. This mode of bidentate MAO coordination is illustrated in Scheme 4.
It is a significant advantage of this MAO coordination mode that the second pseudo-axial position remains vacant and ready for the coordination of an alkene. For the olefin polymerization to proceed by the mechanism outlined in the bottom row of Scheme 2, it is essential that the MAO species is placed next to the olefin, and this requires complexes of the type shown in Scheme 4 with an axial chlorine and an equatorial MAO species.
The desirable [L(R)2FeClax(MAO)eq]+ complexes compete with the isomeric [L(R)2FeCleq(MAO)ax]+ complexes. These isomeric complexes also contain one O → Fe and one Cl → Al contact and one vacancy. However, the vacancy is next to the chloride rather than the MAO species.
We computed both the [L(R)2FeClax(MAO)eq]+ and [L(R)2FeCleq(MAO)ax]+ complexes for the three MAO species 4, 7 and 5b for both the organic ligands L(Ph)2 and L(Me)2. Molecular models of optimized structures of these twelve complexes are shown in Figure 4, Figure 5 and Figure 6 for 4, 7 and 5b, respectively. In each figure, the complexes with the L(Ph)2 ligand are shown in the top row and the complexes in the left column contain equatorial chloride. The energies of the [L(R)2FeCleq(MAO)ax]+ complex relative to the [L(R)2FeClax(MAO)eq]+ complex are included in Table 2, and the energetically preferred isomers are shown with green frames in the figures.
The MAO species 4 and 7 are symmetric and there is only one way for them to enter the coordination sphere of the [L(R)2FeClax]+ complex. On the other hand, the ligand 5 is distinctly asymmetric. The structure of free 5 suggested that 5b and 5c as the best structural isomers with one or two methyl groups in a bridging position (Figure 1). However, in the complexes, the ligand has the topology of 5e (Scheme 1), that is, the exocyclic AlMe2 moiety is almost in the Al2O2 plane with the possibility of O → AlMe2 dative bonding. This topology occurs in all four complexes with 5 (Figure 6). One of the in-ring Al atoms has only one methyl attached and coordinates to chloride (Ala), while the other in-ring Al atom is bonded to two methyl groups (Alb). For complex L(Ph)2FeClax(5)eq, the distances involving Ala are d(Ala–C) = 1.95 Å and d(Ala–Cl) = 2.34 Å, and the d(Alb–C) distances involving Alb are 1.97 Å and 1.98 Å. The regiochemistry of the entry of 5 into the pre-catalyst [L(R)2FeClax]+ is predetermined because chloride binds the AlaMe moiety of the Al2O2 ring. Hence, the AlbMe2 moiety of the Al2O2 ring must be placed in the proximity of the vacancy, which is trans-apical with respect to chloride.
The structure of L(Ph)2FeClax(5)eq exemplifies perfectly the mechanism postulated in the bottom row of Scheme 2, that is, the idea that the MAO species does not alkylate the transition metal and instead coordinates to the transition metal. The regiochemistry of this coordination places the AlMe2 moiety right next to the coordination site of an H2C=CH2 monomer and methyl transfer directly from the Al would result in the formation of an Al(Me)(CH2–CH2–Me) moiety. The catalysis would continue to cycle, converting Al(Me)([CH2–CH2]n–Me) into Al(Me)([CH2–CH2]n+1–Me).

2.4. Kinetic vs. Thermodynamic Control of Complex Formation

We discussed the structure of the MAO complexed transition metal catalysts and our discussions emphasized the advantage of isomer L(R)2FeClax(5)eq for olefin polymerization. In Figure 3, Figure 4 and Figure 5, the more stable isomers are highlighted by a green frame. As can be seen in Figure 4, there is a clear preference for the L(R)2FeClax(4)eq complexes with axial chloride irrespective of the nature of R. For the L(R)2FeCl(7) complexes in Figure 5, the complex with axial chloride is only preferred for the complex with the N,N-diphenyl ligand, L(Ph)2FeClax(4)eq. This result provides guidance for the improvement of the organic ligand. The structure suggests that a cyclic Al2O2 system will tend to align with one N-arene group to benefit from parallel stacking. The interaction of the MAO species with the pre-catalyst will depend on the nature of the groups attached to the imine-Ns and the core of the ligand, which determines the distance between the imines. Inspection of Figure 6 and the isomer stabilities in Table 2 show a preference for the L(R)2FeCleq(5)ax complexes, even for the one with N-phenyl groups.
It is important to distinguish between thermodynamic and kinetic control of isomer formation. In Table 2, we do not only provide thermodynamic isomer stabilities (Reactions (4)–(6) and (13)–(15)), but we also provide the reaction energies for the addition of the MAO species 4, 7, and 5 to the pre-catalyst leading to the isomeric complexes (Reactions (7)–(12) and (16)–(21)). The binding energies of the MAO species to iron are high and suggest irreversible MAO addition to the iron catalyst in fast reactions that will proceed without pronounced regioselectivity and not lead to thermodynamic equilibria. The positive consequence of this situation is that there will always be some active catalyst for the olefin polymerization. The art of catalyst design basically addresses the question as to how to provide a kinetic advantage for the formation of the desired L(R)2FeClax(MAO)eq isomer.

2.5. Absolute and Relative Oxophilicities and MAO Topology

The absolute oxophilicity Ω of the iron pre-catalyst [L(R)2FeCl]+ is defined as the negative of the binding energies of the MAO species which are given in Table 2, and we discuss the ∆G data. For example, the placement of 4, 7, and 5 in the equatorial position of [L(Ph)2FeCl]+ results in oxophilicities of 18.8 (R8), 35.4 (R10), and 35.0 (R12) kcal/mol, respectively. For the N-methyl species [L(Me)2FeCl]+, the respective numbers are 26.1 (R17), 36.5 (R19), and 41.3 (R21) kcal/mol respectively. These oxophilicities reflect the larger steric demand of the N-phenyl groups (decreasing Ω) and the opportunity for π stacking (increasing Ω).
We argued above that the oxygen in a cyclic aluminoxane would be a better donor in O → Fe dative bonding (Scheme 3). The results of the potential energy surface analysis allow to quantify this notion and we examined relative oxophilicities for MAO species in the equatorial position. The relative oxophilicity Ωr(8,10) of the iron pre-catalyst [L(Ph)2FeCl]+ for cyclic 7 relative to acyclic 4 is given by the difference of the binding energies of reactions R8 and R10 in Table 2. In analogy, the relative oxophilicity Ωr(8,12) specifies the preferences for cyclic 5 relative to acyclic 4 based on reactions R8 and R12. The values Ωr(17,19) and Ωr(17,21) are defined in the same way for the N-methyl ligand. The computed oxophilicities Ωr are summarized in Table 3 and we discuss the ΔG values. The difference values ΔΩr(10,12) and ΔΩr(19,21) allow for a comparison of the two cyclic species 7 and 5, and ΔΩr < 0 indicates that 5 is more strongly bonded ligand.
The Ωr values all are greater than 10 kcal/mol, and in strong support of our hypothesis. Both 7 and 5 gave Ωr values of ΔG ≈ 16 kcal/mol for the phenyl-substituted ligands with a small preference ΔΩr(10,12). The respective Ωr values for the methyl-substituted ligands are smaller and vary more; ΔG ≈ 10 kcal/mol for 7 and ΔG ≈ 15 kcal/mol for 5, respectively.

2.6. Mechanistic Hypothesis and Suggested Experimental Approach

The traditional mechanism for MAO-assisted transition metal olefin polymerization considers the MAO species solely as to purveyor of methyl anion to generate an initial Fe–Me species and the turnover of the catalytical cycle keeps adding olefin to generate Fe–alkyl species as shown in Equation (1). We are proposing an alternative mechanism which can be summarized by Equation (2). It is the essential feature of this mechanism that the polymerization occurs at aluminum. The transition metal is directly bonded to aluminum by way of an Fe–O–Al bridge and ensures that the iron-bonded olefin is positioned well for its aluminum-centered polymerization.
(H2C=CH2)LFe([CH2–CH2]n–Me) → LFe([CH2–CH2]n+1–Me)
(H2C=CH2)LFe(Cl)–O–Al([CH2–CH2]n–Me) → LFe(Cl)–O–Al([CH2–CH2]n+1–Me)
It is an essential feature of the proposed mechanism (Equation (2)) that the atoms of one Al–O bond of the MAO species coordinate both to the iron center and a remaining chloride ligand. This feature of the polymerization reaction described by Equation (2) is stressed in the bottom row of Scheme 2. The formation of O → Fe and Cl → Al dative bonds results in a four-membered ring, which should manifest itself in characteristic vibrational signatures. Therefore, we computed both the infrared and Raman intensities of the L(R)2FeClax(X)eq complexes with X = 4, 5b, and 7. In Table S3 we compiled the vibrational modes composed of any combination of Fe–O, Fe–Cl, Al–O, and Al–Cl stretches. The visualization of molecular vibrations is not trivial, but depictions of atom displacement vectors associated with normal modes are beneficial. In Figure S4, we show atom displacement vectors of the three most characteristic vibrational modes of complex L(R)2FeClax(5b)eq. We hope that these vibrational data encourage experimental studies.

3. Computational Methods

Computations were performed with Gaussian09 [36] on the Lewis3 cluster [37] of the high-performance computing center at the University of Missouri. Potential energy surface (PES) analyses of the MAO species were performed with second-order Møller-Plesset perturbation theory (MP2) [38] with all electrons included in the active space and in conjunction with the 6-31G* basis set [39,40]. Vibrational analyses were carried out analytically for each stationary structure at the level of optimization to compute vibrational frequencies and molecular thermal energies, enthalpies, and entropies.
Considering the size of the complexes, the complexes were studied with density functional theory, the popular B3LYP hybrid functional [41], and with the 6-31G* basis set. The six d-electrons of Fe2+ are localized at the Fe center and their specific distribution has little effects on structures and complexation energies. We studied all systems as closed shell singlet using restricted Kohn-Sham orbitals.

4. Conclusions

We have presented the results of computational studies of the association of pre-catalysts [L(Ph)2FeCl]+ with three MAO species. We have shown that the oxygen in cyclic MAO species is a much better ligand in iron complexes compared to the oxygen in acyclic iron species. Importantly, the cyclic MAO species does not have to be a cycloaluminoxane such as [(AlMe)O]2, 7 with all Al–O bonds being localized, but it can be an aluminoxane such as (MeAl)[O(AlMe2)]2, 5, which adopts a cyclic structure because of intramolecular O → Al dative bonding.
It is thought that the pre-catalyst L(R2)FeCl2 goes through dissociation of one chloride and complexation of a MAO species to generate [L(R)2FeCl(MAO)]+ complexes with O → Fe and Cl → Al dative bonding. To enable olefin polymerization, the remaining chloride must be in the pseudo-axial position relative to the Fe-pyridine plane. According to our calculations, the [L(R)2FeClax(MAO)eq]+ complex formations are highly exothermic and exergonic (ΔG values of −19, −35, and −35 kcal/mol for MAO species 4, 7, and 5, respectively), and the regiochemistry of their formation most likely is under kinetic control. Our study suggests that catalyst development should focus on the design of organic ligands (core and N-substituents) that will enhance the propensity for retaining one chloride in the axial position. Computed infrared and Raman spectra are reported and may guide experimental studies.

Supplementary Materials

The Supporting Information is available free of charge at https://www.mdpi.com/article/10.3390/catal12030312/s1, Tables S1 and S2 with the thermodynamic properties of MAO species and iron pre-catalysts, Table S3 with pertinent vibrational data, several figures related to the stationary structures of MAO species, illustrations of vibrational modes, and Cartesian coordinates of MAO species determined at the MP2(full)/6-31G* level and of all iron complexes determined at B3LYP/6-31G* level.

Author Contributions

Conceptualization, K.Y. and R.G.; Data curation, K.Y.; Funding acquisition, R.G.; Supervision, R.G.; Visualization, K.Y.; Writing—original draft, K.Y.; Writing—review & editing, R.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the petroleum Research Fund of the American Chemical Society grant number 53415-ND4, and, in part, by National Science Foundation grant number 1665487.

Data Availability Statement

Included in the Supplementary Materials.

Acknowledgments

This work was supported by the Missouri University of Science and Technology, by a grant from the Petroleum Research Fund of the American Chemical Society (#53415-ND4), and, in part, by a grant from the National Science Foundation #1665487.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Ziegler, K. Folgen und Werdegang einer Erfindung Nobel-Vortrag am 12. Dezember 1963. Angew. Chem. 1964, 76, 545–553. [Google Scholar] [CrossRef]
  2. Natta, G. Von der Stereospezifischen Polymerisation zur Asymmetrischen Autokatalytischen Synthese von Makromolekülen Nobel-Vortrag am 12. Dezember 1963. Angew. Chem. 1964, 76, 553–566. [Google Scholar] [CrossRef]
  3. Nowlin, T.E.; Mink, R.I.; Kissin, Y.V. Supported Magnesium/Titanium-Based Ziegler Catalysts for Production of Polyethylene. In Handbook of Transition Metal Polymerization Catalysts; John Wiley & Sons, Inc.: St. Hoboken, NJ, USA, 2010; pp. 131–155. [Google Scholar]
  4. Severn, J.; Jones, R.L. Stereospecific α-Olefin Polymerization with Heterogeneous Catalysts. In Handbook of Transition Metal Polymerization Catalysts; John Wiley & Sons, Inc.: St. Hoboken, NJ, USA, 2010; pp. 157–230. [Google Scholar]
  5. Sinn, H.; Kaminsky, W. Ziegler-Natta Catalysis. Adv. Organomet. Chem. 1980, 18, 99–149. [Google Scholar]
  6. Andresen, A.; Cordes, H.G.; Herwig, J.; Kaminsky, W.; Merck, A.; Mottweiler, R.; Pein, J.; Sinn, H.; Vollmer, H.J. Halogen-Free Soluble Ziegler Catalysts for the Polymerization of Ethylene. Control of Molecular Weight by Choice of Temperature. Angew. Chem. 1976, 88, 689–690. [Google Scholar] [CrossRef]
  7. Severn, J. Methylaluminoxane (MAO), Silica and a Complex: The “Holy Trinity” of Supported Single-Site Catalyst. In Tailor-Made Polymers: Via Immobilization of Alpha-Olefin Polymerization Catalysts; John Wiley & Sons, Inc.: St. Hoboken, NJ, USA, 2008; pp. 95–138. [Google Scholar]
  8. Zaccaria, F.; Budzelaar, P.; Zuccaccia, C.; Cipullo, R.; Macchioni, A.; Busico, V.; Ehm, C. Chain Transfer to Solvent and Monomer in Early Transition Metal Catalyzed Olefin Polymerization: Mechanisms and Implications for Catalysis. Catalysts 2021, 11, 215. [Google Scholar] [CrossRef]
  9. Kumar, S.; Dholakiya, B.; Jangir, R. Role of Organometallic Complexes in Olefin Polymerization: A Review Report. J. Organomet. Chem. 2021, 953, 122066. [Google Scholar] [CrossRef]
  10. Köppl, A.; Alt, H.H.; Phillips, M.D. Partially Hydrolyzed Trimethylaluminum (PHT) as Heterogeneous Cocatalyst for the Polymerization of Olefins with Metallocene Complexes. J. Appl. Polym. Sci. 2001, 80, 454–466. [Google Scholar] [CrossRef]
  11. Small, B.L.; Brookhart, M.; Bennett, A.M.A. Highly Active Iron and Cobalt Catalysts for the Polymerization of Ethylene. J. Am. Chem. Soc. 1998, 120, 4049–4050. [Google Scholar] [CrossRef]
  12. Small, B.L.; Brookhart, M. Iron-Based Catalysts with Exceptionally High Activities and Selectivities for Oligomerization of Ethylene to Linear α-Olefins. J. Am. Chem. Soc. 1998, 120, 7143–7144. [Google Scholar] [CrossRef]
  13. Ittel, S.D.; Johnson, L.K.; Brookhart, M. Late-Metal Catalysts for Ethylene Homo- and Copolymerization. Chem. Rev. 2000, 100, 1169–1203. [Google Scholar] [CrossRef]
  14. Zhang, T.; Sun, W.H.; Li, T.; Yang, X. Influence of Electronic Effect on Catalytic Activity of Bis(imino)pyridyl Fe(II) and Bis(imino)pyrimidyl Fe(II) Complexes. J. Mol. Catal. A 2004, 218, 119–124. [Google Scholar] [CrossRef]
  15. Zhang, S.; Sun, W.-H.; Xiao, T.; Hao, X. Ferrous and Cobaltous Chlorides Bearing 2,8-Bis(imino)quinolines: Highly Active Catalysts for Ethylene Polymerization at High Temperature. Organometallics 2010, 29, 1168–1173. [Google Scholar] [CrossRef]
  16. Guo, D.; Han, L.; Zhang, T.; Sun, W.H.; Li, T.; Yang, X. Temperature Dependence of the Activity of a Late Transition Metal Catalyst by Molecular Modeling. Macromol. Theory Simul. 2002, 11, 1006–1012. [Google Scholar] [CrossRef]
  17. Zhang, T.; Guo, D.; Jie, S.; Sun, W.H.; Li, T.; Yang, X. Influence of Electronic Effect on Catalytic Activity of Salicylaldiminato Nickel(II) Complexes. J. Polym. Sci. Part A 2004, 42, 4765–4774. [Google Scholar] [CrossRef]
  18. Kong, S.; Guo, C.Y.; Yang, W.; Wang, L.; Sun, W.H.; Glaser, R. 2,6-Dibenzhydryl-N-(2-phenyliminoacenaphthylenylidene)-4-Chloroaniline nickel Dihalides: Synthesis, Characterization and Ethylene Polymerization for Polyethylenes with High Molecular Weights. J. Organomet. Chem. 2013, 725, 37–45. [Google Scholar] [CrossRef]
  19. Lai, J.; Zhao, W.; Yang, W.; Redshaw, C.; Liang, T.; Liu, Y.; Sun, W.H. 2-[1-(2,4-Dibenzhydryl-6-methylphenylimino)ethyl]-6-[1-(arylimino)ethyl]pyridylcobalt(II) Dichlorides: Synthesis, Characterization and Ethylene Polymerization Behavior. Polym. Chem. 2012, 3, 787–793. [Google Scholar] [CrossRef]
  20. Britovsek, G.J.P.; Bruce, M.; Gibson, V.C.; Kimberley, B.S.; Maddox, P.J.; Mastroianni, S.; McTavish, S.J.; Redshaw, C.; Solan, G.A.; Strömberg, S.; et al. Iron and Cobalt Ethylene Polymerization Catalysts Bearing 2,6-Bis(Imino)Pyridyl Ligands:  Synthesis, Structures, and Polymerization Studies. J. Am. Chem. Soc. 1999, 121, 8728–8740. [Google Scholar] [CrossRef]
  21. Zuo, W.; Sun, W.-H.; Zhang, S.; Hao, P.; Shiga, A. Highly Active Ethylene Polymerization and Copolymerization with Norbornene Using Bis(imino-indolide) Titanium Dichloride–MAO System. J. Polym. Sci. Part A 2007, 45, 3415–3430. [Google Scholar] [CrossRef]
  22. Suzuki, Y.; Kinoshita, S.; Shibahara, A.; Ishii, S.; Kawamura, K.; Inoue, Y.; Fujita, T. Trimerization of Ethylene to 1-Hexene with Titanium Complexes Bearing Phenoxy−Imine Ligands with Pendant Donors Combined with MAO. Organometallics 2010, 29, 2394–2396. [Google Scholar] [CrossRef]
  23. Velthoen, M.; Muñoz-Murillo, A.; Bouhmadi, A.; Cecius, M.; Diefenbach, S.; Weckhuysen, B. The Multifaceted Role of Methylaluminoxane in Metallocene-Based Olefin Polymerization Catalysis. Macromolecules 2018, 51, 343–355. [Google Scholar] [CrossRef]
  24. Barron, A.R. A New Understanding of the Co-Catalytic Activity of Alumoxanes: The Opening of a Black Box! Macromol. Symp. 1995, 97, 15–25. [Google Scholar] [CrossRef]
  25. Watanabi, M.; McMahon, C.N.; Harlan, C.J.; Barron, A.R. Reaction of Trimethylaluminum with [(tBu)Al(μ3-O)]6:  Hybrid tert-Butylmethylalumoxanes as Cocatalysts for Olefin Polymerization. Organometallics 2001, 20, 460–467. [Google Scholar] [CrossRef]
  26. Pasynkiewicz, S. Alumoxanes: Synthesis, Structures, Complexes and Reactions. Polyhedron 1990, 9, 429–453. [Google Scholar] [CrossRef]
  27. Sinn, H.J. Proposals for Structure and Effect of Methylalumoxane Based on Mass Balances and Phase Separation Experiments. Macromol. Symp. 1995, 97, 27–52. [Google Scholar] [CrossRef]
  28. Zurek, E.; Woo, T.K.; Firman, T.K.; Ziegler, T. Modeling the Dynamic Equilibrium between Oligomers of (AlOCH3)n in Methylaluminoxane (MAO). A Theoretical Study Based on a Combined Quantum Mechanical and Statistical Mechanical Approach. Inorg. Chem. 2001, 40, 361–370. [Google Scholar] [CrossRef]
  29. Negureanu, L.; Hall, R.W.; Butler, L.G.; Simeral, L.A. Methyaluminoxane (MAO) Polymerization Mechanism and Kinetic Model from Ab Initio Molecular Dynamics and Electronic Structure Calculations. J. Am. Chem. Soc. 2006, 128, 16816–16826. [Google Scholar] [CrossRef]
  30. Linnolahti, M.; Severn, J.R.; Pakkanen, T.A. Formation of Nanotubular Methylaluminoxanes and the Nature of the Active Species in Single-Site α-Olefin Polymerization Catalysis. Angew. Chem. Int. Ed. 2008, 47, 9279–9283. [Google Scholar] [CrossRef]
  31. Linnolahti, M.; Severn, J.R.; Pakkanen, T.A. Are Aluminoxanes Nanotubular? Structural Evidence from a Quantum Chemical Study. Angew. Chem. Int. Ed. 2006, 45, 3331–3334. [Google Scholar] [CrossRef]
  32. Falls, Z.; Tyminska, N.; Zurek, E. The Dynamic Equilibrium Between (AlOMe)n Cages and (AlOMe)n·(AlMe3)m Nanotubes in Methylaluminoxane (MAO): A First-Principles Investigation. Macromolecules 2014, 47, 8556–8569. [Google Scholar] [CrossRef]
  33. Tyumkina, T.V.; Islamov, D.N.; Kovyazin, R.V.; Parfenovam, L.V. Chain and Cluster Models of Methylaluminoxane as Activators of Zirconocene Hydride, Alkyl, and Metallacyclopropane Intermediates in Alkene Transformation. Mol. Catal. 2021, 512, 111768. [Google Scholar] [CrossRef]
  34. Glaser, R.; Sun, X. Thermochemistry of the Initial Steps of Methylaluminoxane Formation. Aluminoxanes and Cycloaluminoxanes by Methane Elimination from Dimethylaluminum Hydroxide and Its Dimeric Aggregates. J. Am. Chem. Soc. 2011, 133, 13323–13336. [Google Scholar] [CrossRef] [PubMed]
  35. Zijlstra, H.S.; Harder, S. Methylalumoxane-History, Production, Properties, and Applications. Eur. J. Inorg. Chem. 2015, 19–43. [Google Scholar] [CrossRef]
  36. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09. 2016. Available online: https://gaussian.com/g09citation (accessed on 1 March 2022).
  37. RCSS Documentation. Available online: http://docs.rnet.missouri.edu/lewis-and-clark-clusters (accessed on 1 March 2022).
  38. Møller, C.; Plesset, M.S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46, 0618–0622. [Google Scholar] [CrossRef] [Green Version]
  39. Petersson, G.A.; Bennett, A.; Tensfeldt, T.G.; Al-Laham, M.A.; Shirley, W.A.; Mantzaris, J. A Complete Basis Set Model Chemistry. I. The Total Energies of Closed-Shell Atoms and Hydrides of the First-Row Atoms. J. Chem. Phys. 1988, 89, 2193–2218. [Google Scholar] [CrossRef]
  40. Petersson, G.A.; Al-Laham, M.A. A Complete Basis Set Model Chemistry. II. Open-shell Systems and the Total Energies of the First-Row Atoms. J. Chem. Phys. 1991, 94, 6081–6090. [Google Scholar] [CrossRef]
  41. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Possible Products Formed by Partial Hydrolysis of Trimethylaluminum 1. Methyl-Transfer Isomerization of 5a Provides a Path to Cyclic 5 and the TMA Adduct of 7 without Going Through 3.
Scheme 1. Possible Products Formed by Partial Hydrolysis of Trimethylaluminum 1. Methyl-Transfer Isomerization of 5a Provides a Path to Cyclic 5 and the TMA Adduct of 7 without Going Through 3.
Catalysts 12 00312 sch001
Scheme 2. Mechanisms of Different MAO Co-Catalysts Catalyzing Olefin Polymerization.
Scheme 2. Mechanisms of Different MAO Co-Catalysts Catalyzing Olefin Polymerization.
Catalysts 12 00312 sch002
Scheme 3. Differences in Intramolecular O → Al σ-Dative Bonding in Acyclic and Cyclic Aluminoxanes Affect Their Capacity to Serve as Lewis Acceptors and Lewis Donors in Intermolecular Interactions.
Scheme 3. Differences in Intramolecular O → Al σ-Dative Bonding in Acyclic and Cyclic Aluminoxanes Affect Their Capacity to Serve as Lewis Acceptors and Lewis Donors in Intermolecular Interactions.
Catalysts 12 00312 sch003
Scheme 4. The MAO Species Engages the Fe–Cl Bond as a Bidentate Donor-Acceptor Ligand.
Scheme 4. The MAO Species Engages the Fe–Cl Bond as a Bidentate Donor-Acceptor Ligand.
Catalysts 12 00312 sch004
Figure 1. MP2/6-31G* optimized structures of permethyltrialuminoxane 5: Acyclic allene structure 5a, cyclic, monobridged structure 5b, and cyclic, dibridged structure 5c. Also shown are the computed structures of the transition state structures TSRF(5a,5b) for ring formation, of the transition state structure TSMT(5b,5c) for methyl transfer between 5b and 5c, and of the transition state structure ATS(5c) for automerization 5c5c.
Figure 1. MP2/6-31G* optimized structures of permethyltrialuminoxane 5: Acyclic allene structure 5a, cyclic, monobridged structure 5b, and cyclic, dibridged structure 5c. Also shown are the computed structures of the transition state structures TSRF(5a,5b) for ring formation, of the transition state structure TSMT(5b,5c) for methyl transfer between 5b and 5c, and of the transition state structure ATS(5c) for automerization 5c5c.
Catalysts 12 00312 g001
Figure 2. Iron dichloride pre-catalysts L(R)2FeCl2 are shown in the column on the left and two perspectives of the iron monochloride pre-catalysts [L(R)2FeCl]+ are shown in the other two columns.
Figure 2. Iron dichloride pre-catalysts L(R)2FeCl2 are shown in the column on the left and two perspectives of the iron monochloride pre-catalysts [L(R)2FeCl]+ are shown in the other two columns.
Catalysts 12 00312 g002
Figure 3. Electrostatic potential mapped on the electron density isosurface at ρ = 0.004 e/a.u.3 for the N,N-dimethyl system [L(Me)2FeCl]+. The surface map on the left shows the ligand face from the side of the axial chloride.
Figure 3. Electrostatic potential mapped on the electron density isosurface at ρ = 0.004 e/a.u.3 for the N,N-dimethyl system [L(Me)2FeCl]+. The surface map on the left shows the ligand face from the side of the axial chloride.
Catalysts 12 00312 g003
Figure 4. Computed structure of catalyst with the Sinn monomer 4, O(AlMe2)2 a small acyclic aluminoxane. Top row: N,N-diphenyl ligand L; bottom row: N,N-dimethyl ligand L. Left column: equatorial chloride; right column: axial chloride.
Figure 4. Computed structure of catalyst with the Sinn monomer 4, O(AlMe2)2 a small acyclic aluminoxane. Top row: N,N-diphenyl ligand L; bottom row: N,N-dimethyl ligand L. Left column: equatorial chloride; right column: axial chloride.
Catalysts 12 00312 g004
Figure 5. Computed structure of catalyst with 7, (OAlMe2)2, a small cyclic aluminoxane. Top row: N,N-diphenyl ligand L; bottom row: N,N-dimethyl ligand L. Left column: equatorial chloride; right column: axial chloride.
Figure 5. Computed structure of catalyst with 7, (OAlMe2)2, a small cyclic aluminoxane. Top row: N,N-diphenyl ligand L; bottom row: N,N-dimethyl ligand L. Left column: equatorial chloride; right column: axial chloride.
Catalysts 12 00312 g005
Figure 6. Computed structure of catalyst with 5, the cyclic isomer of the Sinn trimer. Top row: N,N-diphenyl ligand L; bottom row: N,N-dimethyl ligand L. Left column: equatorial chloride; right column: axial chloride.
Figure 6. Computed structure of catalyst with 5, the cyclic isomer of the Sinn trimer. Top row: N,N-diphenyl ligand L; bottom row: N,N-dimethyl ligand L. Left column: equatorial chloride; right column: axial chloride.
Catalysts 12 00312 g006
Table 1. Computed Relative and Reaction Energies of Formation of Permethyltrialuminoxane 5 and Isomerization of 5 (kcal/mol).
Table 1. Computed Relative and Reaction Energies of Formation of Permethyltrialuminoxane 5 and Isomerization of 5 (kcal/mol).
Reaction/ProcessEΔH0ΔH298ΔG298
5a5b (Cyclization)−9.86−8.73−10.25−0.64
5b5c−0.95−0.67−1.040.78
5a5c (Cyclization)−10.82−9.40−11.290.14
Eact, TSRF(5a,5b) vs. 5a1.571.670.805.63
Eact, TSMT(5b,5c) vs. 5b0.120.06−0.561.57
Eact, ATS(5c) vs. 5c21.5820.3920.9316.74
DMAH (2) + TMA (1) → 4 + CH4−49.14−48.20−47.63−46.16
4 + 25a + CH4−48.05−47.26−46.64−45.38
4 + 25c + CH4−58.87−56.66−57.93−45.24
4 + MeAlO (3) → 5c−98.87−96.27−97.03−79.69
4 + 0.5 (MeAlO)2 (7) → 5c−32.22−30.78−31.58−18.71
Table 2. Computed Relative and Reaction Energies (kcal/mol, B3LYP/6-31G *).
Table 2. Computed Relative and Reaction Energies (kcal/mol, B3LYP/6-31G *).
Rxn.Reaction/ProcessΔEΔH0ΔH298ΔG298
N-phenyl ligands L(Ph)2
1L(Ph)2FeCl2, C2 → Cs0.020.030.03−0.18
2[L(Ph)2FeCl]+, I → II3.580.430.470.62
3[L(Ph)2FeCl]+, I → III0.590.430.470.62
4[L(Ph)2FeClax(4)eq]+ → [L(Ph)2FeCleq(4)ax]+4.154.064.143.84
5[L(Ph)2FeClax(7)eq]+ → [L(Ph)2FeCleq(7)ax]+3.833.653.822.82
6[L(Ph)2FeClax(5)eq]+ → [L(Ph)2FeCleq(5)ax]+−4.66−4.72−4.74−4.57
7[L(Ph)2FeCl]+ + 4 → [L(Ph)2FeCleq(4)ax]+−34.60−33.00−32.68−14.96
8[L(Ph)2FeCl]+ + 4 → [L(Ph)2FeClax(4)eq]+−38.75−37.06−36.82−18.80
9[L(Ph)2FeCl]+ + 7 → [L(Ph)2FeCleq(7)ax]+−50.29−49.02−48.49−32.56
10[L(Ph)2FeCl]+ + 7 → [L(Ph)2FeClax(7)eq]+−54.12−52.67−52.31−35.38
11[L(Ph)2FeCl]+ + 5 → [L(Ph)2FeCleq(5)ax]+−54.22−53.40−52.29−39.55
12[L(Ph)2FeCl]+ + 5 → [L(Ph)2FeClax(5)eq]+−49.56−48.69−47.55−34.97
N-methyl ligands L(Me)2
13[L(Me)2FeClax(4)eq]+ → [L(Me)2FeCleq(4)ax]+5.064.624.904.62
14[L(Me)2FeClax(7)eq]+ → [L(Me)2FeCleq(7)ax]+−0.33−0.62−0.43−1.73
15[L(Me)2FeClax(5)eq]+ → [L(Me)2FeCleq(5)ax]+−3.10−3.39−3.29−3.41
16[L(Me)2FeCl]+ + 4 → [L(Me)2FeCleq(4)ax]+−40.51−38.87−38.57−21.44
17[L(Me)2FeCl]+ + 4 → [L(Me)2FeClax(4)eq]+−45.57−43.49−43.47−26.06
18[L(Me)2FeCl]+ + 7 → [L(Me)2FeCleq(7)ax]+−54.33−53.17−52.53−38.27
19[L(Me)2FeCl]+ + 7 → [L(Me)2FeClax(7)eq]+−53.99−52.54−52.09−36.53
20[L(Me)2FeCl]+ + 5 → [L(Me)2FeCleq(5)ax]+−58.65−57.83−56.69−44.70
21[L(Me)2FeCl]+ + 5 → [L(Me)2FeClax(5)eq]+−55.55−54.44−53.41−41.28
Table 3. Cyclic MAO are much Better Ligands Compared to Acyclic MAO Species (kcal/mol, B3LYP/6-31G*).
Table 3. Cyclic MAO are much Better Ligands Compared to Acyclic MAO Species (kcal/mol, B3LYP/6-31G*).
Oxophilicity AdvantageΔEΔH0ΔH298ΔG298
N-phenyl ligands L(Ph)2
Difference Ωr(8,10)15.3615.6015.4816.58
Difference Ωr(8,12)10.8111.6210.7216.17
Difference ΔΩr(10,12)−4.55−3.98−4.76−0.41
N-methyl ligands L(Me)2
Difference Ωr(17,19)8.429.058.6210.47
Difference Ωr(17,21)9.9810.969.9415.22
Difference ΔΩr(19,21)1.561.901.314.75
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, K.; Glaser, R. Transition Metal-Catalyzed and MAO-Assisted Olefin Polymerization; Cyclic Isomers of Sinn’s Dimer Are Excellent Ligands in Iron Complexes and Great Methylating Reagents. Catalysts 2022, 12, 312. https://doi.org/10.3390/catal12030312

AMA Style

Yang K, Glaser R. Transition Metal-Catalyzed and MAO-Assisted Olefin Polymerization; Cyclic Isomers of Sinn’s Dimer Are Excellent Ligands in Iron Complexes and Great Methylating Reagents. Catalysts. 2022; 12(3):312. https://doi.org/10.3390/catal12030312

Chicago/Turabian Style

Yang, Kaidi, and Rainer Glaser. 2022. "Transition Metal-Catalyzed and MAO-Assisted Olefin Polymerization; Cyclic Isomers of Sinn’s Dimer Are Excellent Ligands in Iron Complexes and Great Methylating Reagents" Catalysts 12, no. 3: 312. https://doi.org/10.3390/catal12030312

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