Next Article in Journal
Red Blood Cell Stiffness and Adhesion Are Species-Specific Properties Strongly Affected by Temperature and Medium Changes in Single Cell Force Spectroscopy
Next Article in Special Issue
The Reaction of Hydrogen Halides with Tetrahydroborate Anion and Hexahydro-closo-hexaborate Dianion
Previous Article in Journal
Insights into the Behavior of Triple-Negative MDA-MB-231 Breast Carcinoma Cells Following the Treatment with 17β-Ethinylestradiol and Levonorgestrel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ti Group Metallocene-Catalyzed Synthesis of 1-Hexene Dimers and Tetramers

by
Pavel V. Kovyazin
,
Almira Kh. Bikmeeva
,
Denis N. Islamov
,
Vasiliy M. Yanybin
,
Tatyana V. Tyumkina
and
Lyudmila V. Parfenova
*
Institute of Petrochemistry and Catalysis of Russian Academy of Sciences, Prospekt Oktyabrya, 141, 450075 Ufa, Russia
*
Author to whom correspondence should be addressed.
Submission received: 10 April 2021 / Revised: 2 May 2021 / Accepted: 4 May 2021 / Published: 8 May 2021

Abstract

:
1-Hexene transformations in the catalytic systems L2MCl2–XAlBui2 (L = Cp, M = Ti, Zr, Hf; L = Ind, rac-H4C2[THInd]2, M = Zr; X = H, Bu i) and [Cp2ZrH2]2-ClAlR2 activated by MMAO-12, B(C6F5)3, or (Ph3C)[B(C6F5)4] in chlorinated solvents (CH2Cl2, CHCl3, o-Cl2C6H4, ClCH2CH2Cl) were studied. The systems [Cp2ZrH2]2-MMAO-12, [Cp2ZrH2]2-ClAlBui2-MMAO-12, or Cp2ZrCl2-HAlBui2-MMAO-12 (B(C6F5)3) in CH2Cl2 showed the highest activity and selectivity towards the formation of vinylidene head-to-tail alkene dimers. The use of chloroform as a solvent provides further in situ dimer dimerization to give a tetramer yield of up to 89%. A study of the reaction of [Cp2ZrH2]2 or Cp2ZrCl2 with organoaluminum compounds and MMAO-12 by NMR spectroscopy confirmed the formation of Zr,Zr-hydride clusters as key intermediates of the alkene dimerization. The probable structure of the Zr,Zr-hydride clusters and ways of their generation in the catalytic systems were analyzed using a quantum chemical approach (DFT).

Graphical Abstract

1. Introduction

Alkene dimers and oligomers represent a large class of compounds that are used as comonomers in ethylene polymerization and as raw materials for the production of adhesives, surfactants, fragrances, synthetic lubricating fuel additives, etc. [1,2,3,4,5,6]. The following industrial processes have been successfully developed and implemented for the production of olefin oligomers: (i) oligomerization of ethylene in the presence of triethylaluminum with subsequent oxidation to higher alcohols (Ziegler–Alfol process), (ii) Philips process of ethylene oligo—and polymerization on chromium catalysts, (iii) Ni-catalyzed synthesis of linear α-olefins via ethylene oligomerization (shell higher olefin process (SHOP)), (iv) oligomerization of ethylene to linear C4–C10 olefins by α-select technology (Axens) or α-SABLIN technology (Sabic and Linde) [7]. Currently, some research groups are developing strategies for the production of highly efficient jet and diesel fuels via oligomerization of alkenes (1-butene, 1-hexene) synthesized from renewable plant raw materials [2,8].
Among the catalysts, Ti group metallocene complexes showed high efficiency in alkene di-, oligo-, and polymerization [9,10,11,12]. In this way, the process of zirconocene-catalyzed dimerization of terminal alkenes to give products with a vinylidene moiety >C=CH2 is well-known (Scheme 1) [5,6,13,14,15,16,17,18,19,20].
Despite significant progress in this area, the mechanism of the dimerization reaction is not fully understood. Earlier, the hypothesis on the key role of the hydride cation [Cp2ZrH]+ was put forward [14]. Subsequently, it was postulated that bimetallic hydride complexes L2Zr(µ-H)3AlR2 can act as catalytically active species in the alkene dimerization [21,22]. Recently, we studied the systems Cp2ZrCl2-XAlBui2 (X = H, Bui) and [Cp2ZrH2]2-ClAlR2 (R = Me, Et, Bui) activated by MMAO-12 (modified methylaluminoxane) or organoboron compounds and found new biszirconium hydride intermediates (Cp2ZrH2.Cp2ZrHCl.ClAlR2), which provide selective formation of head-to-tail dimerization products [19,20,23]. However, the questions on the possible structure of the key intermediates and their formation in the catalytic systems remain open.
Thus, our work was aimed at investigating the effect of transition metal η5-complexes and solvents (chlorine-containing solvents) on the activity, chemo- and regioselectivity of the systems L2MCl2–XAlBui2 (L = Cp, M = Ti, Zr, Hf; L = Ind, rac-H4C2[THInd]2, M = Zr; X = H, Bui) and [Cp2ZrH2]2-ClAlR2 activated by MMAO-12, B(C6F5)3, or (Ph3C)[B(C6F5)4] in alkene oligomerization. NMR studies and quantum chemical calculations were applied to establish the possible structure of the hydride intermediates formed in these systems. The results showed that the M,M-type bimetallic hydrides and their action mechanisms are unique for the dimerization reaction.

2. Results and Discussion

2.1. Transformations of 1-Hexene with Cp2MY2 (M = Ti, Zr, Hf; Y = H, Cl)-XAlBui2 (X = H, Bui) Catalytic Systems Activated by MMAO-12, B(C6F5)3, or (Ph3C)[B(C6F5)4]

The conditions for the selective synthesis of vinylidene dimers in the presence of catalytic systems Cp2ZrY2 (Y = H, Cl)-XAlBui2 (X = H, Bui)-activator (Scheme 2, catalytic systems A and B), developed in our previous study, were taken as the starting point for the experiment (Table 1, entry 1) [19,20].
The reaction proceeds in toluene or benzene at 20–60 °C to give the target products in 81–98% yields within 5 to 150 min, depending on the type of activator (Table 1, entries 1, 15, 21, 27, and 43). The use of chlorine-containing solvents in these systems was found to accelerate the reaction. For example, the system [Cp2ZrH2]2-ClAlBui2-MMAO-12 with a molar ratio [Zr]:[Al]:[MMAO‑12]:[1-hexene] = 1:3:30:400 at a temperature of 40 °C in CH2Cl2 gives 1-hexene dimer in a yield of more than 98% by the 15th minute of the reaction (Table 1, entry 2). When the reaction is carried out in chloroform, the relative content of dimers decreases to 92% (entry 4). Under the same conditions, the dimers produced in the first minutes of the reaction serve as substrates for the subsequent dimerization to afford dimers of dimer 7 in 75–79% yields in 180 min (entries 7, 8, Figures S26 and S27). According to 1H and 13C NMR data, the structures of these products correspond to those proposed earlier [16,22,24,25,26]. Dimerization of 5 occurs probably via participation of cationic species formed upon the reaction of CHCl3 with ClAlR2 or MMAO-12.
The replacement of MMAO-12 by B(C6F5)3 in the system [Cp2ZrH2]2-ClAlBui2-activator leads to a significant reduction in substrate conversion. The reaction occurs only in chloroform, and the conversion of 1-hexene in 16 h is 75% with a 71% content of dimers in the product mixture (entries 18, 19). Unlike B(C6F5)3, compound (Ph3C)[B(C6F5)4] can activate the system [Cp2ZrH2]2-ClAlBui2. Under these conditions, the conversion of 1-hexene in 3 h was higher in CH2Cl2 (99%, entry 26), and the reaction predominantly gave tetramer 7 in 69% yield. In contrast, in chloroform, the 1-hexene conversion was 81% in 16 h, and oligomeric products 6, including compound 7, prevailed in the product mixture (entries 24, 25).
The use of chlorine-containing solvents CH2Cl2 and CHCl3 in the case of the system [Cp2ZrH2]2-MMAO-12 at the ratio [Zr]:[MMAO-12]:[1-hexene] = 1:30:400 allows the reaction to be carried out in the absence of ClAlR2 with a >99% conversion of 1-hexene and up to 89–91% yield of dimer 5 (entries 9, 10). In o-dichlorobenzene and 1,2-dichloroethane, the reaction proceeds to a lower substrate conversion and a lower yield of the dimeric product (entries 11–14). The composition of the reaction products formed in the system [Cp2ZrH2]2-MMAO-12-1-hexene for 16 h in any of the solvents (CH2Cl2, CHCl3, o-Cl2C6H4, ClCH2CH2Cl) did not change significantly. Attempts to activate [Cp2ZrH2]2 with boron-containing activators in the absence of ClAlR2 were unsuccessful (entries 16, 17, 22, 23).
Thus, vinylidene dimer 5 is formed in a high yield and with high selectivity under the action of [Cp2ZrH2]2-ClAlBui2-MMAO-12 or [Cp2ZrH2]2-MMAO-12 in CH2Cl2. Chloroform facilitates the formation of a non-classical tetramer of 1-hexene (7), which could be used as a component of lubricants [16,24,25,26].
The catalytic system based on Cp2ZrCl2, HAlBui2, and MMAO-12 also provides vinylidene dimers 5 in chlorinated solvents. In CH2Cl2 and CHCl3, the reaction proceeds in 30 min with the conversion of 1-hexene above 99% and the yield of dimers of 98% (entries 28, 34). Increasing 1-hexene initial concentration to 1000 equivalents leads to decreased alkene conversion to 82% and the yield of dimeric product to 80% in CH2Cl2. Under these conditions, the conversion and the yield of dimers are higher in CHCl3 than in CH2Cl2 and amount to >99% and 90–98% in 30 min, respectively (entries 32, 35).
The reaction carried out in the presence of Cp2ZrCl2 and MMAO-12 at the ratio [Zr]:[MMAO-12]:[1-hexene] = 1:30:400 in chlorinated solvents (CH2Cl2, o-Cl2C6H4, ClCH2CH2Cl) at 40 °C results in 91–96% yield of dimers (entries 30, 31, 39–42). Carrying out the reaction in chloroform gives dimers of dimer 7 in a yield of up to 89% (entry 37). A decrease in the relative amount of MMAO-12 in the system to [Zr]:[MMAO-12]:[1-hexene] = 1:10:400 provides the selective formation of dimerization products at the level of 91% within 16 h.
The highest yield of dimer 5 at the level of 99% is achieved by using a neutral boron-containing activator B(C6F5)3 in the system Cp2ZrCl2-HAlBui2 at the ratio [Zr]:[Al]:[B(C6F5)3]:[1-hexene] = 4:16:1:1000 and in dichloromethane as a solvent (entry 44).
Activation of system Cp2ZrCl2-HAlBui2 by (Ph3C)[B(C6F5)4] at a ratio [Zr]:[Al]:[(Ph3C)[B(C6F5)4] ]:[1-hexene] = 4:16:1:1000 and a temperature 40 °C affords alkene oligomers in both CH2Cl2 and CHCl3 (entries 50–52, 54, 55). Selective formation of a dimer (92%, entry 49) occurs in CH2Cl2 at 20 °C, with the reaction time being up to 180 min.
Subsequently, the content of dimers decreases to 55%, and the amount of tetramer 7 increases to 33% (entry 50). Using chloroform as a solvent raises the yield of 7 to 65–72% (entries 54, 55).
The activity of Cp2TiCl2 (1a) and Cp2HfCl2 (1c) in the studied systems was lower than that of Cp2ZrCl2 (1b). Indeed, the conversion of 1-hexene in CH2Cl2 at 40 °C was 80% in 60 min of the reaction in the case of the Ti complex (Table 2, entries 1) and 84% in 120 min for the Hf complex (entries 5). The use of chloroform as a solvent increases substrate conversion to 93% for Cp2TiCl2 and decreases the conversion to 60% for Cp2HfCl2 (entries 3,8). Moreover, these catalysts are significantly inferior in chemo- and regioselectivity to zirconocene dichloride. Thus, in the presence of Cp2TiCl2, the yield of oligomeric products 6 and tetramer 7, which represent a mixture of regioisomers, increases in total to 89%. The use of Cp2HfCl2 in CH2Cl2 leads to increased the content of vinylidene oligomers up to 28–37%. In chloroform, the yield of regioisomeric oligomers also increases in the first 60 min of the reaction (entry 8). Further, the appearance of products of toluene mono-, di-, and trialkylation with 1-hexene was found (entries 9,10). It should be noted that in the presence of HAlBui2, upon replacing MMAO-12 with boron-containing activators, the systems based on Cp2TiCl2 and Cp2HfCl2 completely lost their activity (Table 2, entries 11–14).
The effect of the ligand structure of zirconocenes on the activity of catalytic systems and the reaction route was also studied. For this purpose, we tested Zr η5-complexes (C5Me5)2ZrCl2 (1d), Ind2ZrCl2 (1e), and rac-H4C2(THInd)2ZrCl2 (1f) (Scheme 2) as components of the systems L2ZrCl2-HAlBui2-MMAO-12 at molar ratio [Zr]:[Al]:[MMAO-12]:[1-hexene] = 1:3:30:400 and temperature of 40 °C. It turned out that (C5Me5)2ZrCl2 does not lead to the formation of the target products either in CH2Cl2 or in CHCl3 (Table 3, entries 1,2). In this case, 1-hexene was consumed for toluene alkylation (toluene is present in the system as a solvent for MMAO-12). The complex Ind2ZrCl2 showed lower activity than Cp2ZrCl2 (Table 3, entries 3–7) and shifted the reaction route toward the oligomerization (up to 87%) both in CH2Cl2 and CHCl3. Ansa-zirconocene rac-H4C2(THInd)2ZrCl2 was inactive in CHCl3 (entry 9); however, it catalyzed alkene oligomerization in CH2Cl2 (1-hexene conversion of >99%, entry 8). Thus, the zirconocene ligand structure significantly affects the course of the reaction, and some tested complexes mainly provided the formation of oligomeric products.

2.2. NMR Study of Intermediate Structure in MMAO-12-Activated Systems Cp2ZrY2 (Y = H, Cl)-OAC in Chlorinated Solvents

The effect of chlorinated solvents (CD2Cl2 and CDCl3) on the structure of the intermediates responsible for the alkene dimerization in the catalytic systems [Cp2ZrH2]2-MMAO-12 was studied by NMR. The addition of MMAO-12 to a solution of [Cp2ZrH2]2 in CD2Cl2 gives rise to a triplet at −5.95 ppm (J = 17.0 Hz) in the 1H NMR spectra (Figure 1c, Figures S21–S23). In the COSY HH spectrum, the triplet correlates with the signal at −0.47 ppm, which is superimposed with the region of resonance lines of methyl group protons of MMAO-12 (−0.82–0.32 ppm). The spectra also exhibited signals for the Cp rings at 6.16 ppm (108.4 ppm in 13C NMR spectra), 6.39 ppm (112.8 ppm in 13C NMR spectra), and 6.61 ppm (116.2 ppm in 13C NMR spectra). The NOESY spectra showed the cross-peaks of the Cp-ring signal at 6.16 ppm with the upfield signals at −5.95 ppm and −0.47 ppm. Relying on these results and the data from previous studies [19,20], these signals were assigned to the biszirconium trihydride complex 9. The signals at 6.61 ppm and 6.39 ppm correspond to the Cp rings of Cp2ZrCl2 and Cp2ZrMeCl (11), respectively [27].
As shown in Scheme 3, the system [Cp2ZrH2]2-MMAO-12 in CD2Cl2 produces complex 9 via in situ formation of Cp2ZrHCl and Cp2ZrCl2 by the reaction of zirconocene dihydride with a chlorine-containing solvent [28]. Residual AlMe3 present in the MMAO-12 solution reacts with Cp2ZrCl2 to give methyl chloride complex 11 and ClAlMe2. The reaction of Cp2ZrHCl with the starting Cp2ZrH2 and ClAlMe2 makes it possible to selectively obtain complex 9 and then high-molecular-weight associates with MMAO-12 (9·MAO), which are active in the alkene dimerization [19,20].
The 1H and 13C NMR spectra of the system [Cp2ZrH2]2-ClAlBui2 (Figure 1a, Figures S14 and S15) in CD2Cl2 exhibited signals of previously described intermediates [19,23,28,29]: complex 8 (broadened signals of hydride atoms at δH −0.86 and −1.91 ppm), dimeric complex 10 (broadened signals of hydride atoms at δH −1.42 and −2.38 ppm), and biszirconium trihydride complex 9 (doublet signal at δH −0.70 ppm (2J = 17.0 Hz) and triplet at −5.87 ppm). The addition of MMAO-12 to the equilibrium mixture of the complexes leads to the vanishing of signals of complex 10 from the 1H NMR spectrum and the appearance of an additional triplet at −6.01 ppm attributable to the adduct 9·MAO (Figure 1b, Figures S16–S18).
A similar picture is observed in the 1 H and 13C NMR spectra of Cp2ZrCl2-HAlBui2 in CD2Cl2 (Figures S5 and S6). When MMAO-12 is added to the system, the signals of the dimeric complex 10 also disappear, the intensity of the signals of the trihydride 8 decreases, and additional triplets of 9·MAO appear in the upfield region at −6.21 ÷ −5.92 ppm (Figures S7 and S8).
The reaction of [Cp2ZrH2]2 with MMAO-12 in either CDCl3 or CD2Cl2 gives complexes Cp2ZrCl2Cp 6.76 ppm) and Cp2ZrMeCl (δCp 6.58 ppm) (Figure 2c, Figures S19 and S20). Moreover, the 1 H NMR spectrum showed a significant broadening of the signals of the hydride atom at −6.12 ppm and Cp rings of the heavy adduct 9·MAO, which disappeared from the spectra due to the precipitation of the heavy fraction to the bottom of the NMR tube.
In the reaction of Cp2ZrCl2 with HAlBui2 (1:4) in CDCl3, only the Zr,Al-trihydride complex 8 was identified (Figure 2a, Figures S1 and S2). The signals of complexes 9 and 9·MAO appeared in the 1 H NMR spectra after the addition of MMAO-12 and 1-hexene to the system (Figure 2b, Figures S3 and S4). The formation of complexes 810 was observed in the reaction of zirconocene dihydride with ClAlBui2 taken in a 1:2 ratio in CDCl3 (Figure S9). The addition of MMAO-12 to this system results in the disappearance of signals of complex 10, broadening of complex 8 signals, and the appearance of an additional broadened multiplet at −6.03 ppm, corresponding to the heavy adduct 9·MAO (Figure S11). NMR monitoring of the system’s reaction with 1-hexene showed the consumption of 9·MAO adduct and the accumulation of the vinylidene dimer (Figures S24 and S25).
Thus, biszirconium complex 9 is readily formed in the systems Cp2ZrCl2-HAlBui2, [Cp2ZrH2]2-MMAO-12, and [Cp2ZrH2]2-ClAlBui2 both in CD2Cl2 and in CDCl3. This intermediate reacts with methylaluminoxane to give a heavy adduct, which selectively provides vinylidene dimers of 1-hexene.

2.3. DFT Study of the Structure of Biszirconium Complex 9

To refine the structure of complex 9, probable structures of cyclic isomers 9a9d were optimized using the PBE/3ζ quantum chemical method [30,31,32]. Their structure is in line with the obtained NMR spectral data on the ratio of the signal intensities of the constituent moieties and the symmetry of the complex (Scheme 4). As follows from Scheme 4, the structures of the complexes significantly differ from one another. For example, in complex 9c, all three hydrides are located between the zirconium atoms and form a trihydride bridge, while in molecule 9a, there is only one bridging H atom. If the optimized structures of the two most energetically favorable hydride complexes 9a and 9c (Figure 3) are considered in detail, the length of the Zr–H bond varies, which may cause differences in the reactivity of the studied complexes. Thus, the lengths of both Zr–H bonds in the Zr–H–Zr moiety of complex 9a is 2.09 Å, while the lengths of two terminal Zr–H bonds are dZr1-H = dZr2-H = 1.83 Å. It is comparable with the Zr–H bond length in Cp2ZrHCl calculated by the same method (dZr-H = 1.84 Å).
In isomer 9c, all three hydrogen atoms are inside the biszirconium cage. Meanwhile, the lengths of bridging Zr–H bonds are also increased compared to those in Cp2ZrHCl, but are not equivalent to each other: dZr1-H1 = dZr2-H1 = 2.01 Å, dZr1-H2 = dZr2-H3 = 1.97 Å, dZr1-H3 = dZr2-H2 = 1.99 Å. Thus, one hydrogen atom is equidistant from both zirconium atoms, while the other two H atoms, forming an “inner” bridge, are characterized by some displacement towards one of the zirconium atoms. It is quite obvious that structural differences also determine energy differences (Table 4, Table S1).
The most thermodynamically stable complex is 9a, in which the chlorine atoms are part of the Zr–Cl–Al moieties. Isomer 9b is higher in energy by 3.3 kcal/mol. The presence of bulky Cl atoms in the inner bridge of the biszirconium cage of structure 9d makes this complex least thermodynamically stable relative to other compounds.
To identify the structure of the complex observed by NMR spectroscopy, we compared theoretical and experimental NMR chemical shifts of hydride atoms in each of them (Table 5). It was found that in complex 9a, the hydrogen atom of the Zr–H–Zr moiety is significantly shielded (δH1 = −4.2 ppm), while two other hydride atoms experience a de-shielding effect (δH2 = δH3 = 2.9 ppm). Thus, the difference in chemical shifts between the considered hydrogen atoms is ∆δ = 7.1 ppm, which is in good agreement with the experimental data. A similar trend is also observed for complex 9c, in which one of the hydride atoms of the inner Zr–H–Zr bridge should be in the upfield region of the 1H NMR spectrum. As follows from Table 5, the calculated NMR data for complexes 9b and 9d are in poor agreement with the NMR spectral parameters observed. It should be noted for comparison that the calculated chemical shifts of the hydride atoms of the open structure 9e proposed earlier [23] also do not agree well with the NMR experiment. As a result of a comprehensive analysis of chemical shifts and relative Gibbs energy, the complex 9a was proposed as the most probable structure.

3. Materials and Methods

3.1. General Procedures

All operations for organometallic compounds were performed under argon according to the Schlenk technique. Zirconocene 1 was prepared from ZrCl4 (99.5%, Merck, Darmstadt, Germany) using the standard procedure [33]. The synthesis of [Cp2ZrH2]2 (2) from (1) was carried out as described previously [28,34]. The solvents (CHCl3, CH2Cl2, ClCH2CH2Cl) were distilled from P2O5 immediately before use. The solvent o-Cl2C6H4 (anhydrous, 99%, Merck) was used without further purification. Commercially available Cp2HfCl2 (98%, Strem, Newburyport, MA, USA), Cp2TiCl2 (99%, Strem), (C5Me5)2ZrCl2 (97%, Acros), rac-C2H4(THInd)2ZrCl2 (97%, Merck), HAlBui2 (99%, Merck), ClAlBui2 (97%, Strem), MMAO-12 (7% wt Al in toluene, Merck), (Ph3C)[B(C6F5)4] (97%, Abcr, Karlsruhe, Germany), B(C6F5)3 (95%, Merck), and 1-hexene (97%, Fisher Scientific, Pittsburgh, Pennsylvania, USA) were used for the reactions.
CAUTION: pyrophoric nature of aluminum alkyl and hydride compounds requires special safety precautions in their handling.
1H and 13C NMR spectra were recorded on a Bruker AVANCE−400 spectrometer (400.13 MHz (1H), 100.62 MHz (13C)) (Bruker, Rheinstetten, Germany). As the solvents and the internal standards, CD2Cl2 and CDCl3 were employed. 1D and 2D NMR spectra (COSY HH, HSQC, HMBC, NOESY) were recorded using standard Bruker pulse sequences.
The products were analyzed using a GC–MS-QP2010 Ultra gas chromatograph–mass spectrometer (Shimadzu, Tokyo, Japan) equipped with the GC−2010 Plus chromatograph (Shimadzu, Tokyo, Japan), TD-20 thermal desorber (Shimadzu, Tokyo, Japan), and an ultrafast quadrupole mass-selective detector (Shimadzu, Tokyo, Japan). Details on the GC–MS analysis of dimers and oligomers are given in the SI.

3.2. Reaction of [Cp2ZrH2]2 with ClAlBui2, Activators (MMAO-12, (Ph3C)[B(C6F5)4] or B(C6F5)3) and 1-Hexene

A flask with a magnetic stirrer was filled under argon with 10 mg (0.045 mmol) [Cp2ZrH2]2 (all ratios are given relative to the monomer), 0.026 mL (0.135 mmol) ClAlBui2, 0.58 mL (1.35 mmol) MMAO-12, 2.24 mL or 5,63 mL (17.9 or 45 mmol) 1-hexene, and 2 mL solvent (CHCl3, CH2Cl2, o-Cl2C6H4, ClCH2CH2Cl). For organoboron activators (Ph3C)[B(C6F5)4] or B(C6F5)3 the following amounts were used: 10 mg (0.045 mmol) [Cp2ZrH2]2, 0.018–0.036 mL (0.10–0.18 mmol) ClAlBui2, 6 mg (0.011 mmol) B(C6F5)3 or 10 mg (0.011 mmol) (Ph3C)[B(C6F5)4], 0.6 mL (4.5 mmol) 1-hexene, and 2 mL solvent (CHCl3, CH2Cl2, o-Cl2C6H4, ClCH2CH2Cl). The reaction was carried out with stirring at a temperature of 40 °C. After 5, 10, 15, 30, 60, 90, 120, 180, and 960 min, samples (0.1 mL) were syringed into tubes filled with argon and decomposed with 10% HCl at 0 °C. The products were extracted with CH2Cl2, and the organic layer was dried over Na2SO4. The yield of dimers or oligomers was determined by GC/MS.

3.3. Reaction of L2MCl2 (1af) with HAlB i2, MMAO-12, (Ph3C)[B(C6F5)4] or B(C6F5)3 and 1-Hexene

A flask with a magnetic stirrer was filled under argon with 0.034 mmol (9–15 mg) L2MCl2, 0.019 mL (0.108 mmol) HAlBui2, 0.44 mL (1.02 mmol) MMAO-12, 1.7 or 4.2 mL (13.6 or 34 mmol) 1-hexene, and 2 mL solvents (CHCl3, CH2Cl2, o-Cl2C6H4, ClCH2CH2Cl). For organoboron activators (Ph3C)[B(C6F5)4] or B(C6F5)3 the following amounts were used: 0.034 mmol (9–15 mg) L2MCl2, 0.018–0.036 mL (0.10–0.18 mmol) HAlBui2, 5 mg (0.0085 mmol) B(C6F5)3 or 10–16 mg (0.011–0.017 mmol) (Ph3C)[B(C6F5)4], 1 or 1.7 mL (8.5 or 13.6 mmol) 1-hexene, and 2 mL solvents (CHCl3, CH2Cl2, o-Cl2C6H4, ClCH2CH2Cl). The reaction was carried out with stirring at a temperature of 40 °C. After 15, 20, 30, 60, 90, 120, 180, and 960 min, samples (0.1 mL) were syringed into tubes filled with argon and decomposed with 10% HCl or DCl at 0 °C. The products were extracted with CH2Cl2, and the organic layer was dried over Na2SO4. The yields of the products were determined by GC/MS.

3.4. NMR Study of the Reaction of [Cp2ZrH2]2 with ClAlR2 and MMAO-12

An NMR tube was charged with 0.045 mmol (10 mg) [Cp2ZrH2]2 in an argon-filled glovebox. The tube was cooled to 0 °C, and 0.045–0.2 mmol (8–35.3 mg) ClAlBui2 was added dropwise. Then CDCl3 or CD2Cl2 was added. The mixture was stirred, and the formation of complexes 810 was monitored by NMR at room temperature. Then 0.135–0.9 mmol (60–300 mg) MMAO-12 was added, and the formation of intermediates was monitored by NMR at room temperature (Figures S9–S18).
In the case of the system [Cp2ZrH2]2-MMAO-12, the number of components was the same as described above (Figures S19–S23). The deuterated solvent was added to the NMR tube as the last component.

3.5. NMR Study of the Reaction of Cp2ZrCl2 with HAlBui2 and MMAO-12

An NMR tube was charged with 0.034 mmol (10 mg) Cp2ZrCl2 in an argon-filled glovebox. The tube was cooled to 0 °C, and 0.068–0.136 mmol (10–19.3 mg) HAlBui2 was added dropwise. Then CDCl3 or CD2Cl2 was added. The mixture was stirred, and the formation of complexes 810 was monitored by NMR at room temperature. Then 0.135–0.9 mmol (60–300 mg) MMAO-12 was added, and the formation of intermediates was monitored by NMR at room temperature (Figures S1–S8).

3.6. Computational Details

DFT calculations were carried out in Priroda-06 program [32,35]. Geometry optimization, vibrational frequency analysis, and calculation of absolute chemical shielding, entropy, and thermodynamic corrections to the total energy of the compounds were carried out using the Perdew−Burke−Ernzerhof (PBE) functional [30]. PBE functional was used in combination with a 3ζ basis set [31]. The electronic configurations of the molecular systems were described by the orbital basis sets of contracted Gaussian-type functions of size (5s1p)/[3s1p] for H, (11s6p2d)/[6s3p2d] for C, (15s11p2d)/[10s6p2d] for Al and Cl, and (20s16p11d)/[14s11p7d] for Zr, which were used in combination with the density-fitting basis sets of uncontracted Gaussian-type functions of size (5s2p) for H, (10s3p3d1f) for C, (14s3p3d1f1g) for Al and Cl, and (22s5p5d4f4g) for Zr. No symmetry of internal coordinate constraints was applied during optimizations. Thermodynamic parameters were determined at 298.15. Normal-mode vibrational frequency analysis was performed to confirm minima structures. Computations of absolute chemical shielding, σ, were carried out in the GIAO approach [36,37]. Chemical shifts were calculated as δ = σTMS − σ, where σTMS was the calculated shielding constant of tetramethylsilane. For comparison, structural and energetic parameters of complexes 9ae were calculated using Gaussian 09 [38] at the PBE0 level of theory [39] employing the def2-TZVP basis set [40,41] with and without Grimme’s D3(0) empirical dispersion correction (GD3) [42] (Tables S2 and S3). Moreover, the method was successfully employed to calculate chemical shifts of Au hydrides [43]. Calculated chemical shifts obtained for 9a by two methods was comparable (PBE0/def2-TZVP: δH1 = −3.5 ppm, δH2 = 3.1 ppm, δH3 = 3.1 ppm). Therefore, the PBE/3ζ method was used for the calculation of the chemical shifts of other complexes. As follows from Table S3 (Supporting Information), using the GD3 correction does not lead to significant changes in the energy of the studied complexes. Furthermore, we carried out the study on the solvent effect (CH2Cl2 and CHCl3) using the conductor-like polarizable continuum model (CPCM) [44,45]. The data indicated minor energy changes (Supporting Information, Tables S4 and S5).
The program ChemCraft [46] was used to visualize obtained quantum chemical data. The energy at 0 K, the ZPVE correction, the enthalpy, the Gibbs free energy in gas, the temperature multiplied by the entropy, and Cartesian coordinates for all optimized structures are given in the Supporting Information.

4. Conclusions

As a result of studying the catalytic transformations of 1-hexene under the action of Ti group metallocenes, organoaluminum compounds, and activators MMAO-12, B(C6F5)3 or (Ph3C)[B(C6F5)4] in chlorinated solvents, it was found that systems based on Zr complexes [Cp2ZrH2]2-ClAlBui2-MMAO-12, Cp2ZrCl2-HAlBui2-MMAO-12 in CH2Cl2 selectively afford dimeric products in high yields. The use of CHCl3 as the solvent facilitates the formation of non-classical tetramers of 1-hexene, the products of dimer dimerization.
NMR studies showed that systems Cp2ZrCl2-HAlBui2-MMAO-12, [Cp2ZrH2]2-MMAO-12, and [Cp2ZrH2]2-ClAlBui2-MMAO-12 provide the adduct of biszirconium complex (Cp2ZrH2∙Cp2ZrHCl∙ClAlR2) with an activator both in CD2Cl2 and CDCl3. The adduct reacts with 1-hexene and produces a vinylidene dimer.
The probable structure of the key biszirconium hydride complex was proposed based on a comparison of experimental and theoretical NMR data and estimation of the thermodynamical stability of the complexes. These studies are necessary to further understand the mechanism of key intermediate activation by methylaluminoxane or organoboron compounds for selective alkene dimerization.

Supplementary Materials

Supplementary Materials are available online. Figures S1–S27: NMR spectra of catalytic systems, Figures S28 and S29: Examples of GC-MS analysis of dimers and oligomers, Tables S1–S5: Calculated thermodynamic parameters of isomeric complexes 9a–d.

Author Contributions

Conceptualization, L.V.P.; methodology, P.V.K. and T.V.T; validation, L.V.P., P.V.K. and T.V.T.; formal analysis, P.V.K. and T.V.T.; investigation, P.V.K., A.K.B., D.N.I. and V.M.Y.; resources, P.V.K. and L.V.P.; data curation, P.V.K., T.V.T. and L.V.P.; writing—original draft preparation, P.V.K. and T.V.T.; writing—review and editing, L.V.P.; visualization, P.V.K. and D.N.I.; supervision, L.V.P.; project administration, P.V.K.; funding acquisition, P.V.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 19-73-10122.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The structural studies of compounds were carried out at the Center for Collective Use “Agidel” at the Institute of Petrochemistry and Catalysis, Russian Academy of Sciences.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Organometallic Reactions and Polymerization; Osakada, K. (Ed.) Springer-Verlag: Berlin/Heidelberg, Germany, 2014; p. 301. [Google Scholar] [CrossRef]
  2. Nicholas, C.P. Applications of light olefin oligomerization to the production of fuels and chemicals. Appl. Cat. A Gen. 2017, 543, 82–97. [Google Scholar] [CrossRef]
  3. de Klerk, A. Oligomerization. In Fischer-Tropsch Refining; de Klerk, A., Ed.; Wiley-VCH Verlag: Hoboken, NJ, USA, 2011; pp. 369–391. [Google Scholar] [CrossRef]
  4. McGuinness, D.S. Olefin Oligomerization via Metallacycles: Dimerization, Trimerization, Tetramerization, and Beyond. Chem. Rev. 2011, 111, 2321–2341. [Google Scholar] [CrossRef]
  5. Nifant’ev, I.; Ivchenko, P.; Tavtorkin, A.; Vinogradov, A.; Vinogradov, A. Non-traditional Ziegler-Natta Catalysis in a-Olefin Transformations: Reaction Mechanisms and Product Design. Pure Appl. Chem. 2017, 89, 1017–1032. [Google Scholar] [CrossRef]
  6. Janiak, C. Metallocene and Related Catalysts for Olefin, Alkyne and Silane Dimerization and Oligomerization. Coord. Chem. Rev. 2006, 250, 66–94. [Google Scholar] [CrossRef]
  7. Comyns, A.E. Encyclopedic Dictionary of Named Processes in Chemical Technology, 4th ed.; CRC Press: Boca Raton, FL, USA, 2014; p. 416. [Google Scholar] [CrossRef]
  8. Harvey, B.G.; Meylemans, H.A. 1-Hexene: A Renewable C6 Platform for Full-performance Jet and Diesel fuels. Green Chem. 2014, 16, 770–776. [Google Scholar] [CrossRef]
  9. Chen, E.Y.-X.; Marks, T.J. Cocatalysts for Metal-Catalyzed Olefin Polymerization:  Activators, Activation Processes, and Structure−Activity Relationships. Chem. Rev. 2000, 100, 1391–1434. [Google Scholar] [CrossRef]
  10. Resconi, L.; Cavallo, L.; Fait, A.; Piemontesi, F. Selectivity in Propene Polymerization with Metallocene Catalysts. Chem. Rev. 2000, 100, 1253–1346. [Google Scholar] [CrossRef] [PubMed]
  11. Kaminsky, W. The Discovery of Metallocene Catalysts and Their Present State of the Art. J. Polym. Sci. A Polym. Chem. 2004, 42, 3911–3921. [Google Scholar] [CrossRef]
  12. Collins, R.A.; Russell, A.F.; Mountford, P. Group 4 Metal Complexes for Homogeneous Olefin Polymerisation: A Short Tutorial Review. App. Petroch. Res. 2015, 5, 153–171. [Google Scholar] [CrossRef] [Green Version]
  13. Christoffers, J.; Bergman, R.G. Catalytic Dimerization Reactions of α-Olefins and α,ω-Dienes with Cp2ZrCl2/Poly(methylalumoxane):  Formation of Dimers, Carbocycles, and Oligomers. J. Am. Chem. Soc. 1996, 118, 4715–4716. [Google Scholar] [CrossRef]
  14. Christoffers, J.; Bergman, R.G. Zirconocene-alumoxane (1:1)—A Catalyst for the Selective Dimerization of α-Olefins. Inorg. Chim. Acta 1998, 270, 20–27. [Google Scholar] [CrossRef]
  15. Nifant'ev, I.E.; Vinogradov, A.A.; Vinogradov, A.A.; Ivchenko, P.V. Zirconocene-Catalyzed Dimerization of 1-Hexene: Two-stage Activation and Structure–Catalytic Performance Relationship. Cat. Commun. 2016, 79, 6–10. [Google Scholar] [CrossRef]
  16. Nifant’ev, I.E.; Vinogradov, A.A.; Vinogradov, A.A.; Sedov, I.V.; Dorokhov, V.G.; Lyadov, A.S.; Ivchenko, P.V. Structurally Uniform 1-Hexene, 1-Octene, and 1-Decene Oligomers: Zirconocene/MAO-Catalyzed Preparation, Characterization, and Prospects of Their Use as low-viscosity Low-temperature Oil Base Stocks. Appl. Cat. A Gen. 2018, 549, 40–50. [Google Scholar] [CrossRef]
  17. Nifant'ev, I.E.; Vinogradov, A.A.; Vinogradov, A.A.; Churakov, A.V.; Bagrov, V.V.; Kashulin, I.A.; Roznyatovsky, V.A.; Grishin, Y.K.; Ivchenko, P.V. The Catalytic Behavior of Heterocenes Activated by TIBA and MMAO under a Low Al/Zr Ratios in 1-Octene Polymerization. Appl. Cat. A Gen. 2019, 571, 12–24. [Google Scholar] [CrossRef]
  18. Nifant’ev, I.; Vinogradov, A.; Vinogradov, A.; Karchevsky, S.; Ivchenko, P. Experimental and Theoretical Study of Zirconocene-Catalyzed Oligomerization of 1-Octene. Polymers 2020, 12, 1590. [Google Scholar] [CrossRef] [PubMed]
  19. Parfenova, L.V.; Kovyazin, P.V.; Bikmeeva, A.K. Bimetallic Zr,Zr-Hydride Complexes in Zirconocene Catalyzed Alkene Dimerization. Molecules 2020, 25, 2216. [Google Scholar] [CrossRef]
  20. Parfenova, L.V.; Kovyazin, P.V.; Bikmeeva, A.K.; Palatov, E.R. Catalytic Systems Based on Cp2ZrX2 (X = Cl, H), Organoaluminum Compounds and Perfluorophenylboranes: Role of Zr,Zr- and Zr,Al-Hydride Intermediates in Alkene Dimerization and Oligomerization. Catalysts 2021, 11, 39. [Google Scholar] [CrossRef]
  21. Nifant’ev, I.; Vinogradov, A.; Vinogradov, A.; Karchevsky, S.; Ivchenko, P. Zirconocene-Catalyzed Dimerization of α-Olefins: DFT Modeling of the Zr-Al Binuclear Reaction Mechanism. Molecules 2019, 24, 3565. [Google Scholar] [CrossRef] [Green Version]
  22. Nifant’ev, I.; Ivchenko, P. Fair Look at Coordination Oligomerization of Higher α-Olefins. Polymers 2020, 12, 1082. [Google Scholar] [CrossRef]
  23. Parfenova, L.V.; Kovyazin, P.V.; Tyumkina, T.V.; Islamov, D.N.; Lyapina, A.R.; Karchevsky, S.G.; Ivchenko, P.V. Reactions of bimetallic Zr,Al- hydride complexes with methylaluminoxane: NMR and DFT study. J. Organomet. Chem. 2017, 851, 30–39. [Google Scholar] [CrossRef]
  24. Dong, S.Q.; Mi, P.K.; Xu, S.; Zhang, J.; Zhao, R.D. Preparation and Characterization of Single-Component Poly-α-olefin Oil Base Stocks. Energy Fuels 2019, 33, 9796–9804. [Google Scholar] [CrossRef]
  25. Zhao, R.; Mi, P.; Xu, S.; Dong, S. Structure and Properties of Poly-α-olefins Containing Quaternary Carbon Centers. ACS Omega 2020, 5, 9142–9150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Shao, H.; Gu, X.; Wang, R.; Wang, X.; Jiang, T.; Guo, X. Preparation of Lubricant Base Stocks with High Viscosity Index through 1-Decene Oligomerization Catalyzed by Alkylaluminum Chloride Promoted by Metal Chloride. Energy Fuels 2020, 34, 2214–2220. [Google Scholar] [CrossRef]
  27. Parfenova, L.V.; Kovyazin, P.V.; Gabdrakhmanov, V.Z.; Istomina, G.P.; Ivchenko, P.V.; Nifant'Ev, I.E.; Khalilov, L.M.; Dzhemilev, U.M. Ligand Exchange Processes in Zirconocene Dichloride-Trimethylaluminum Bimetallic Systems and Their Catalytic Properties in Reaction with Alkenes. Dalton Trans. 2018, 47, 16918–16937. [Google Scholar] [CrossRef] [PubMed]
  28. Parfenova, L.V.; Pechatkina, S.V.; Khalilov, L.M.; Dzhemilev, U.M. Mechanism of Cp2ZrCl2-Catalyzed Olefin Hydroalumination by Alkylalanes. Russ. Chem. Bull. 2005, 54, 316–327. [Google Scholar] [CrossRef]
  29. Parfenova, L.V.; Kovyazin, P.V.; Nifant’ev, I.E.; Khalilov, L.M.; Dzhemilev, U.M. Role of Zr,Al Hydride Intermediate Structure and Dynamics in Alkene Hydroalumination with XAlBui2 (X = H, Cl, Bui), Catalyzed by Zr η5-Complexes. Organometallics 2015, 34, 3559–3570. [Google Scholar] [CrossRef]
  30. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  31. Laikov, D.N. Razvitiye ekonomnogo podkhoda k raschetu molekul metodom funktsionala plotnosti i yego primeneniye k resheniyu slozhnykh khimicheskikh zadach. Ph.D Thesis, Moscow State University, Moscow, Russia, 2000. (In Russian). [Google Scholar]
  32. Laikov, D.N. Fast Evaluation of Density Functional Exchange-Correlation Terms Using the Expansion of the Electron Density in Auxiliary Basis Sets. Chem. Phys. Lett. 1997, 281, 151–156. [Google Scholar] [CrossRef]
  33. Freidlina, R.K.; Brainina, E.M.; Nesmeyanov, A.N. The Synthesis of Mixed Pincerlike Cyclopentadienyl Compounds of Zirconium. Dokl. Acad. Nauk SSSR 1961, 138, 1369–1372. [Google Scholar]
  34. Shoer, L.I.; Gell, K.I.; Schwartz, J. Mixed-Metal Hydride Complexes Containing Zr-H-Al Bridges. Synthesis and Relation to Transition-Metal-Catalyzed Reactions of Aluminum Hydrides. J. Organomet. Chem. 1977, 136, c19–c22. [Google Scholar] [CrossRef]
  35. Laikov, D.N.; Ustynyuk, Y.A. PRIRODA-04: A quantum-chemical program suite. New possibilities in the study of molecular systems with the application of parallel computing. Russ. Chem. Bull. 2005, 54, 820–826. [Google Scholar] [CrossRef]
  36. Ditchfield, R. Self-consistent Perturbation Theory of Diamagnetism. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  37. Wolinski, K.; Hinton, J.F.; Pulay, P. Efficient Implementation of the Gauge-independent Atomic Orbital Method for NMR Chemical Shift Calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [Google Scholar] [CrossRef]
  38. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09 Rev. D.01; Gaussian: Wallingford, CT, USA, 2009. [Google Scholar]
  39. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  40. Weigend, F.; Ahlrichs, R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef] [PubMed]
  41. Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjustedab Initio Pseudopotentials for the Second and Third Row Transition Elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  42. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  43. Rocchigiani, L.; Fernandez-Cestau, J.; Chambrier, I.; Hrobárik, P.; Bochmann, M. Unlocking Structural Diversity in Gold(III) Hydrides: Unexpected Interplay of cis/trans-Influence on Stability, Insertion Chemistry, and NMR Chemical Shifts. J. Am. Chem. Soc. 2018, 140, 8287–8302. [Google Scholar] [CrossRef]
  44. Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar] [CrossRef]
  45. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem. 2003, 24, 669–681. [Google Scholar] [CrossRef]
  46. Zhurko, G.A.; Zhurko, D.A. ChemCraft 1.6; Informer Technologies, Inc.: Roseau, Dominica, 2009. [Google Scholar]
Scheme 1. Zirconocene-catalyzed dimerization of terminal alkenes.
Scheme 1. Zirconocene-catalyzed dimerization of terminal alkenes.
Molecules 26 02775 sch001
Scheme 2. 1-Hexene dimerization and oligomerization in catalytic systems AD.
Scheme 2. 1-Hexene dimerization and oligomerization in catalytic systems AD.
Molecules 26 02775 sch002
Scheme 3. The reaction of Cp2ZrCl2 and [Cp2ZrH2]2 with OACs and MMAO-12.
Scheme 3. The reaction of Cp2ZrCl2 and [Cp2ZrH2]2 with OACs and MMAO-12.
Molecules 26 02775 sch003
Figure 1. 1H NMR spectrum of the system [Cp2ZrH2]2-ClAlBui2-MMAO-12 in CD2Cl2 (T = 298 K): (a) [Cp2ZrH2]:[ClAlBui2]:[MMAO-12] = 1:2:0; (b) [Cp2ZrH2]:[ClAlBui2]:[MMAO-12] = 1:2:5; (c) [Cp2ZrH2]:[ClAlBui2] [MMAO-12] = 1:0:7.
Figure 1. 1H NMR spectrum of the system [Cp2ZrH2]2-ClAlBui2-MMAO-12 in CD2Cl2 (T = 298 K): (a) [Cp2ZrH2]:[ClAlBui2]:[MMAO-12] = 1:2:0; (b) [Cp2ZrH2]:[ClAlBui2]:[MMAO-12] = 1:2:5; (c) [Cp2ZrH2]:[ClAlBui2] [MMAO-12] = 1:0:7.
Molecules 26 02775 g001
Figure 2. 1H NMR spectrum of systems Cp2ZrCl2-HAlBui2-MMAO-12 and [Cp2ZrH2]2-MMAO-12 in CDCl3 (T = 298 K): (a) [Cp2ZrCl2]:[HAlBui2]:[MMAO-12] = 1:4:0; (b) [Cp2ZrCl2]:[HAlBui2]:[MMAO-12]:[1-hexene] = 1:2:11:0.1; (c) [Cp2ZrH2]:[MMAO-12] = 1:12.
Figure 2. 1H NMR spectrum of systems Cp2ZrCl2-HAlBui2-MMAO-12 and [Cp2ZrH2]2-MMAO-12 in CDCl3 (T = 298 K): (a) [Cp2ZrCl2]:[HAlBui2]:[MMAO-12] = 1:4:0; (b) [Cp2ZrCl2]:[HAlBui2]:[MMAO-12]:[1-hexene] = 1:2:11:0.1; (c) [Cp2ZrH2]:[MMAO-12] = 1:12.
Molecules 26 02775 g002
Figure 3. Optimized structures of isomers 9a and 9c.
Figure 3. Optimized structures of isomers 9a and 9c.
Molecules 26 02775 g003
Scheme 4. Theoretically calculated structures of isomers of complex 9.
Scheme 4. Theoretically calculated structures of isomers of complex 9.
Molecules 26 02775 sch004
Table 1. Catalytic activity and chemoselectivity of systems A and B in the reaction with 1-hexene.
Table 1. Catalytic activity and chemoselectivity of systems A and B in the reaction with 1-hexene.
EntryCatalytic Systems[Zr]: [Al]: [Activator]:[1-Hexene]SolventT, °CTime, minAlkene Conversion, % Product Composition, % h
Zr Complex OAC aActivator4567
n = 1n = 2n = 3
1 [19][Cp2ZrH2]2ClAlBui2MMAO-121:3:30:100C6H5CH34015>99 b 86
2ClAlBui2MMAO-121:3:30:400CH2Cl24015>99 981.3
3180 98.21.8
4ClAlBui2MMAO-121:3:30:400CHCl3401592 92
560>99 98
6105 971.31.6
7180912.61.42 75
896017-1.42.6 79
9-MMAO-121:30:400CHCl34030961911
10CH2Cl2>99<1898
11o-Cl2C6H4608617310<1
1296093<1741331
13(CH2Cl)26090 835
1496099 9141
15 [20]ClAlBui2B(C6F5)34:8:1:400C6H6409081 81
16-B(C6F5)34:1:400CHCl3409600
17CH2Cl29600
18ClAlBui24:16:1:400CHCl318042 402
1996075 714
20CH2Cl29600
21 [20]ClAlEt2(Ph3C)[B(C6F5)4]4:8:1:400C6H6409091 c 86
22-(Ph3C)[B(C6F5)4]4:1:400CHCl3409600
23CH2Cl29600
24ClAlBui24:16:1:400CHCl318044 920- 15
2596081 173613 15
26CH2Cl2180>99 1318 69
27 [19]Cp2ZrCl2HAlBui2MMAO-121:3:30:100C6H5CH34015>99 d 91
28HAlBui2MMAO-121:3:30:400CH2Cl2403098 971
296099 981
30-1:30:400CH2Cl24018098 96--
3196099 924<1
32HAlBui2MMAO-121:3:30:1000CH2Cl2403082 802
336088 772 9
34HAlBui2MMAO-121:3:30:400CHCl34030>99 982
35HAlBui2MMAO-121:3:30:1000CHCl34030>99 907 3
36HAlBui2MMAO-121:3:30:400CHCl320180>99 972
37-MMAO-121:30:400CHCl340180>99532 89
38-MMAO-121:10:400CHCl34096091 91<1
39-MMAO-121:30:400o-Cl2C6H4403099 934-
4096099 9151
41-MMAO-121:30:400(CH2Cl)2403099 952
42960>99 962
43 [20]HAlBui2B(C6F5)34:16:1:1000C6H64060>99 e 93
44HAlBui2B(C6F5)34:16:1:1000CH2Cl24060>99 991
45CHCl36083 821
46-B(C6F5)34:1:1000CH2Cl2409600
47CHCl3409600
48 [20]HAlBui2(Ph3C)[B(C6F5)4]4:16:1:1000C6H6609097 f 67
49HAlBui2(Ph3C)[B(C6F5)4]4:16:1:1000CH2Cl220180>99 926 2
5096015583 33
51HAlBui2(Ph3C)[B(C6F5)4]4:16:1:1000CH2Cl24030>99 g36061 27
52960 1212443
53HAlBui2(Ph3C)[B(C6F5)4]4:16:1:1000CHCl320180>99 7682 13
5440180 15811 65
55960 8812 72
56-(Ph3C)[B(C6F5)4]4:1:1000CH2Cl2409600
57CHCl39600
a OAC—organoaluminum compound; b hydrometalation product, 4%; saturated dimer, 2%; unsaturated methylalumination product, 8% [19]; c hydrometalation product, 5% [20]; d hydrometalation product, 5%; saturated dimer, 1%; unsaturated methylalumination product, 2%; e hydrometalation product, 3%; trimers, 3%; f oligomers, 23%; g 6-mers (29%) are identified; h wt % in the reaction mixture, determined by GC–MS of deuterolysis or hydrolysis products (the details of the GC–MS analysis is shown in Supporting Information, Figures S28 and S29).
Table 2. Catalytic activity and chemoselectivity of systems Cp2MCl2-HAlBui2-activator (M = Ti, Hf) in the reaction with 1-hexene.
Table 2. Catalytic activity and chemoselectivity of systems Cp2MCl2-HAlBui2-activator (M = Ti, Hf) in the reaction with 1-hexene.
EntryCatalytic Systems[Zr]: [Al]: [Activator]:[1-Alkene]SolventT, °CTime, minAlkene Conversion, % Product Composition, % d
Zr Complex OACActivator4567
n = 1n = 2n = 3
1Cp2TiCl2 aHAlBui2MMAO-121:3:30:400CH2Cl2406080 20344 15
218091 17445 20
3CHCl36093 16366528
4960>99544111730
5Cp2HfCl2HAlBui2MMAO-121:3:30:400CH2Cl240120842532062
6180942542182
796096-5922834
8MMAO-121:3:30:400CHCl3406060-42153 1
912074 b-30164 5
10960>99 c6171715 9
11Cp2TiCl2HAlBui2(Ph3C)[B(C6F5)4]4:16:1:400CH2Cl2409600
12CHCl39600
13Cp2HfCl2HAlBui2(Ph3C)[B(C6F5)4]CH2Cl2409600
14CHCl39600
a regioselectivity is significantly reduced due to double bond migration; b byproducts of toluene mono- (13%), di- (4%) and tri- (2%) alkylation with 1-hexene are formed; c byproducts of toluene mono- (22%), di- (7%), and tri- (4%) alkylation with 1-hexene are formed; d wt % in the reaction mixture, determined by GC–MS of deuterolysis or hydrolysis products (the details of the GC–MS analysis is shown in Supporting Information).
Table 3. Catalytic activity and chemoselectivity of systems L2ZrCl2-HAlBui2-activator in the reaction with 1-hexene.
Table 3. Catalytic activity and chemoselectivity of systems L2ZrCl2-HAlBui2-activator in the reaction with 1-hexene.
EntryCatalytic Systems[Zr]: [Al]: [activator]:[1-alkene]SolventT, °CTime, minAlkene Conversion, % Product Composition, % e
Zr Complex OACActivator456
n = 1n = 2n = 3
1(C5Me5)2ZrCl2HAlBui2MMAO-121:3:30:400CHCl34030>99 a
2CH2Cl230
3Ind2ZrCl2HAlBui2MMAO-121:3:30:400CH2Cl2403019 11521
418048 201963
596085 b 9181816
6CHCl340180>99 13512412
7960>99 c 16211918
8rac-H4C2[THInd]2ZrCl2HAlBui2MMAO-121:3:30:400CHCl34030>99 d 3938164
9CH2Cl29600
a 1-hexene is completely consumed for toluene alkylation; b 6-mers (15%) and 7-mers (9%); c 6-mers (17%) and 7-mers (9%); d 6-mers (3%); e wt % in the reaction mixture, determined by GC–MS of deuterolysis or hydrolysis products (the details of the GC–MS analysis is shown in Supporting Information).
Table 4. Relative thermodynamic parameters of isomeric complexes 9.
Table 4. Relative thermodynamic parameters of isomeric complexes 9.
Complex∆E, Hartree∆EZPVE,
Hartree
∆H,
kcal/mol
∆G,
kcal/mol
T∆S,
cal/mol
9a0.0000000.0000000.00.02475.2
9b0.0131030.0129488.29.61139.2
9c0.0030220.0033982.43.31517.2
9d0.0221460.02311114.416.90.0
Table 5. Calculated and experimental chemical shifts of isomeric complexes 9ad (numbering according to Scheme 4).
Table 5. Calculated and experimental chemical shifts of isomeric complexes 9ad (numbering according to Scheme 4).
Complexδ(H 1), ppmδ(H 2), ppmδ(H 3), ppmδ(Cp), ppm
9a−4.22.92.95.9
9b−2.31.81.86.2
9c−3.63.73.76.2
9d0.70.80.86.0
9e−1.6−0.4−0.46.1
9 (experimental)−5.9−0.7−0.76.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kovyazin, P.V.; Bikmeeva, A.K.; Islamov, D.N.; Yanybin, V.M.; Tyumkina, T.V.; Parfenova, L.V. Ti Group Metallocene-Catalyzed Synthesis of 1-Hexene Dimers and Tetramers. Molecules 2021, 26, 2775. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26092775

AMA Style

Kovyazin PV, Bikmeeva AK, Islamov DN, Yanybin VM, Tyumkina TV, Parfenova LV. Ti Group Metallocene-Catalyzed Synthesis of 1-Hexene Dimers and Tetramers. Molecules. 2021; 26(9):2775. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26092775

Chicago/Turabian Style

Kovyazin, Pavel V., Almira Kh. Bikmeeva, Denis N. Islamov, Vasiliy M. Yanybin, Tatyana V. Tyumkina, and Lyudmila V. Parfenova. 2021. "Ti Group Metallocene-Catalyzed Synthesis of 1-Hexene Dimers and Tetramers" Molecules 26, no. 9: 2775. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26092775

Article Metrics

Back to TopTop