Next Article in Journal
A Bifunctional BiOBr/ZIF-8/ZnO Photocatalyst with Rich Oxygen Vacancy for Enhanced Wastewater Treatment and H2O2 Generation
Previous Article in Journal
In Vitro Antiviral Potential, Antioxidant, and Chemical Composition of Clove (Syzygium aromaticum) Essential Oil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Properties of Zirconocene-Based Systems in 1-Hexene Oligomerization and Structure of Metal Hydride Reaction Centers

by
Lyudmila V. Parfenova
1,*,
Pavel V. Kovyazin
1,
Almira Kh. Bikmeeva
1,
Eldar R. Palatov
1,
Pavel V. Ivchenko
2,3,
Ilya E. Nifant’ev
2,3 and
Leonard M. Khalilov
1
1
Institute of Petrochemistry and Catalysis, Ufa Federal Research Center, Russian Academy of Sciences, Prosp. Oktyabrya, 141, 450075 Ufa, Russia
2
Department of Chemistry, Lomonosov Moscow State University, 1-3 Leninskiye Gory, 119991 Moscow, Russia
3
A.V. Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences, Leninsky Prosp. 29, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2420; https://doi.org/10.3390/molecules28062420
Submission received: 6 February 2023 / Revised: 2 March 2023 / Accepted: 3 March 2023 / Published: 7 March 2023
(This article belongs to the Section Organic Chemistry)

Abstract

:
Despite large-scale investigations of homogeneous single-site metallocene catalysts and systems based on them, there are still unsolved problems related to the control of their activity and chemo- and stereoselectivity. A solution to these problems is required to develop efficient methods for the synthesis of practically useful products of alkene transformations, such as dimers, oligomers, and polymers. Here we studied the catalytic activity of structurally diverse zirconocenes (L2ZrCl2, L = Cp, C5Me5, Ind, L2 = Me2CCp2, Me2SiCp2, Me2C2Cp2, rac-Me2CInd2, rac-H4C2Ind2, BIPh(Ind)2, H4C2[THInd]2), and co-catalysts activating the system, namely HAlBui2, MMAO-12, and (Ph3C)[B(C6F5)4], at low activator/Zr ratios in a 1-hexene oligomerization reaction. The influence of catalyst structure and system composition on the alkene conversion, the type of products, and the reaction stereoselectivity were investigated. The composition of hydride intermediates formed in the L2ZrCl2–HAlBui2–activator system (L2 = ansa-Me2CCp2, Ind) was studied by NMR spectroscopy. Participation of the bis-zirconium hydride complex as the precursor of catalytically active sites of the alkene dimerization reaction was shown.

1. Introduction

Catalytic systems based on metallocenes, organoaluminum compounds (OACs), and activators are well known and have been studied for almost 70 years [1]. A large amount of research is related to the possibility of the efficient synthesis of practically important products: alkene dimers and oligomers, as well as polymers widely known as polyalphaolefins (PAO) [2,3,4,5,6,7]. The use of metallocene complexes in these reactions made it possible to transfer the classical Ziegler–Natta catalysis from a heterogeneous to a homogeneous medium. Hence, highly active and stereoselective single-site catalysts for the polymerization of unsaturated compounds became more controllable, and a detailed study of the reaction mechanism became possible. The high activity and chemo- and stereoselectivity of metallocene complexes are due to their structural features. They provide the most efficient stabilization of the electronic and steric environment of transition metal atoms owing to the high energy of metal–ligand bonds and allow variation of the electrophilicity and geometry of catalytically active sites due to broad possibilities of π-ligand modification. It has been shown that the electronic and steric features of the ligand, the nature of the activator, and the reaction conditions [8,9,10,11,12,13,14,15] determine the productivity and selectivity of catalytic systems.
There is a long-standing interest in the mechanisms of action of homogeneous single-site catalysts, which are constantly being improved and supplemented by new facts [16,17,18,19,20,21,22]. One of the trends in this field is the investigation of the role of hydride intermediates in alkene transformation reactions, especially in dimerization and oligomerization [6,7,23,24,25,26,27,28,29,30,31], the mechanism of which remains an open question.
Our interest in the structural features and reactivity of group 4B metal hydrides is due to the recent data on the alkene transformation reactions catalyzed by zirconocene dichloride Cp2ZrCl2 or zirconocene dihydride [Cp2ZrH2]2 and OACs in the presence of aluminum- or boron-containing activators [32,33,34,35]. The formation of the bimetallic Zr,Zr-hydride intermediate in these systems was established, and its ability to serve as the precursor of the specific and selective catalytically active sites for alkene dimerization was demonstrated.
The present work is aimed at the study of the catalytic activity of structurally diverse zirconocenes (10 examples of cyclopentadienyl and indenyl complexes) in the presence of activators, namely HAlBui2, MMAO-12, and (Ph3C)[B(C6F5)4], at low activator/Zr ratios in 1-hexene oligomerization. The influence of catalyst structure and system composition on the alkene conversion and reaction chemo- and stereoselectivity was investigated. The structure of hydride intermediates formed in the L2ZrCl2–HAlBui2–activator system (L2 = ansa-Me2CCp2, Ind) was studied by means of NMR spectroscopy.

2. Results and Discussion

2.1. Catalytic Effect of L2ZrCl2–HAlBui2–Activator Systems in 1-Hexene Oligomerization

We studied the catalytic effect of complexes 1aj in the oligomerization reaction in the presence of HAlBui2 and the activators MMAO-12 (modified methylaluminoxane) and [Ph3C][B(C6F5)4]. We used di(isobutyl)alane HAlBui2 as a co-catalyst similar to AlBui3 [6,7,26,27,28,30] to initiate the formation of zirconocene hydrides, which act as catalytically active sites in these reactions [29,30,33,34,35]. The reactant ratios were chosen on the basis of our earlier studies [33,34,35]. During the reaction under the selected conditions, the conversion of 1-hexene was approximately 99% in 1–3 h in most of the experiments. A decrease in the alkene conversion down to 85% was observed in the reaction catalyzed by complex 1e in the presence of MMAO-12. 1-Hexene dimer (5) and various oligomers (6) were formed as the major reaction products, which attests to predominant chain termination via β-H elimination mechanism [36] (Scheme 1, Figures S1 and S2). The presence of oligomers with various double bond positions may be due to the migration of zirconium cation upon the formation of allyl complexes [28,37]. Small amounts of M–C bonded oligomers 4 were generally observed in systems containing MMAO-12. In the same systems, oligomers with a methyl starting group were found. It should be noted that the application of other OACs (AlBui3, AlEt3, or AlMe3) in the reaction provided the identical nature of the dimerization and oligomerization products (see, for example, Refs. [26,27,28,31,33,34]). The origin of hydride intermediates in these systems is a result of the β-H elimination process in isobutyl and ethyl groups, as well as in branched alkyl chains bonded with Zr atoms [36].
It was shown that in the system based on zirconocene dichloride and [Ph3C][B(C6F5)4] reacting in toluene, the best result was achieved at a lower content of activator [Ph3C][B(C6F5)4] relative to the Zr complex. Thus, Cp2ZrCl2 catalyzes the oligomerization reaction at the [Zr]:[B] ratio of 1:(0.3–1). When [Zr]:[B] = 1:0.3, the catalytic system was more selective towards dimerization (Table 1, entries 1, 4). An increase in the [Ph3C][B(C6F5)4] concentration in the range of 0.3–1 equiv. resulted in a higher proportion of heavier oligomers with n = 1–5. Simultaneously, the selectivity of the reaction decreased, as follows from the GC-MS and NMR spectra of the products, representing a mixture of stereoisomers and isomers differing in the double bond position. The 13C NMR spectra of the oligomers formed under these conditions exhibited signals for the internal double bond at δC 124–131 ppm. When the reaction was carried out at the [Zr]:[B] ratio of 1:2, alkylation products, hexyl-substituted toluenes, were formed as the major products. An increase in the relative content of zirconocene and HAlBui2 provided the formation of heavier products with bimodal distribution (Mw 5571 and 1578 Da), along with dimers and light oligomers (Table 1, entry 4). In the presence of MMAO-12, this system showed higher activity and selectivity to dimers (entry 5) [31].
In the presence of complexes 1c and 1e with hydrocarbon-bridged cyclopentadienyl ligands, dimer 5 was formed in >94% yield (entries 8, 10, and 14). The activity and chemoselectivity of complex 1e appeared to be higher with activator [Ph3C][B(C6F5)4] (98% dimer) compared to MMAO-12 (78% dimer) (entry 15).
Significant differences in the action of activators were also found when the reaction was carried out in the presence of Me2Si-bound complex 1d (entries 12, 13). For example, the use of MMAO-12 promoted the formation of dimerization product 5 in up to 80% yield, whereas the application of [Ph3C][B(C6F5)4] resulted in the predominant formation of atactic oligomers 6, including heavy oligomers with Mw 6130.
Carrying out the reaction in the presence of bulky pentamethylcyclopentadienyl complex 1b gave mainly oligomer 6, whose mass distribution depended on the type of activator used (entries 6, 7). Under the action of [Ph3C][B(C6F5)4], short-chain oligomers with vinylidene end groups were formed predominantly (13C NMR data). The heavy fraction obtained in the presence of MMAO-12 contained oligomers with a bimodal distribution (Mw 5605 and 861 Da). Moreover, the number of heavy oligomers was much higher in this case than in the experiment with sterically non-hindered Cp2ZrCl2. This indicates a significant predominance of the chain propagation rate over the chain termination rate under these conditions, which may be due to the steric hindrance of the catalytically active sites preventing β-H elimination or chain transfer to the monomer or OAC. Apparently, the steric hindrance is generated by not only the pentamethyl-substituted ligand on Zr but also the bulky methylaluminoxane counterion.
The appearance of indeneyl ligands (1f), in particular bridged ligands (1g–j), in the complexes also led to an increased yield of oligomeric products, both for MMAO-12 and for [Ph3C][B(C6F5)4] (entries 16–33). The reaction catalyzed by Me2CInd2ZrCl2 (1g) mainly afforded dimers and light oligomers with n = 1–4 (entries 23–25). Heavy oligomers were formed in the system based on rac-H4C2Ind2ZrCl2 (1h) and [Ph3C][B(C6F5)4] or MMAO-12 activator (entries 26, 27). Hydrogenation of the ligand benzene rings in complex 1h and the transition to 1j resulted in a considerable decrease in the catalytic system activity: an alkene conversion at the 90–99% level was achieved in 3 h (entries 31–33). Moreover, the oligomeric products obtained in these experiments had, as a rule, a multimodal distribution.
For the isolated higher hexene oligomers, the isotacticity in mmmm% was determined using 13C NMR spectroscopy [37,38,39,40]. The highest stereoselectivity was found for zirconocenes with indenyl ligands. For example, complexes 1f, 1g, and 1h showed isotacticities of 67, 93, and 71% mmmm, respectively (Table 1, entries 20, 23, and 26) (Figures S3 and S4). The stereoselectivity of the catalytic systems based on zirconocenes 1i and 1j was below 58% mmmm. It is noteworthy that the stereoselectivity of this reaction depended on the nature of the activator used. Under the action of complex 1f, an oligomer with isotacticity of 67% was obtained in the presence of MMAO-12, while [Ph3C][B(C6F5)4] furnished an atactic product. The opposite trend was observed for complex 1g with bound ligands: the highest stereoselectivity was attained in the presence of [Ph3C][B(C6F5)4]. The complex H4C2Ind2ZrCl2 (1h) provides the formation of isotactic oligohexenes (61–71% mmmm) regardless of the activator type. For complexes 1i and 1j, some increase in the stereoselectivity was found in the presence of MMAO-12. The observed dependence indicates a fine-tuning of the catalytically active sites in the homogeneous Ziegler–Natta type systems. Probably, the stereocontrol may be provided by not only the ligand on the transition metal atom (site control) or the growing chain (chain control) but also the activator occurring as a part of the active site and determining its structure and dynamics on an equal basis with other factors. As stated earlier [12], the complete separation of the cation–anion pair may increase the activity of the system but may lead to the loss of stereoselectivity. Strong cation–anion interactions could minimize the epimerization of the last inserted unit and thus increase the stereoselectivity, but at the same time, they cause a decrease in the activity by hampering the monomer insertion.
Thus, it was shown that the dimerization pathway is implemented with the participation of Zr complexes with sterically non-hindered cyclopentadienyl ligands (L = Cp, ansa-Me2CCp2, ansa-(Me2C)2Cp2, and ansa-Me2SiCp2), whereas the oligomerization is determined by the action of complexes with bulky cyclopentadienyl (L = C5Me5, rac-H4C2[THInd]2) or electron-withdrawing indenyl (L = Ind, Me2CInd2, H4C2Ind2, BIPh(Ind)2) ligands. The difference between the behaviors of the complexes is primarily determined by the composition and reactivity of the intermediates. It was assumed that hydride bimetallic intermediates may act as the active sites of dimerization and oligomerization reactions [29,30,33,34,35]. Our previous NMR study of the reaction of Cp2ZrCl2 or [Cp2ZrH2]2 with OACs showed the existence of Zr,Zr-hydride complexes, which are activated by MMAO-12 or organoboron compound ([Ph3C][B(C6F5)4] or B(C6F5)3), being thus converted to active species responsible for dimer formation [32,33,34,35].

2.2. NMR Study of the L2ZrCl2–HAlBui2–Activator Catalytic Systems

In order to elucidate the effect of the ligand environment of the transition metal on the structure and reactivity of hydride intermediates, we studied systems consisting of zirconocenes with ansa-cyclopentadienyl and indenyl ligands (1c and 1f), HAlBui2, and MMAO-12 or the ionic activator [Ph3C][B(C6F5)4].
It was found that the composition of the reaction mixture formed in the reaction of zirconocenes with HAlBui2 largely depends on the starting reactant ratio. For example, the 1H NMR spectrum of the mixture of 1c and HAlBui2 taken in the molar ratio [Zr]:[Al] = 1:3 showed hydride signals at δH −1.44 (H1) and 1.29–1.43 ppm (H2). These signals were correlated with each other in the COSY HH experiment (at 250 K) (Figures S5 and S6, Table 2). The H1:HCp:HCp ratio of signal intensities was 1:2:2. The NOESY spectrum exhibited opposite phase cross-peaks of these protons with Cp ring signals at δH 6.11 and 5.19 ppm, and a signal for the methyl group of the isopropylidene bridge at δH 1.03 ppm (Figure S7). At 298 K, the hydride signals were significantly broadened, and peaks corresponding to the chemical exchange between H1, H2, and HAlBui2 monomers and oligomers appeared in the NOESY spectra (Figure S8). Considering these results and literature data for similar molecules with other π-ligands [41,42,43,44], we assigned the observed signals to the trihydride complex 8c as the most probable structure (Scheme 2, Table 2). Moreover, the C2v symmetry of the complex was also in line with the observed NMR pattern.
Under conditions of HAlBui2 deficiency ([Zr]:[Al] = 1:1.7), the spectrum exhibited high-field signals at δH −1.10 (H1) and −0.81 (H2) ppm, correlated with each other in the COSY HH spectrum, in addition to the signals of unreacted complex 1c (Figures S9 and S10). The low-field region of the spectrum contained signals of four non-equivalent protons of the cyclopentadienyl ligand at δH 6.28, 6.23, 5.66, and 5.37 ppm. The observed decrease in the symmetry of the structure, the H1:H2:HCp:HCp:HCp:HCp intensity ratio of 1:1:2:2:2, and literature data [42,44] make it possible to identify the structure as 9c. Additionally, the 1H NMR spectrum showed a low-intensity triplet at δH −3.58 ppm, which was correlated with a doublet at δH −1.34 ppm in the COSY HH spectra (Figure S10). The spin–spin coupling constants 2J = 17.9 Hz for these signals are typical of the bis-zirconium structures [Cp2ZrH2∙Cp2ZrHCl∙ClAlR2], which we detected earlier [32,33,34,35]. Therefore, these resonance lines were attributed to complex 10c.
The addition of MMAO-12 to the mixture obtained in the reaction of Me2CCp2ZrCl2 and HAlBui2 in a ratio of 1:(3–8) and containing the trihydride intermediate 8c induced broadening of both hydride signals and cyclopentadienyl ring signals of complex 8c (Figure 1 and Figure S13), which may indicate its association with methylaluminoxane. According to theoretical data, zirconium hydride complexes have a high affinity for this activator and are able to produce stable adducts [45]. In addition, it is known that methylaluminoxane tends to absorb hydrides, which exchange with methyl groups to provide H-substituted MAO (methylaluminoxane) [46]. This was also confirmed in our study. The NOESY spectra of a mixture of MMAO-12 and HAlBui2 showed a negative effect for methyl groups (−0.14 ppm), isobutyl groups (0.56 and 1.29 ppm), and H-Al(MAO) (3.26–4.11 ppm) (Figures S11 and S12), indicating that these groups belong to the macromolecule. It also follows from the NOESY spectra that the association of 8c with MAO preserves the possibility of hydride exchange with H-Al(MAO) (Figure S14).
Under conditions of HAlBui2 deficiency, when complexes 9c and 10c existed in the catalytic system, the addition of MMAO-12 was accompanied by the appearance of a heavy fraction and by doubling of the triplet at δH −3.58 ppm belonging to 10c, whereas the signals of complex 9c remained virtually unchanged (Figures S15 and S16). The negative NOE indicated an increase in the molecular weight of the complex, which may be due to the formation of a stable 10c∙MAO conjugate similar to the bis-cyclopentadienyl complex [33,35] (Figure S17).
The addition of [Ph3C][B(C6F5)4] to complex 8c formed in the Me2CCp2ZrCl2–HAlBui2 system (1:1) led to an equilibrium shift due to the reaction of HAlBui2 with the organoboron reagent, resulting in the recovery of the original zirconocene dichloride and the appearance of complexes 9c and 10c (Figures S18 and S19). The NMR signals of complex 9c disappeared over time.
In the reaction of Ind2ZrCl2 (1f) with HAlBui2, taken in a ratio of [Zr]:[Al] = 1:(3–8), complex 8f was formed [43]. The complex was characterized by the presence of broadened high-field signals of hydride atoms in the 1H NMR spectra at room temperature, which indicated the participation of the molecule in exchange processes (Figures S21 and S22). The addition of the ionic activator [Ph3C][B(C6F5)4] to 8f, obtained in the 1f-HAlBui2 system (1:3), provided complex 10f and initial 1f by analogy with the ansa-bis-cyclopentadienyl complex (Figure 2 and Figures S23 and S24, Table 2). Moreover, the spectra exhibited a set of indenyl proton signals at δH 4.2–7.8 ppm and hydride atom signals at δH −3–0 ppm, which can be tentatively assigned to the cationic hydride species. Their content increased as the concentration of HAlBui2 increased in the system (Figure 2 and Figure S25).
As in the case of the cyclopentadienyl complex 8c, the addition of MMAO-12 to 8f, obtained in the reaction of 1f with an excess of HAlBui2 (1:3), gave the associate 10f∙MAO, which was characterized by hydride signals at –6.11 and 0.05 ppm in the 1H NMR spectra (Figures S26 and S27). An increase in the content of HAlBui2 to the ratio [Zr]:[Al] = 1:8 in the 1f-HAlBui2-MMAO-12 ([Ph3C][B(C6F5)4]) systems contributed to an increase the intensity of the 1H NMR signals corresponding probably to the cationic species [L2ZrH]+ (Figure 2b and Figures S25 and S28). It should be noted that under conditions of a large excess of HAlBui2 (1:8), the complex 8f did not provide associates with methylalumoxane similar to 8c, which followed from the permanence of the shape and width of NMR spectral lines with the appearance of the activator.
To clarify the reactivity of the complexes towards the alkene, 1-hexene was added to the systems containing various intermediates in different ratios. After the addition of alkene to the reaction media containing an associate of 8c with MAO, hydrometalation products were formed (Figure S20). The formation of dimers was detected only in systems containing intermediate 10c when the reaction mixture was heated to 60 °C for several minutes (Figure 3). It is noteworthy that the reaction catalyzed by the cyclopentadienyl complex 1c at the ratio of reagents [Zr]:[HAlBui2]:[(Ph3C)(B(C6F5)4)] (MMAO-12) = 1:1:0.1(11), when complexes 9c and 10c are present in the catalytic system, was accompanied by a lower conversion of the alkene and the yield of the reaction product 5 (Table 1, entries 9, 10). Apparently, an additional amount of HAlBui2 is required both for the preparation of active forms of co-catalysts [34] and for their subsequent reaction with 10c to give reactive species.
A more pronounced dependency of the reactivity and chemoselectivity of the catalytic system on the composition of the intermediates was observed in the case of the indenyl complex. At a lower content of HAlBui2, when both 10f and cationic species were present in the systems, both dimers and oligomers were found (Table 1, entries 18, 21; Figures S29 and S30). With an increase HAlBui2 concentration and, consequently, cationic intermediates, the proportion of oligomers in the reaction products increased (Table 1, entries 19, 22).
Thus, the NMR study of L2ZrCl2–HAlBui2–activator systems demonstrated the formation of various hydride clusters, among which the bis-zirconium hydride intermediate [L2ZrH2∙L2ZrHCl∙ClAlR2] may be the precursor of the active sites responsible for dimerization.

3. Materials and Methods

3.1. General Procedures

All operations for organometallic compounds were performed under argon according to the Schlenk technique. The complexes Cp2ZrCl2 (1a) [47], Me2CCpZrCl2 (1c) [48], Me2SiCp2ZrCl2 (1d) [49], (Me2C)2CpZrCl2 (1e) [50], Ind2ZrCl2 (1f) [51], rac- Me2C(Ind)2ZrCl2 (1g) [52], and BIPh(Ind)ZrCl2(1i) [53] were synthesized from ZrCl4. Commercially available compounds, including (C5Me5)2ZrCl2 (1b) (97%, Acros), rac-C2H4Ind2ZrCl2 (1h) (Strem), rac-C2H4(THInd)2ZrCl2 (1j) (97%, Merck), HAlBui2 (99%, Merck), MMAO-12 (7% wt Al in toluene, Merck), (Ph3C)[B(C6F5)4] (97%, Abcr), 1-hexene (97%, Acros), were involved into the reactions. The solvents (THF, toluene, benzene) were dried over AlBui3 and distilled immediately prior to use.
CAUTION: pyrophoric nature of aluminum alkyl and hydride compounds require 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). C7D8 and CDCl3 were employed as the solvents and the internal standards. 1D and 2D NMR spectra (COSY HH, HSQC, HMBC, NOESY) were recorded using standard Bruker pulse sequences.
The analysis of the oligo(1-hexenes) by NMR spectroscopy was carried out according to Refs. [37,38,39,40].
The ratio of the light and heavy fractions and the Mw and Mn of oligo(1-hexenes) were determined on a Shimadzu gel permeation chromatograph with a RID-20A detector equipped with a PSS column. Toluene was used as a solvent at a flow rate of 1 mL/min. The measurements were carried out at 25 °C. The standard was polystyrene.
The light products were analyzed using a gas chromatograph-mass spectrometer GCMS-QP2010 Ultra (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).

3.2. Reaction of L2ZrCl2 (1a-j) with HAlBui2, MMAO-12 or (Ph3C)[B(C6F5)4] and 1-Hexene

A flask with a magnetic stirrer was filled under argon with 10 mg (0.018–0.034 mmol) of L2ZrCl2, 0.01–0.018 mL (0.055–0.103 mmol) of HAlBui2, 0.23–0.44 mL (0.55–1.02 mmol) of MMAO-12, 0.002–0.004 mL (0.018–0.034 mmol) of 1-hexene and 0.5 mL of C7D8. For organoboron activator (Ph3C)[B(C6F5)4], the following amounts were used: 10 mg (0.018–0.034 mmol) of L2MCl2, 0.0097–0.018 mL (0.0546–0.103 mmol) of HAlBui2, 4–63 mg (0.0045–0.0684 mmol) of (Ph3C)[B(C6F5)4], 0.002–1.7 mL (0.018–13.7 mmol) of 1-hexene, and 0.5 mL of toluene. The reaction was carried out with stirring at a temperature of 60 °C. After 60 and 180 min, samples (0.1 mL) were syringed into tubes filled with argon, and the samples were decomposed with 10% HCl or DCl at 0 °C. Products were extracted with CH2Cl2, and the organic layer was dried over Na2SO4. The yields of products were determined by GC/MS.

3.3. NMR Study of the Reaction of Me2CCp2ZrCl2 or Ind2ZrCl2 with HalBui2 and Activator (MMAO-12, (Ph3C)[B(C6F5)4])

The NMR tube was charged with 0.029–0.045 mmol of Me2CCp2ZrCl2 or Ind2ZrCl2 and C7D8 in an argon-filled glovebox. The tube was cooled to 0 °C, and 0.03–0.25 mmol (3.8–36 mg) of HAlBui2 was added dropwise. The mixture was stirred, and the formation of complexes 810 was monitored by NMR. Then, 0.17–0.34 mmol (65.4–0.131 mg) of MMAO-12 or 0.005–0.022 mmol (4.7–20 mg) of (Ph3C)[B(C6F5)4] were added.

4. Conclusions

To reveal the regularity of metallocene-AOC-activator catalytic systems’ action, we studied the activity and chemo- and stereoselectivity of η5-zirconium complexes in the presence of HAlBui2, MMAO-12 or [Ph3C][B(C6F5)4] in 1-hexene oligomerization. Depending on the composition of the catalytic system, the formation of 1-hexene vinylidene dimer and oligomers, 1,2-insertion products with different double bond positions, were observed.
The dimerization pathway is implemented, as a rule, in the presence of Zr complexes with sterically non-hindered cyclopentadienyl ligands (L = Cp, ansa-Me2CCp2, ansa-(Me2C)2Cp2, ansa-Me2SiCp2). The use of zirconocenes with bulky cyclopentadienyl (L = C5Me5, rac-H4C2[THInd]2) or electron-withdrawing indenyl (L = Ind, Me2CInd2, H4C2Ind2, BIPh(Ind)2) ligands resulted in the formation of oligomers.
Indenyl complexes (L = Ind, Me2CInd2, H4C2[Ind]2) showed the highest values of isotacticity at 67, 93, 71% mmmm, respectively. Moreover, stereoselectivity was notably dependent on the type of activator, suggesting a significant influence of the co-catalyst on the stereoregulation process in the course of alkene coordination by the catalytically active species.
Considerable differences in catalyst action are indicative of the realization of dimerization and oligomerization processes through different active centers, including hydride intermediates. In the example of the reaction of ansa-Me2CCp2ZrCl2 and Ind2ZrCl2 with HAlBui2 and activators, the composition of hydride intermediates was studied by NMR spectroscopy. As a result, the bis-zirconium complex of type [L2ZrH2∙L2ZrHCl∙ClAlR2], which is the precursor of catalytically active centers of the alkene dimerization, was found.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28062420/s1, Figures S1–S30: NMR spectra of reaction products and intermediates.

Author Contributions

Conceptualization, L.V.P.; methodology, P.V.K.; validation, L.V.P., P.V.K. and A.K.B.; formal analysis, P.V.K.; investigation, P.V.K., A.K.B. and E.R.P.; data curation, L.V.P., P.V.K., P.V.I. and I.E.N.; writing—original draft preparation, P.V.K.; writing—review and editing, L.V.P. and L.M.K.; visualization, L.V.P.; supervision, L.V.P. and L.M.K.; project administration, L.V.P.; funding acquisition, L.V.P. 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 22-23-00818.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

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.

References

  1. Kaminsky, W.; Sinn, H. Methylaluminoxane: Key Component for New Polymerizati on Catalysts, in Polyolefins: 50 Years after Ziegler and Natta II; Advances in Polymer Science; Kaminsky, W., Ed.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 1–28. [Google Scholar]
  2. Janiak, C. Metallocene and Related Catalysts for Olefin, Alkyne and Silane Dimerization and Oligomerization. Coord. Chem. Rev. 2006, 250, 66–94. [Google Scholar] [CrossRef]
  3. Shubkin, R.L.; Kerkemeyer, M.E. Tailor-making polyalphaolefins. J. Synth. Lubr. 1991, 8, 115–134. [Google Scholar] [CrossRef]
  4. Ray, S.; Rao, P.V.C.; Choudary, N.V. Poly-α-olefin-based synthetic lubricants: A short review on various synthetic routes. Lubr. Sci. 2012, 24, 23–44. [Google Scholar] [CrossRef]
  5. Whiteley, K.S.; Heggs, T.G.; Koch, H.; Mawer, R.L.; Immel, W. Polyolefins. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, Germany, 2000. [Google Scholar]
  6. 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]
  7. Nifant’ev, I.; Ivchenko, P. Fair Look at Coordination Oligomerization of Higher α-Olefins. Polymers 2020, 12, 1082. [Google Scholar] [CrossRef]
  8. Kaminsky, W.; Külper, K.; Brintzinger, H.H.; Wild, F.R.W.P. Polymerization of Propene and Butene with a Chiral Zirconocene and Methylalumoxane as Cocatalyst. Angew. Chem. Int. Ed. Engl. 1985, 24, 507–508. [Google Scholar] [CrossRef]
  9. Kaminsky, W.; Niedoba, S.; Möller-Lindenhof, N.; Rabe, O. Isotactic Olefin Polymerization with Optically Active Catalysts. In Catalysis in Polymer Synthesis; American Chemical Society: Washington, DC, USA, 1992; pp. 63–71. [Google Scholar]
  10. Möhring, P.C.; Coville, N.J. Homogeneous group 4 metallocene ziegler-natta catalysts: The influence of cyclopentadienyl-ring substituents. J. Organomet. Chem. 1994, 479, 1–29. [Google Scholar] [CrossRef]
  11. Resconi, L.; Cavallo, L.; Fait, A.; Piemontesi, F. Selectivity in Propene Polymerization with Metallocene Catalysts. Chem. Rev. 2000, 100, 1253–1346. [Google Scholar] [CrossRef]
  12. Xiao, A.; Wang, L.; Liu, Q.; Dong, X. Recent research progress in influence of the ansa-zirconcene catalytic system on the polypropylene microstructure. Des. Monomers Polym. 2007, 10, 281–295. [Google Scholar] [CrossRef] [Green Version]
  13. 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]
  14. 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]
  15. 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]
  16. Desert, X.; Carpentier, J.-F.; Kirillov, E. Quantification of active sites in single-site group 4 metal olefin polymerization catalysis. Coord. Chem. Rev. 2019, 386, 50–68. [Google Scholar] [CrossRef]
  17. Desert, X.; Proutiere, F.; Welle, A.; Den Dauw, K.; Vantomme, A.; Miserque, O.; Brusson, J.-M.; Carpentier, J.-F.; Kirillov, E. Zirconocene-Catalyzed Polymerization of α-Olefins: When Intrinsic Higher Activity Is Flawed by Rapid Deactivation. Organometallics 2019, 38, 2664–2673. [Google Scholar] [CrossRef]
  18. Moscato, B.M.; Zhu, B.; Landis, C.R. Mechanistic Investigations into the Behavior of a Labeled Zirconocene Polymerization Catalyst. Organometallics 2012, 31, 2097–2107. [Google Scholar] [CrossRef]
  19. Sillars, D.R.; Landis, C.R. Catalytic Propene Polymerization:  Determination of Propagation, Termination, and Epimerization Kinetics by Direct NMR Observation of the (EBI)Zr(MeB(C6F5)3)propenyl Catalyst Species. J. Am. Chem. Soc. 2003, 125, 9894–9895. [Google Scholar] [CrossRef] [PubMed]
  20. Christianson, M.D.; Tan, E.H.P.; Landis, C.R. Stopped-Flow NMR: Determining the Kinetics of [rac-(C2H4(1-indenyl)2)ZrMe][MeB(C6F5)3]-Catalyzed Polymerization of 1-Hexene by Direct Observation. J. Am. Chem. Soc. 2010, 132, 11461–11463. [Google Scholar] [CrossRef] [PubMed]
  21. Novstrup, K.A.; Travia, N.E.; Medvedev, G.A.; Stanciu, C.; Switzer, J.M.; Thomson, K.T.; Delgass, W.N.; Abu-Omar, M.M.; Caruthers, J.M. Mechanistic Detail Revealed via Comprehensive Kinetic Modeling of [rac-C2H4(1-indenyl)2ZrMe2]-Catalyzed 1-Hexene Polymerization. J. Am. Chem. Soc. 2010, 132, 558–566. [Google Scholar] [CrossRef]
  22. Sian, L.; Dall’Anese, A.; Macchioni, A.; Tensi, L.; Busico, V.; Cipullo, R.; Goryunov, G.P.; Uborsky, D.; Voskoboynikov, A.Z.; Ehm, C.; et al. Role of Solvent Coordination on the Structure and Dynamics of ansa-Zirconocenium Ion Pairs in Aromatic Hydrocarbons. Organometallics 2022, 41, 547–560. [Google Scholar] [CrossRef]
  23. Baldwin, S.M.; Bercaw, J.E.; Brintzinger, H.H. Alkylaluminum-Complexed Zirconocene Hydrides: Identification of Hydride-Bridged Species by NMR Spectroscopy. J. Am. Chem. Soc. 2008, 130, 17423–17433. [Google Scholar] [CrossRef] [Green Version]
  24. Baldwin, S.M.; Bercaw, J.E.; Brintzinger, H.H. Cationic Alkylaluminum-Complexed Zirconocene Hydrides as Participants in Olefin Polymerization Catalysis. J. Am. Chem. Soc. 2010, 132, 13969–13971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kissin, Y.V. Oligomerization reactions of 1-hexene with metallocene catalysts: Detailed data on reaction chemistry and kinetics. Mol. Catal. 2019, 463, 87–93. [Google Scholar] [CrossRef]
  26. 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]
  27. 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]
  28. 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]
  29. 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] [PubMed] [Green Version]
  30. 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]
  31. Nifant’ev, I.E.; Vinogradov, A.A.; Vinogradov, A.A.; Bagrov, V.V.; Churakov, A.V.; Minyaev, M.E.; Kiselev, A.V.; Salakhov, I.I.; Ivchenko, P.V. A competetive way to low-viscosity PAO base stocks via heterocene-catalyzed oligomerization of dec-1-ene. Mol. Catal. 2022, 529, 112542. [Google Scholar] [CrossRef]
  32. 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]
  33. 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]
  34. 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]
  35. 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. [Google Scholar] [CrossRef] [PubMed]
  36. Chirik, P.J.; Bercaw, J.E. Cyclopentadienyl and olefin substituent effects on insertion and β-hydrogen elimination with group 4 metallocenes. kinetics, mechanism, and thermodynamics for zirconocene and hafnocene alkyl hydride derivatives. Organometallics 2005, 24, 5407–5423. [Google Scholar] [CrossRef]
  37. Asakura, T.; Demura, M.; Nishiyama, Y. Carbon-13 NMR spectral assignment of five polyolefins determined from the chemical shift calculation and the polymerization mechanism. Macromolecules 1991, 24, 2334–2340. [Google Scholar] [CrossRef]
  38. Babu, G.N.; Newmark, R.A.; Chien, J.C.W. Microstructure of Poly(1-hexene) Produced by ansa-Zirconocenium Catalysis. Macromolecules 1994, 27, 3383–3388. [Google Scholar] [CrossRef]
  39. Ahmadjo, S. Preparation of ultra high molecular weight amorphous poly(1-hexene) by a Ziegler–Natta catalyst. Polym. Adv. Technol. 2016, 27, 1523–1529. [Google Scholar] [CrossRef]
  40. Yan, Y.; Wang, D.; He, S.; Ren, H.; Xu, Y. Study on the synthesis of hexene-1 catalyzed by Ziegler-Natta catalyst and polyhexene-1 applications. e-Polymers 2019, 19, 511–518. [Google Scholar] [CrossRef]
  41. 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]
  42. 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]
  43. 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]
  44. Parfenova, L.V.; Vil’danova, R.F.; Pechatkina, S.V.; Khalilov, L.M.; Dzhemilev, U.M. New effective reagent [Cp2ZrH2·ClAlEt2]2 for alkene hydrometallation. J. Organomet. Chem. 2007, 692, 3424–3429. [Google Scholar] [CrossRef]
  45. Tyumkina, T.V.; Islamov, D.N.; Kovyazin, P.V.; Parfenova, L.V. Chain and cluster models of methylaluminoxane as activators of zirconocene hydride, alkyl and metallacyclopropane intermediates in alkene transformations. Mol. Catal. 2021, 512, 111768. [Google Scholar] [CrossRef]
  46. Babushkin, D.E.; Panchenko, V.N.; Timofeeva, M.N.; Zakharov, V.A.; Brintzinger, H.H. Novel Zirconocene Hydride Complexes in Homogeneous and in SiO2-Supported Olefin-Polymerization Catalysts Modified with Diisobutylaluminum Hydride or Triisobutylaluminum. Macromol. Chem. Phys. 2008, 209, 1210–1219. [Google Scholar] [CrossRef] [Green Version]
  47. 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]
  48. Koch, T.; Blaurock, S.; Somoza, F.B.; Voigt, A.; Kirmse, R.; Hey-Hawkins, E. Unexpected P−Si or P−C Bond Cleavage in the Reaction of Li2[(C5Me4)SiMe2PR] (R = Cyclohexyl, 2,4,6-Me3C6H2) and Li[(C5H4)CMe2PHR] (R = Ph, tBu) with ZrCl4 or [TiCl3(thf)3]:  Formation and Molecular Structure of the ansa-Metallocenes [{(η-C5Me4)2SiMe2}ZrCl2] and [{(η-C5H4)2CMe2}MCl2] (M = Ti, Zr). Organometallics 2000, 19, 2556–2563. [Google Scholar]
  49. Bajgur, C.S.; Tikkanen, W.; Petersen, J.L. Synthesis, structural characterization, and electrochemistry of [1]metallocenophane complexes, [Si(alkyl)2(C5H4)2]MCl2, M = Ti, Zr. Inorg. Chem. 1985, 24, 2539–2546. [Google Scholar] [CrossRef]
  50. Schwemlein, H.; Brintzinger, H.H. ansa-Metallocene derivatives: V. Synthesis of tetramethylethylene-bridged titanocene and zirconocene derivatives via reductive fulvene coupling. J. Organomet. Chem. 1983, 254, 69–73. [Google Scholar] [CrossRef] [Green Version]
  51. Piccolrovazzi, N.; Pino, P.; Consiglio, G.; Sironi, A.; Moret, M. Electronic effects in homogeneous indenylzirconium Ziegler-Natta catalysts. Organometallics 1990, 9, 3098–3105. [Google Scholar] [CrossRef]
  52. Nifant’ev, I.E.; Ivchenko, P.V. Synthesis of Zirconium and Hafnium ansa-Metallocenes via Transmetalation of Dielement-Substituted Bis(cyclopentadienyl) and Bis(indenyl) Ligands. Organometallics 1997, 16, 713–715. [Google Scholar] [CrossRef]
  53. Ellis, W.W.; Hollis, T.K.; Odenkirk, W.; Whelan, J.; Ostrander, R.; Rheingold, A.L.; Bosnich, B. Synthesis, structure, and properties of chiral titanium and zirconium complexes bearing biaryl strapped substituted cyclopentadienyl ligands. Organometallics 1993, 12, 4391–4401. [Google Scholar] [CrossRef]
Scheme 1. Alkene transformations in the presence of L2ZrCl2–HAlBui2–activator catalytic systems.
Scheme 1. Alkene transformations in the presence of L2ZrCl2–HAlBui2–activator catalytic systems.
Molecules 28 02420 sch001
Scheme 2. Reaction of Me2CCp2ZrCl2 (1c) or Ind2ZrCl2 (1f) with HAlBui2, activators, and alkene.
Scheme 2. Reaction of Me2CCp2ZrCl2 (1c) or Ind2ZrCl2 (1f) with HAlBui2, activators, and alkene.
Molecules 28 02420 sch002
Figure 1. 1H NMR spectrum of the Me2CCp2ZrCl2 (1c)−HAlBui2−MMAO-12 system in C7D8 (298 K): (a) [Zr]:[Al]:[AlMAO] = 1:8:0; (b) [Zr]:[Al]:[AlMAO] = 1:8:11.
Figure 1. 1H NMR spectrum of the Me2CCp2ZrCl2 (1c)−HAlBui2−MMAO-12 system in C7D8 (298 K): (a) [Zr]:[Al]:[AlMAO] = 1:8:0; (b) [Zr]:[Al]:[AlMAO] = 1:8:11.
Molecules 28 02420 g001
Figure 2. 1H NMR spectrum of the Ind2ZrCl2 (1f)−HAlBui2−[Ph3C][B(C6F5)4] system in C7D8 (298 K): (a) [Zr]:[Al]:[B] = 1:3:0.2; (b) [Zr]:[Al]:[B] = 1:8:0.2.
Figure 2. 1H NMR spectrum of the Ind2ZrCl2 (1f)−HAlBui2−[Ph3C][B(C6F5)4] system in C7D8 (298 K): (a) [Zr]:[Al]:[B] = 1:3:0.2; (b) [Zr]:[Al]:[B] = 1:8:0.2.
Molecules 28 02420 g002
Figure 3. NMR monitoring of the reaction of hydride complexes with 1-hexene in C6D5CD3 (intensity of high-field signals is increased): (a) Me2CCp2ZrCl2 (1c)−HAlBui2–(Ph3C)[B(C6F5)4], [Zr]:[Al]:[B] = 1:1:0.1; (b) [Zr]:[Al]:[B]:[1-alkene] = 1:1:0.1:1.5, 5 min; (c) [Zr]:[Al]:[B]:[1-alkene] = 1:1:0.1:1.5, after keeping the tube at 60 °C for 5 min.
Figure 3. NMR monitoring of the reaction of hydride complexes with 1-hexene in C6D5CD3 (intensity of high-field signals is increased): (a) Me2CCp2ZrCl2 (1c)−HAlBui2–(Ph3C)[B(C6F5)4], [Zr]:[Al]:[B] = 1:1:0.1; (b) [Zr]:[Al]:[B]:[1-alkene] = 1:1:0.1:1.5, 5 min; (c) [Zr]:[Al]:[B]:[1-alkene] = 1:1:0.1:1.5, after keeping the tube at 60 °C for 5 min.
Molecules 28 02420 g003
Table 1. Catalytic activity and chemoselectivity of L2ZrCl2–HAlBui2–activator systems in the reaction with 1-hexene in toluene, 60 °C.
Table 1. Catalytic activity and chemoselectivity of L2ZrCl2–HAlBui2–activator systems in the reaction with 1-hexene in toluene, 60 °C.
EntryCatalytic SystemsMole Ratio
[Zr]:
[HAlBui2]: [Activator]:
[1-Alkene]
Time, MinAlkene Conver-sion, %Light Fraction Yield, wt.%Light Fraction Product Composition, % (GC-MS)Heavy Fraction Yield, wt.% (GPC)MW, DaMN, DaMW/
MN
Tacticity, mmmm%Oligomer Structural Type
ComplexActivator56
n = 1n = 2n = 3n = 4n = 5
1Cp2ZrCl2
(1a)
(Ph3C)[B(C6F5)4]1:3:0.3:40060959861211062-2621161291.01atactic6bd
21:3:1:400>99993028201084-
31:3:2:400>99 a 33527752081.01
44:16:1:400>9988615128852557154591.02
10157812111.30
5MMAO-12 [31]1:3:30:400598 97 -
6(C5Me5)2ZrCl2
(1b)
(Ph3C)[B(C6F5)4]4:16:1:4001809944027183024 5511099331.19atactic6a
7MMAO-121:3:30:400609932823171616 71560559061.05atactic4, 6a d
268617981.08
8Me2CCp2ZrCl2
(1c)
(Ph3C)[B(C6F5)4]4:16:1:40060>99 954 -
91:1:0.1:40018091 865
10MMAO-121:3:30:40060>99 946 -
111:1:11:40018077 752
12Me2SiCp2ZrCl2 (1d)(Ph3C)[B(C6F5)4]4:16:1:400180>9951728152118 89613060411.02atactic6a, 6d
6276926721.04
13MMAO-121:3:30:40060>999980164 -
14(Me2C)2Cp2ZrCl2 (1e)(Ph3C)[B(C6F5)4]4:16:1:40060>9999981 -
15MMAO-121:3:30:4006085 787 -
16Ind2ZrCl2
(1f)
(Ph3C)[B(C6F5)4]1:3:0.3:400180>9996723322117 3630762001.02atactic6ad
174:16:1:40060>9963022202631 369888691.14atactic6a
18 e1:3:0.2:51074 f 121616
19 e1:8:0.2:8010>99 g 1316161098
20MMAO-121:3:30:40060>9927615171943 72218215991.36isotactic (67%)4, 6a,d
21 e1:3:7:310>99 h 468
22 e1:8:10:8010>99 i 27221784
23rac-Me2CInd2ZrCl2
(1g)
(Ph3C)[B(C6F5)4]1:3:0.3:4006097702221201918 27635962911.01isotactic (93%)6d
244:16:1:40060>99727910542 24558551441.08isotactic (76%)6ad
39038711.04
25MMAO-121:3:30:40060>9955433331812 4410819271.17atactic4, 6ad d
26rac-H4C2Ind2ZrCl2
(1h)
(Ph3C)[B(C6F5)4]4:16:1:40060>99 99639662851.02isotactic (71%)6ad
27MMAO-121:3:30:40060>995722222425 94634762871.01isotactic (61%)4, 6a,d
28BIPh(Ind)2ZrCl2
(1i)
(Ph3C)[B(C6F5)4]1:3:0.3:40018047 b2513492315 22603761251.01
294:16:1:40060>99 b59721261432 40109610031.09isotactic (33%)6a
30MMAO-121:3:30:40060>99 b9320252329 20574256571.02isotactic (45%)4, 6a d
70215915321.4
31rac-H4C2[THInd]2ZrCl2
(1j)
(Ph3C)[B(C6F5)4]1:3:0.3:40018089 c6913182015141210613160481.02
109609091.06
324:16:1:400180>99112919181717 88553050331.10isotactic (38%)6ad
33MMAO-121:3:30:4001809581124242318 38615961031.01isotactic (58%)6ad
19314531141.01
30207419711.06
a Monoalkyl-substituted toluene (77%), dialkyl-substituted toluene (19%), and trialkyl-substituted toluene (4%) were detected; b tetramers, that is, “dimer of dimers” (15%), were formed in the presence of both activators; c the yield of oligomers with n = 6 were up to 7%; d Me starting group; e reaction was carried out in an NMR tube at 20 °C; f hydrometalation product (8%), oligomer 4 (m = 1: 15%, m = 2: 7%, R = H); g hydrometalation product (4%), oligomer 4 (m = 1: 18%, R = H); h hydrometalation product (22%), oligomer 4 (m = 1: 24%, R = H); i hydrometalation product (4%), oligomer 4 (m = 1: 18%, R = H).
Table 2. 1H and 13C NMR (δ, ppm, 400.13 MHz (1H), 100.62 MHz (13C)) of hydride complexes.
Table 2. 1H and 13C NMR (δ, ppm, 400.13 MHz (1H), 100.62 MHz (13C)) of hydride complexes.
ComplexActivatorT, KδH CpδC CpδH H1δH H2
8c 250 K6.11 (s, 4H)
5.19 (s, 4H)
1.03 (s, 12H)
110.9
110.3
100.8
36.8
22.3
−1.44
(br.s, 2H)
1.29–1.43
(br.s, 1H)
9c 6.28 (s, 2H)
6.23 (s, 2H)
5.66 (s, 2H)
5.37 (s, 2H)
1.15 (NOESY)
1.12 (NOESY)
−1.10
(br.s, 1H)
−0.81
(br.s, 1H)
10c 298 K −3.58 (t, 17.9 Hz, 1H)−1.34 (d, 17.9 Hz, 2H)
8c∙MAOMMAO−12
298 K6.14 (NOESY)
4.93–5.69 (br.m)
−1.62− −1.22 (br.s)1.25–1.48 (COSY HH)
−0.38 − −0.02 (MAO)
3.26–4.11
(H-MAO)
10c∙MAOMMAO−12298 K6.12 (NOESY)
5.67 (NOESY)
5.42 (NOESY)
4.85 (NOESY)
0.98 (NOESY)
119.6
108.7
101.2
98.2
21.1
−3.57
(t, 17.6 Hz, 2/3H)
−3.71 (t, 18.0 Hz, 1/3H)
−1.32 (d, 17.6 Hz, 2H)

−0.56 − 0.15 (MAO)
10c a[Ph3C][B(C6F5)4] 298 K −3.59 (t, 18.3 Hz, 2H)−1.34 (d, 18.3 Hz, 2H)
8f [43] 220 K7.37 (m, 4H)
6.87 (m, 4H)
6.32 (m, 2H)
5.19 (m, 4H)
−1.11 (d, 5.6 Hz, 2H)0.07 (t, 7.6 Hz, 1H)
8f 292 K7.43 (m, 4H)
6.89 (m, 4H)
6.38 (m, 2H)
5.36 (m, 4H)
−0.92
(br.s, 2H)
0.07
(NOESY)
10f∙MAOMMAO−12298 K5.92 (m, 4H)
5.08 (m, 4H)
4.62 (m, 4H)
−6.11 (t, 18.5 Hz, 1H)0.05
(COSY HH)
10f a[Ph3C][B(C6F5)4]298 K7.49 (d, 8.3 Hz, 4H)
7.19 (d, 8.3 Hz, 4H)
6.99–7.16 (m, 8H)
5.82 (br.t, 3.0 Hz, 4H)
5.02 (m, 4H)
4.57 (m, 4H)
127.9
126.6
119.0
117.3
99.1
96.2
−6.00 (t, 18.1 Hz, 1H)−0.04 (d, 18.1 Hz, 2H)
a The complex was observed after the addition of [Ph3C][B(C6F5)4].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Parfenova, L.V.; Kovyazin, P.V.; Bikmeeva, A.K.; Palatov, E.R.; Ivchenko, P.V.; Nifant’ev, I.E.; Khalilov, L.M. Catalytic Properties of Zirconocene-Based Systems in 1-Hexene Oligomerization and Structure of Metal Hydride Reaction Centers. Molecules 2023, 28, 2420. https://doi.org/10.3390/molecules28062420

AMA Style

Parfenova LV, Kovyazin PV, Bikmeeva AK, Palatov ER, Ivchenko PV, Nifant’ev IE, Khalilov LM. Catalytic Properties of Zirconocene-Based Systems in 1-Hexene Oligomerization and Structure of Metal Hydride Reaction Centers. Molecules. 2023; 28(6):2420. https://doi.org/10.3390/molecules28062420

Chicago/Turabian Style

Parfenova, Lyudmila V., Pavel V. Kovyazin, Almira Kh. Bikmeeva, Eldar R. Palatov, Pavel V. Ivchenko, Ilya E. Nifant’ev, and Leonard M. Khalilov. 2023. "Catalytic Properties of Zirconocene-Based Systems in 1-Hexene Oligomerization and Structure of Metal Hydride Reaction Centers" Molecules 28, no. 6: 2420. https://doi.org/10.3390/molecules28062420

Article Metrics

Back to TopTop