Next Article in Journal
Single and Multiple Gate Design Optimization Algorithm for Improving the Effectiveness of Fiber Reinforcement in the Thermoplastic Injection Molding Process
Next Article in Special Issue
Polyisobutylenes with Controlled Molecular Weight and Chain-End Structure: Synthesis and Actual Applications
Previous Article in Journal
Assessing the Effect of Cellulose Nanocrystal Content on the Biodegradation Kinetics of Multiscale Polylactic Acid Composites under Controlled Thermophilic Composting Conditions
Previous Article in Special Issue
Research of Potential Catalysts for Two-Component Silyl-Terminated Prepolymer/Epoxy Resin Adhesives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MAO- and Borate-Free Activating Supports for Group 4 Metallocene and Post-Metallocene Catalysts of α-Olefin Polymerization and Oligomerization

by
Ilya E. Nifant’ev
1,2,
Pavel D. Komarov
1,
Oksana D. Kostomarova
3,
Nikolay A. Kolosov
3 and
Pavel V. Ivchenko
1,2,*
1
A.V. Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences, Leninsky Av. 29, 119991 Moscow, Russia
2
Chemistry Department, M.V. Lomonosov Moscow State University, Leninskie Gory 1-3, 119991 Moscow, Russia
3
NIOST LLC, Kuzovlevsky Tr. 2-270, 634067 Tomsk, Russia
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(14), 3095; https://doi.org/10.3390/polym15143095
Submission received: 13 June 2023 / Revised: 15 July 2023 / Accepted: 17 July 2023 / Published: 19 July 2023
(This article belongs to the Special Issue Catalytic Applications in Polymerization)

Abstract

:
Modern industry of advanced polyolefins extensively uses Group 4 metallocene and post-metallocene catalysts. High-throughput polyolefin technologies demand the use of heterogeneous catalysts with a given particle size and morphology, high thermal stability, and controlled productivity. Conventional Group 4 metal single-site heterogeneous catalysts require the use of high-cost methylalumoxane (MAO) or perfluoroaryl borate activators. However, a number of inorganic phases, containing highly acidic Lewis and Brønsted sites, are able to activate Group 4 metal pre-catalysts using low-cost and affordable alkylaluminums. In the present review, we gathered comprehensive information on MAO- and borate-free activating supports of different types and discussed the surface nature and chemistry of these phases, examples of their use in the polymerization of ethylene and α-olefins, and prospects of the further development for applications in the polyolefin industry.

1. Introduction

Polyolefins still remain the most valuable plastics, with an annual production of ~200 Mt [1]. Most polyolefins are manufactured using Ziegler-Natta titanium catalysts and Phillips chromium catalysts [2,3,4,5,6]. Group 4 metal single-site metallocene and post-metallocene catalysts outperform industrial catalysts on the parameters of activity and homogeneity of ethylene/α-olefin copolymers [2,3,4,7,8]. The advantages of single-site Group 4 catalysts can be primarily attributed to the nature of the catalytic species that represent cationic alkyl-metal centers in basic ligand environments, which have a high ability to π-complexation and insertion of α-olefin molecule [9] (Scheme 1a). Active sites of highly efficient Group 4 metal catalysts are stabilized by weakly coordinating anions. These anions are formed during the activation of pre-catalysts by methylalumoxane (MAO) [10,11] or by trialkylaluminums in combination with B(C6F5)3 or perfluoroaryl borates [9,11,12,13] (Scheme 1b) and other, less commonly used, reagents [14,15,16,17]. High catalytic performance of single-site Group 4 metal catalysts is usually observed when using 103–104 molar equivalents of MAO; perfluoroarylborates can be applied in equal amounts [11,18]. Among trialkylaluminums, Me3Al can act as catalyst poison, especially in the case of zirconocenes, due to the formation of dormant L2Zr-(μ-Me)2-AlMe2+ species [19,20,21], and one of the functions of excess of MAO under homogeneous conditions is to be a ‘Me3Al sponge’. When using zirconocenes and perfluoroaryl borates, activation mechanisms and the structure of catalytic species essentially depend on the type of R3Al. In particular, pre-activation of LMCl2 by iBu3Al results in the formation of heterometallic hydrides [22,23,24,25], which form cationic species during reaction with borates [26,27]; in the case of Et3Al, metallacycles can be formed [28,29] (Scheme 1c). The type of borate activator can also have an impact on the activity, e.g., activation by [PhNMe2H][B(C6F5)4] is followed by the formation of PhNMe2 capable of coordination at the catalytic center [20]. An important aspect of homogeneous iBu3Al/perfluoroaryl borate activation is the reaction of iBu3Al with borate. As early as 1998, catalyst deactivation caused by the reaction of [CPh3][B(C6F5)4] with iBu3Al was observed [30]; this process leads to CPh3H, isobutylene, and highly reactive iBu2Al+ [31]. When using iBu3Al/[PhNHMe2][B(C6F5)4], isobutane was detected in reaction products [32]. Therefore, the mechanism of activation by iBu3Al requires considering the direct participance of cationic iBu2Al+ species, which were isolated and characterized recently [33].
The use of a large excess of organoaluminum compounds often leads to hydroalumination [34,35], formation of inactive Zr–CH2CH2–Al species [36] (iBu3Al), and cycloalumination (Et3Al) [37,38] side processes. The problems of side reactions during homogeneous activation can be resolved during the adaptation of Group 4 catalysts to actual polyolefin technologies that are tuned for use in gas-phase or slurry processes and require supported catalysts [39]. Heterogenization of single-site catalysts was and is a crucial challenge for the polyolefin industry, and this problem was intensively studied over the past three decades [39,40,41,42,43]. The most evident way to solve this problem is based on the reaction of MAO with active support (silica, alumina, etc.) [44,45,46,47,48,49,50,51] or on functionalization of the support by perfluoroaryl borate or aluminate fragments [52,53,54] with a formation of negatively charged particles that provide binding with catalytically active cationic species through Coulomb attraction (Scheme 2a); stabilization of the catalyst occurs through the same mechanism as for homogeneous catalysts [39,40,44]. The studies of immobilization by covalent binding of the ligand precursors with the support surface (Scheme 2b) were also carried out with varying results, this type of heterogenization also requires MAO or borate activators [39,55,56,57,58,59]. Finally, a number of research groups investigated acidic supports that can react with Group 4 metal complexes with a formation of active cationic catalytic species (Scheme 2c) [46,60]. Activating supports are particularly attractive since they do not require MAO or borates (only R3Al are needed as alkylating agents and scavengers), thereby reducing the overall cost of the catalysts.
The very idea of the activating supports for Group 4 metallocenes and post-metallocenes is based on early studies of the activation of homoleptic alkyl derivatives of Ti(IV) and Zr(IV) by Al2O3 and SiO2; the activating effect of metal oxides was attributed to the reaction of surface–OH groups of the support (S) with MR4 that resulted in the formation of (S)(–O)2MR2 species (Scheme 3a) [61,62,63,64,65,66]. Note that the chemical nature of catalytically active (≡Si–O)2ZrBn2 species, formed on SiO2 surface, was confirmed experimentally just in 2013 [67]. Activation of bis(arene) Ti(0) and Zr(0) complexes by Al2O3 presumably proceeds via oxidative addition of a surface O–H bond to give a divalent supported species (Scheme 3b) [65,66,67,68,69].
The first experimental proof of the ability of support to activate metallocenes with a formation of highly active cationic catalytic species according to Scheme 2c comes from the studies of Marks and colleagues [70,71,72]. In 1985, they investigated the reaction of (η5-C5Me5)2ThMe2 with dehydroxylated γ-Al2O3 (DA); the transfer of methyl groups from Th to surface Al sites with a formation of cationic species was detected using 13C cross-polarization/magic angle spinning (CP MAS) NMR [70]. In 1988, they showed that DA is an efficient activator for (η5-C5H5)2ZrMe2, (η5-C5Me5)2ZrMe2, and (η5-C5Me5)ZrMe3 in ethylene polymerization [71]; Zr→Al migrations of Me group were also confirmed by 13C CP/MAS NMR experiments [65,69].
In the early 1990s, Soga and Kaminaka [73,74,75] studied the propylene polymerization activity of different zirconocene dichlorides activated by AlR3, with SiO2, Al2O3, MgCl2, MgF2, CaF2, or AlF3 supports. The absence of activity, observed for SiO2 and MgO, was attributed to a lack of acidity of these oxides compared with the other materials. MgCl2 was then studied as an activating support for Ti(IV)-based post-metallocenes [76,77,78,79] and zirconocenes [73,80,81]. Since unmodified silica and alumina provided only partial activation of metal complexes [82], in the current century, the focus was on sulfated metal oxides, studied mainly by the scientific group of Marks [83,84,85,86,87], and fluorinated metal oxides, studied by the scientists from Elf [88], Chevron Phillips [89,90,91,92,93,94,95,96,97,98], INEOS [99,100,101], and Total in cooperation with the scientific group of Boisson [102,103,104,105,106,107].
The use of activating supports in the catalytic chemistry of α-olefins has been discussed in a number of review papers and book chapters. The first promising results in the surface chemistry of sandwich complexes of actinides and Group 4 metals have been discussed by Marks and Eisen in the early 1990s [72,108]. Some issues regarding metallocene activation by silica [109], and the difference in activating chemistry and efficiency between silica and alumina [110], were discussed by Copéret et al. The importance of achieving supported catalyst structural uniformity was also demonstrated in a review from Wegener et al [111]. The chapters, devoted to MgCl2-supported single-site catalysts [112] and activation of metallocenes by solid acids [113], were published in the collection ‘Tailor-Made Polymers, Via Immobilization of Alpha-Olefin Polymerization Catalysts’ (2008). In 2018, activating supports for metallocenes and related catalysts were briefly discussed by Hoff in a chapter of the ‘Handbook of Transition Metal Polymerization Catalysts’ [114]. Activating supports were also mentioned in recent reviews on supported catalysts [41,42,43,46,60].
In the present work, we summarized and analyzed published data on activating supports of different types and compared mechanisms and efficiency of Group 4 metallocene and post-metallocene activation and reactivity. The prospects of the use of activating supports in the polyolefin industry are also discussed.

2. Metal Halides

2.1. MgCl2

MgCl2 is a widely used support for conventional titanium–magnesium Ziegler–Natta catalysts (TMZNCs) of third and subsequent generations [115,116,117,118,119,120]. In view of the industrial importance of TMZNCs, efficient methods for controlling the particle size, porosity, and morphology of MgCl2 have been developed, and it is not surprising that MgCl2 was studied as a support for Group 4 metal single-site catalysts [39,112]. MgCl2 supports may also have the advantage of easier fragmentation than silica or alumina supports, thereby facilitating polymerization and improving the morphology of polymer particles formed. Whilst TMZNCs are not the subject of our review, the role of MgCl2 as an activating support for absorbed Ti catalytic species is not in doubt. According to current theoretical concepts about the nature of TMZNCs, pre-catalysts represent TiCl4 adsorbed on MgCl2 crystallites, and binding strength depends on the crystal structure of the MgCl2 support. In their fundamental works, Cavallo and colleagues assumed the MgCl2 bulk to be in the α crystalline phase with the surface comprising (104) and (110) lateral cuts that have Mg centers with coordination numbers (CNMg) of 5 and 4, respectively (Figure 1) [116,121].
In terms of the reaction mechanism, the method of the activation, i.e., the introduction of R3Al after or before/simultaneously with the interaction of the pre-catalyst with MgCl2, could be essential. In the latter case, the chemistry of R3Al on the MgCl2 surface is an important point in examining the activation of Group 4 metal complexes. Using 27Al MAS NMR spectroscopy, Potapov and colleagues have studied organoaluminum compounds EtnAlCl3−n supported on a highly dispersed MgCl2, prepared by the reaction of Mg with n-BuCl [122]. In the 27Al MAS NMR spectrum of the sample Et2AlCl/MgCl2 they detected a characteristic signal of Al species with coordination number (CNAl) = 5. An attempt to record the 27Al MAS NMR spectrum of Et3Al/MgCl2 failed; however, analysis of the spectrum obtained after treatment with Bu2O allowed to propose the formation of dimeric Et2Al(μ-Et)2AlEt2 species on the MgCl2 surface (CNAl = 5). However, the reaction of MgCl2 with R3Al can not be called fully studied; a lot of research on TMC activation was focused on R3Al interactions with Ti species [123,124] without analyzing concurrent processes on the MgCl2 surface.
In the 1990s, a number of MgCl2-supported Group 4 metal complexes were studied in polymerization experiments after activation by alkylaluminum compounds. Kaminaka and Soga activated rac-[1,2-ethylenebis(η5-4,5,6,7-tetrahydroinden-1-yl)]ZrCl2 (rac-[EBTHI]ZrCl2, Zr1) by R3Al (R = Me, Et) in the presence of MgCl2 [73]. The supported catalyst was prepared by the grounding of the support and Zr1 in toluene in the presence of R3Al. In propylene polymerization, supported catalysts were significantly inferior to homogeneous Zr1/MAO by activity, but at the same time, polypropylene (PP), formed when using a supported catalyst, had higher isotacticity. During a comparative study of (η5-C5H5)2ZrCl2 (Cp2ZrCl2, Zr2) and [Me2C(η5-fluoren-9-yl)(η5-C5H4)]ZrCl2 (Me2C(Flu)(Cp)ZrCl2, Zr3), activated by Me3Al in the presence of commercial MgCl2, atactic and syndiotactic polypropylenes were obtained; both zirconocenes have demonstrated high activities comparable with the activity of MAO-activated complexes [74]. In further studies, the mechanism of the activating of Cp2ZrX2 (X = Cl, Me, Et) by Zr→Mg transfer of X with a formation of Cp2ZrX+ was proposed [75]. Note that with an increase of the Al:Zr ratio, catalytic activity goes through the maximum; this was attributed to R3Al binding with the MgCl2 surface. The binding of Cp2Zr fragments with MgCl2 surface is rather weak: so, for example, Ochędzan-Siodłak and Nowakowska showed that the addition of MAO to the Zr2/MgCl2/R3Al system results in the formation of a homogeneous catalyst [125].
Soga and colleagues have studied the catalytic behavior of CpTiCl3 (Ti1), supported on MgCl2 and activated by iBu3Al, in the polymerization of propylene [126]. Surprisingly, when using a MgCl2-supported catalyst, isotactic PP was obtained. However, in a later study, Kang and colleagues reported a lack of activity when using CpMCl3/MgCl2 (M = Ti, Zr) pre-catalysts prepared in the presence of 2-ethylhexanol [127].
Bis- and mono-cyclopentadienyl Ti(IV) complexes Cp2TiCl2 (Ti2) and Ti1 were supported on MgCl2 by physical (grinding) or chemical (in the presence of Et3Al) methods [128]. UV-vis spectra showed coordination of Cl ligands of Ti1 and Ti2 to surface Mg Lewis acidic sites (LAS), which were more pronounced for Ti1. After activation by Et3Al or MAO, supported catalysts were studied in propylene polymerization. For both Ti complexes, isotactic PP was obtained, and the authors attributed this fact to the formation of Ti(III) catalytic species.
The first study of the activation of Group 4 post-metallocenes using MgCl2 and different alkylaluminums in propylene polymerization was conducted by Soga and colleagues in 1997 [129] on the example of β-diketonate L2TiCl2 complexes Ti3Ti6 (Scheme 4a). Pre-catalysts were prepared by the reaction of Ti3Ti6 with MgCl2 in toluene, and the activation by alkylaluminums was performed in n-heptane. The polymerization rate when using Ti3 decreased in the following order: MAO > Me3Al > Et3Al > iBu3Al > Et2AlCl from 300 to 17 kg∙mol−1∙h−1; mainly atactic PP was formed in all experiments. Ethylene/propylene copolymerization on the example of Ti6 as a pre-catalyst with and without diisopropyldimethoxysilane as an ‘external donor’ was studied the following year [76]. The nature of the catalytic species, formed by Ti3Ti6 after activation, remained unclear, and the fact that the activating support mechanism is applicable to these catalysts did not.
Bis(phenoxy-imine) Group 4 metal complexes (named FI catalysts), combined with MAO or borate activators, display unique polymerization activities in α-olefin polymerization [130,131,132,133,134,135]. Because of the high productivity of FI catalysts and the broad spectrum of accessible polymers, there has been significant interest in developing supported catalysts for industrial applications. The use of harmless and cost-competitive MgCl2 as an activating support seemed a profitable and promising idea. In 2004, the Fujita group reported the results of the study of catalytic behavior of the complexes Ti7 and Zr4 (Scheme 4b), activated by Et3Al in the absence and in the presence of mechanically pulverized MgCl2, in the polymerization of ethylene. At 50 °C and an ethylene pressure of 0.9 MPa, no polymer formed in the absence of MgCl2, whereas in the presence of MgCl2, Ti7 and Zr4 demonstrated activities of 0.32 and 9.10 kg∙mmol−1∙bar−1∙h−1 [79]. The authors proposed that MgCl2 acts as a Lewis acid to form a cationic active species from the ethylated FI catalysts and becomes an anionic species stabilizing cationic active species. Under identical conditions, metallocenes Zr2 and Ti2 provided only a trace amount of PE (<0.1 kg∙mmol−1∙bar−1∙h−1).

2.2. MgCl2/THF Adducts

Supported Cp2TiCl2/MgCl2 pre-catalyst, prepared by the precipitation of Ti2 and MgCl2 solution in THF into pentane, polymerized ethylene after activation by iBu3Al or Et2AlCl, and no activity was detected after treatment of Ti2 by iBu3Al in the absence of MgCl2 [136]. Et2AlCl was found to be a more efficient activator in comparison with iBu3Al, and the catalysts showed a steady-state kinetic behavior. The authors explained these results by Ti→Mg transfer of the alkyl group with a formation of active Cp2TiR+ species.
The MgCl2(THF)2/Et2AlCl phase was proposed as a support for Zr2 by Nowakowska and colleagues [125,137], who assumed the formation of the catalyst precursor Cp2Zr(μ-Cl)(μ-Et)Al(μ-Cl)(μ-THF)Mg on the surface of MgCl2(THF)2, which was activated by MAO without desorption. Activation of this catalyst by R3Al was not studied.

2.3. MgCl2/ROH Adducts

The products of the reaction of MgCl2/ROH adducts with alkylaluminum compounds are another type of MgCl2-based activating supports. Evidently, their chemical nature is qualitatively different from the chemical nature of the product of the reaction of MgCl2 with R3Al.
Cho and Lee prepared adducts MgCl2·4MeOH and MgCl2·3.33EtOH, and thermal treatment of the former complex resulted in the formation of MgClx(OMe)y species [138]. Active supports were obtained by the treatment of the adducts with Me3Al, Et3Al, iBu3Al, or MAO (comparative experiments). A more efficient binding with alkylaluminums was observed for MgCl2·4MeOH due to the formation of Mg–(μ-OMe)–Al species. The proposed impregnation mechanism for Zr2, based on 27Al NMR spectral data, is presented in Scheme 5. Supported catalysts were studied in ethylene/hex-1-ene copolymerization after activation by MAO; therefore, the results of the research MgCl2(ROH)x/R3Al system cannot be regarded as a full-fledged activating support.
Subsequent studies of the use of MgCl2 supports obtained by the reaction of MgCl2/EtOH adducts with AliBu3, AlEt3, or AlEt2Cl have demonstrated high efficiency of iBu3Al and Et3Al pre-treated supports in activation of Ti2, Zr2, and Me2Si(η5-C5Me4)(NtBu)TiCl2 (Ti8) in the absence of MAO and borates [80]. In the homopolymerization of ethylene, catalytic activities reached 300 kg∙mol−1∙bar−1∙h−1, and PE of spherical morphology (Figure 2) and relatively high dispersity ĐM = 2.4–2.7 were obtained. Note that Ti2/MgCl2∙2.1EtOH/iBu3Al catalyzed formation of PE with ĐM = 2.0, which confirms the uniformity and stability of the active species. The activity of rac-[1,2-ethylenebis(η5-inden-1-yl)]ZrCl2 (rac-[EBI]ZrCl2, Zr5)/MgCl2∙2.1EtOH/Et3Al was 5 kg∙mol−1∙bar−1∙h−1, and PE with narrow molecular weight distribution (ĐM = 2.1) was formed.
Supports, prepared by the reaction of MgCl2/EtOH with Et3Al (Et3Al/EtOH = 2), were used for the activation of a series of zirconocenes and hafnocenes in ethylene polymerization [81]. The investigation of substituent effects was first carried out using the support of composition MgCl2∙0.30AlEt2.46(OEt)0.54 (Table 1). Catalyst immobilization was obtained by contacting the support overnight with a metallocene solution in toluene; ethylene polymerization was carried out at 50 °C with an ethylene pressure of 5 bar. The ethylene mass flow remained essentially constant during polymerization, indicating negligible decay in catalyst activity.
As can be seen in Table 1, low activities were obtained for (RCp)2ZrCl2 with R = H or Et, but longer-chain alkyl substituents (Zr7, Zr9) provided much higher activities. Introducing branching into the alkyl substituent (Zr8, Zr10) led to poor catalyst performance. According to the study of the authors, the increased steric bulk imposed by branched substituents evidently impedes effective activation of the magnesium chloride support, but the causes of abnormal activities of Zr7 and Zr8 in comparison with Zr6 remain unclear. The analogous hafnocenes were also investigated; Cp2HfCl2 (Hf1) demonstrated an activity of 40 kg∙mol−1∙bar−1∙h−1, whereas with (n-BuCp)2HfCl2 (Hf2), the activity was 240 kg∙mol−1∙bar−1∙h−1. An increase in the number of Me substituents had no effect on the activity (Zr13, Zr15Zr17). Mixed-ligand zirconocenes (RCp)(Cp)ZrCl2 with R = n-Pr, n-Bu, and n-pentyl (Zr18Zr20) demonstrated activities of 740, 1020, and 560 kg∙mol−1∙bar−1∙h−1, respectively, i.e., a smoothed but clear maximum of activity was observed for R = n-Bu in (RCp)(Cp)ZrCl2 with equally unclear reasons. The variation of the composition of MgCl2/EtOH complexes using Zr7 showed a better performance for support, prepared from MgCl2∙1.1EtOH (4300 kg∙mol−1∙bar−1∙h−1), and a higher EtOH content resulted in a lowering of the activity. When ball-milled MgCl2 was used as a support, activity was 720 kg∙mol−1∙bar−1∙h−1. When using iBu3Al instead of Et3Al, more thermally stable catalysts were obtained, and the activity of the Zr7-based catalyst was 12,680 kg∙mol−1∙bar−1∙h−1. It can be assumed that the higher efficiency of iBu3Al in comparison with Et3Al is due to the possibility of side reactions after the formation of unstable ZrEt2 derivatives (Scheme 1c) [37,38]. Note that the use of (n-PrCp)2ZrMe2 (Zr7′) instead of dichloro complex Zr7 with different organoaluminum co-catalysts did not result in changes of activity and PE characteristics, signifying that the limiting factor in the generation of active species in this system is not the alkylation stage. UV-vis spectral studies showed that the uptake of Zr2 from the solution was much slower than that of Zr6 or Zr7, indicating a beneficial effect of an electron-donating alkyl substituent.
The nature of the active species in MgCl2-supported catalysts is still far from being resolved. Reasoning by analogy with TMZNCs, two possible coordinations were proposed for Cp2ZrMe2 (Zr2′, Figure 3) [81]. The formation of alkyl-zirconocenium cations was attributed to the abstraction of the Me group by acidic Mg centers. Another important factor was the complexation of R2AlOEt with Zr2 or Zr2′; supports with relatively low contents of residual ethoxide would therefore be expected to give higher activities for immobilized zirconocenes, as observed in [81]. In our opinion, the coordination of Zr2′ on the MgCl2 surface, presented in Figure 3, is a very rough model: Mg–(μ-OEt)–Al species, formed during the reaction of MgCl2/EtOH with R3Al, have little choice other than to participate in the formation of active species with Zr2 or Zr2′.
It should also be noted that the reaction of MgCl2/EtOH with Ti2 in the absence and in the presence of R3Al is a complex process with a formation of diverse reaction products: as shown in 2016 by Sobota and colleagues [139], during this reaction, Ti2 loses η5-ligands, and heterometallic polynuclear clusters are formed. Given the high nucleophilicity of Mg and Al alkoxides, the decomposition of metallocene fragments is also possible for Zr complexes, but this issue has not yet been studied.
The MgCl2/ROH/R′3Al system was also successfully used for the activation of Group 4 post-metallocenes. Nakayama et al. [77,78,79] prepared an MgCl2/2-ethyl-1-hexanol adduct, which is soluble in n-decane; treatment of this adduct with iBu3Al resulted in the formation of an MgCl2/iBumAl(OR)n support, which is used for activation of bis(phenoxy-imine) titanium complexes Ti7 and Ti9Ti11 (Scheme 6). At 50 °C and an ethylene pressure of 0.9 MPa, catalytic activities were 36.3, 20.8, 36.0, and 26.2 kg∙mmol−1∙h−1, respectively. Mw of PE was 230–1170 kDa, with broadened molecular weight distribution (ÐM = 2.4–3.5). When using MAO as an activator, PE of the worst morphology was obtained (Figure 4).
To confirm the single-site nature of the catalyst using an MgCl2/iBumAl(OR)n activator, copolymerization of ethylene with propylene was performed [78,79]. Complex Ti10 produced a copolymer with a propylene content of 29.3 mol% with an activity of 28.2 kg∙mmol−1∙h−1 and a narrow MWD (ÐM = 1.70), thus confirming homogeneity of the catalytic species [78,79]. The studies of Ti12 and Ti13, containing C6F5 fragment (Scheme 7a), in the polymerization of propylene resulted in the following: Ti12-based system catalyzed living propylene polymerization at 25 °C and atmospheric pressure. The living nature was confirmed by the linear relationship between Mn and polymerization time and by the narrow MWD of the PP obtained (Mn 53 kDa, ĐM = 1.09). The complex Ti13 activated by MgCl2/iBumAl(OR)n produced highly syndiotactic PP ([rr] = 97%, Tm = 155 °C).
When studying bis(phenoxy-imine) Zr complexes Zr4 (Scheme 4b) and Zr21Zr24 (Scheme 7b) using MgCl2/iBumAl(OR)n activator in ethylene polymerization, a catalytic activity of 1820 kg∙mmol−1∙h−1 (Zr22) was achieved. For Zr21 and Zr22, MAO was a less effective activator in comparison with MgCl2/iBumAl(OR)n [79], and PE had a very high molecular weight (up to Mv = 3 MDa) with an exceptionally high bulk density value of 0.47 g∙mL−1 due to excellent spherical morphology of polymer particles.
The main findings and prospects of FI catalysts when using MgCl2/R′nAl(OR)3−n activating supports were presented and discussed by Fujita and colleagues in conceptual work [140] that summarized and complemented the results of previous studies [77,78,79]. In the case of FI catalysts, MgCl2-based supports pose a real competition to MAO, particularly with regards to polymer morphology (Figure 5).

2.4. MgCl2/SiO2 Supports

To prepare MgCl2/SiO2 bisupport particles, Chung and colleagues used agglomeration of silica gel under the action of aq. MgCl2 [141]. After separation and drying at 80 °C, bisupport particles were treated with Me3Al and Zr2 in toluene media. The supported catalyst demonstrated an activity of 1013 kg∙mol−1∙bar −1∙h−1 in the polymerization of ethylene. One might suppose that the organoaluminum activator in this experiment represents MAO, which was formed by hydrolysis of Me3Al; however, the comparative experiment with MAO was characterized by a qualitatively different kinetic profile with rapid deactivation of the catalyst.
A prospective approach to MAO- and borate-free supported catalysts was proposed by Kissin and colleagues who prepared catalysts by mixing Zr5, Bu2Mg, and Et2AlCl solutions (Zr/Mg/Al ratio ~1:10:30), followed by addition of the mixture to Davison-grade 955 silica [142]. In ethylene/hex-1-ene copolymerization at 90 °C and ethylene pressure of 1.3 MPa, catalytic activities up to 7800 kg∙mol−1∙h−1 were achieved, and the hex-1-ene content in the copolymer was 4.4 mol%. The authors assumed that the reaction of Bu2Mg with Et2AlCl results in the formation of MgCl2 that can abstract R or Cl from the alkylated Zr complexes, and that the resulting anionic species have a broadly distributed negative charge to stabilize rac-[EBI]ZrR+ active species.
In 2018, Ko and colleagues described binary support prepared from MgCl2/EtOH adducts and silica [143]. Subsequent treatment with various alkylaluminum compounds and Zr9 resulted in the formation of supported catalysts for ethylene/hex-1-ene copolymerization. Among alkylaluminum compounds, Et3Al2Cl3 demonstrated the highest efficiency, but the hex-1-ene content was ~1 mol% in all experiments. The scheme of catalyst immobilization, presented in [143], looks doubtful as it suggests the formation of a Zr–(μ-O)–Al bond that hinders the formation of active catalytic species.

2.5. Metal Fluorides

Comparative experiments on the activation of Zr1 by Me3Al using MgF2, CaF2, and AlF3, conducted by Soga et al. [75], demonstrated relatively high activities in ethylene polymerization for Group 2 metal fluorides, which were comparable with the activities when using MgCl2 support. Similar results were obtained for Zr3, which was supported on MgF2 in the presence of Me3Al [75].
The absence of further research on metal fluoride-activated single-site Group 4 metal can be attributed to the greater development of MgCl2-based catalysts, widely used in Ziegler–Natta industrial processes. As a result, we have what we have, and the very idea of the activating supports currently neglects the use of metal fluorides as a basis for catalyst-containing particles. At the same time, positive results of the use of fluorine-modified metal oxides have demonstrated the advantages of fluorine-containing supports (see below); therefore, the use of metal fluorides as catalyst’s supports seems undervalued.

3. Silica and Metal Oxides

3.1. Silica SiO2

Activation of Zr1 [73,75], as well as Zr2 and Zr3 [74], by Me3Al using SiO2 support in propylene polymerization failed. According to the study of the reaction of (η5-C5Me5)TiMe3 (Ti14′) with SiO2 and Me3Al, the formation of Ti–Al and Ti–Al2 heterobimetallic species was proposed [144]; however, ethylene uptake by this catalytic system was only mentioned without any experimental data for comparison.
Chemisorption of Me3Al on dehydrated silica material MCM-41 (Mobil Composition of Matter No. 41 [145]) of hierarchical microstructure was studied by Anwander et al. [146]. In the reaction product, a SiMe/AlMe population ratio of 0.45, as well as the formation of Lewis acidic centers, were detected in model reactions using Y[N(SiHMe2)2]3(THF)2; Group 4 metal complexes on this support were not explored. In 2001, a similar study was conducted by Sano and colleagues [147]. After calcination at 500 °C and thermal pre-treatment at 280 °C (the content of Si–OH groups was ~3 mmol∙g−1), MCM-41 was treated with Me3Al (4 mmol per 1 g of the support) and, after separation, calcined at 500–800 °C in vacuo; AlMCM-41 supports were obtained in this way. Propylene polymerization was conducted at 40 °C using Zr1, and iBu3Al was used as an additional activator/scavenger. The use of MCM-41 in place of the AlMCM-41 provided no polymer. In the case of AlMCM-41, catalytic activity passed through the maximum for support and calcined at 700 °C in vacuo.
Detailed investigation of the solid–liquid reaction of silica with Me3Al by 13C, 27Al, and 29Si MAS NMR [148] showed that AlMen moieties constitute about 70% of the total amount of surface-attached Me groups. The amounts of Si–OMe, SiMe3, and SiMe2 moieties were approximately equal to each other and totally amounted to ~30% of Me groups. Note that the formation of Si–OMe fragments implies the presence of Si–Al bonds in the reaction product (not detected by NMR); however, the formation of Si–Al bonds was only mentioned in this work, without direct experimental proof.
The reaction of mesoporous silica SBA-15 with iBu3Al(Et2O) resulted in the formation of well-defined (SiSO)2AliBu(Et2O) and (SiSO)3SiiBu species via Al→Si transfer of iBu group [149]. More in-depth and comprehensive studies of the reaction of SBA-15 with Et3Al [150], iBu3Al [151,152], and Et2AlCl [153] resulted in the detection of a variety of Al1 and Al2 species, which were chemically bonded with silica surface via SiSOAl fragments. Among the products formed after treatment with iBu3Al, (SiSO)2AlH and (SiSO)3SiH species were detected [152]. The formation of Si–Al bonds was not discussed in these later works, which would make the very idea of Si–Al bond formation during the reaction of silica with R3Al questionable.
In conclusion, we note that the interaction of silica with organoaluminum compounds was used as a first stage of the synthesis of functionalized (e.g., fluorinated) activating supports (see Section 5).

3.2. γ-Al2O3

The γ-Al2O3 represents the metastable spinel-type cubic phase of alumina obtained by the thermal dehydration (calcination) of aluminum hydroxides and oxyhydroxides [154]. Its bulk structure (Figure 6a) contains an fcc sublattice of oxide ions that generates octahedral and tetrahedral interstices that accommodate Al ions [155,156]. The precursor of γ-Al2O3 is a well-organized and stable boehmite (AlOOH) that contains a sublattice of cubic close-packaged O2− anions with interstitially situated Al3+ cations [157] (Figure 6b).
Dehydroxylated and partially dehydroxylated γ-alumina surfaces demonstrate an explicit Lewis acidic character due to the presence of coordinatively unsaturated surface Al sites [158,159,160,161]. From a coordination chemistry point of view, the accessibility of both Brønsted acidic sites (BASs) >Al–OH and Lewis acidic sites (LAS) >Al– on γ-Al2O3 surface offers a unique complexation environment (Figure 7). The nature of these sites changes with the level of hydroxylation, which is defined and achieved by pre-treatment temperature and duration [159,160,161].
During the complex experimental and theoretical study of the surface reactivity of γ-Al2O3 [160], Wischert et al. showed high efficiency of Al LAS in N2 absorption and reactivity of Al,O Lewis acid−base pairs in heterolytic dissociation of H2 and CH4 molecules with a formation of Al–H and Al–Me species. The maximum site density was observed for alumina, thermally treated at 700 °C, thus confirming the significance of the process on (110) termination of γ-Al2O3 (see below).
The marked difference in reactivity of γ-Al2O3, calcined at different temperatures, was demonstrated in a number of works, based on the use of the complexes MR4. Partially dehydroxylated at 500 °C, γ-Al2O3 was grafted with Zr(CH2tBu)4, and reaction products contained 2 CH2tBu per Zr atom [162]. Combined experimental and theoretical (Figure 8) studies allowed the authors to propose the reaction mechanism that includes the formation of (AlSO)2ZrR2 species that further react with the alumina surface through the transfer of one of its R ligands onto an adjacent AlS Lewis center, providing a cationic surface complex [(AlSO)2ZrR]+[(AlS)R]. This alkyl transfer was confirmed by a combination of 13C CP MAS NMR and chemical shift calculations. In further theoretical and experimental works [158,163], the formation and chemistry of hydride species [(AlSO)2Zr(H)(μ-H)-AlVI] and [(AlSO)2Zr(H)(μ-R)AlVI] along with cationic [(AlSO)2Zr(H)]+ species stabilized by tetrahedral aluminum hydrides [(AlIV-H)] by hydrogenolysis of [(AlSO)2ZrR]+[(AlS)R] was proposed. During the formation of these hydrides or the hydrogenolysis of alkanes, the alkanes transformed into lower homologues to provide methane and ethane as final products. These processes go beyond the subject of our review but seem relevant to the current problem of chemical recycling and upcycling of polyolefins [164].
Grafting of Hf(CH2tBu)4 on γ-Al2O3 resulted in different products depending on pre-treatment temperature (Scheme 8): the temperature of 350 °C was not high enough to provide the ability of γ-Al2O3 to react with Hf alkyl as an activating support [159]. These results were in line with previous experimental and theoretical studies of the Hf(CH2tBu)4/γ-Al2O3 system [165].
Another important aspect of the chemistry of γ-Al2O3 as an activating support also involved the own reactivity of γ-Al2O3 towards R3Al. As shown in the work of Mazoyer et al. [166], the reaction of calcined γ-Al2O3 with iBu3Al results in the formation of the surface Al–iBu species, and in the presence of molecular hydrogen, Al–H species are formed. It was found that Al–H species can initiate the polymerization of ethylene (with low activities, ~0.55 kg∙mol−1∙h−1). However, it can be assumed that surface Al–R and Al–H species can participate in the formation of Group 4 metal catalytic centers on the surface of γ-Al2O3 when using organoaluminum activators.
The first study of the activation of Group 4 metallocenes by γ-Al2O3 was conducted by Marks and colleagues in 1988 [71]. In their experiments, highly dehydroxylated alumina (DA, 150 m2∙g−1, ~0.1 surface OH∙nm−2) or partially dehydroxylated alumina (PDA, 150 m2∙g−1, ~4 surface OH∙nm−2) were mixed with solutions of Zr2′, (η5-C5Me5)2ZrMe2 (Zr17′), and (η5-C5Me5)ZrMe3 (Zr25′) in n-pentane, with subsequent elimination of the solvent in He flow and treatment by molecular hydrogen. Hydrogenation of propylene was studied at −63 °C, and turnover frequencies Nt (s−1) amounted to 0.3, 0.2, and 1.1 (DA) and 0.1, 0.06, and 0.3 (PDA) for Zr2′, Zr17′, and Zr25′, respectively. Relative activities in ethylene polymerization were for Zr2′ >> Zr17′Zr25′. The percentage of active sites, determined by their coordination with CO, was found to be 4% for Zr2′/DA and 12% for Zr25′/DA.
Activation of Zr1 by Me3Al and Et3Al using γ-Al2O3 support in propylene polymerization showed moderate efficiency, which was comparable with the efficiency of MgCl2 support [73,75]. A similar pattern was observed for Zr3, whereas the catalytic activity of Zr3 when using γ-Al2O3 was four times lower than the activity in the presence of MgCl2 (42 vs. 172 kg∙mol−1∙h−1) [74]. Ti1/ γ-Al2O3 pre-catalyst, prepared by the reaction of a toluene solution of Ti1 with γ-Al2O3 (110 °C, 3 h), demonstrated low activity in propylene polymerization after activation by iBu3Al [126]. Apparently, this was due to the low calcination temperature for γ-Al2O3 used (200 °C). During comparative studies of the activation of Zr9 by different metal oxides, treated by iBu3Al, γ-Al2O3 demonstrated a moderate efficiency relative to other supports [167].
13C CP MAS NMR and EXAFS studies of the reaction of Zr25′ with γ-Al2O3 (calcined at 500 °C) showed the presence of Al–O–ZrMe25-C5Me5) species; some of the surface complexes underwent further interaction with the γ-Al2O3 surface, either involving bridging methyl groups between Zr and Al or full Zr→Al transfer of a methyl group [82]. This sample showed moderate activity in ethylene polymerization (~20 kg∙mol−1∙bar−1∙h−1). The results of the 13C CP MAS NMR study of the reaction of Zr2′ with γ-Al2O3 confirmed the formation of cationic Cp2Zr+ species, which bonded with alumina via Al–O–Zr covalent or Coulombic interactions [82]. Obviously, only the latter Cp2ZrMe+ species can be considered catalytically active, but their content was low (activity in ethylene polymerization was only 3 kg∙mol−1∙bar−1∙h−1). In [82], the authors proposed the common scheme of the structure–activity relationship for Zr25′ and Zr2′, supported by γ-Al2O3 (Figure 9), but the potential catalytic activity of some complexes, presented in this figure, appears doubtful since direct Si–O–Zr or Al–O–Zr bonding will hardly lead to the formation of highly active catalytic species.
A more thorough and accurate theoretical study of the structural and catalytic properties of Zr2′, chemisorbed on dehydroxylated γ-Al2O3, was conducted by Marks and colleagues [155]. They developed a γ-Al2O3 surface model and studied the interactions of Zr2′ with various surface coordination sites. The results of this modeling were compared and contrasted with data for the homogeneous [Cp2ZrMe+][MeB(C6F5)3] system. Since theoretical and experimental data for γ-Al2O3 indicate that the (110) surface predominates (70–80% of the total area), the modeling was focused on the (110) surface (Figure 10a). The optimized alumina (110) surface was found to exhibit significant rearrangement of Al and O ions relative to the bulk; in particular, bulk octahedral Al centers became pseudotetrahedral (AlIV in Figure 10a) while bulk tetrahedral Al centers became pseudotrigonally planar on the surface (AlIII in Figure 10a). The surface O ions were found to have either μ3-O and μ2-O geometries. The μ3-O species were bound to AlIII and AlIV surface ions and to the octahedral AlVI bulk ion, while the μ2-O species were bound to the AlIV and to the octahedral AlVI bulk ions.
The acidic Al sites activate Zr2′ via Zr–Me scission with a Zr→Al transfer of the Me group to the surface; solid-state NMR data showed that the methide group had been placed on the AlIII center at a distance of ~5–8 Å from the metallocenium center. The possible ion pair interactions involving Cp2ZrMe+ and γ-Al2O3 surface were scrutinized to search for the most stable configurations for the complexes at μ3-O and μ2-O sites (Figure 10b). Comparison of the calculated enthalpies of ion pair formation (ΔHform) and separation (ΔHips) for chemisorbed species and [Cp2ZrMe+][MeB(C6F5)3] (Table 2) showed the stronger coordinative capability of the μ2-O sites vs. the μ3-O sites. Moreover, interactions found in oxo-bridged configurations were significantly stronger than those in the dioxo-bridged configurations. For [Cp2ZrMe+][MeB(C6F5)3] the adduct formation was less exothermic, but the value of ΔHips was comparable to that in the heterogeneous μ3-O dioxo-bridged complex.
π-coordination and insertion of ethylene molecule required only partial ion pair separation. These reaction steps were modeled for all four coordination modes of chemisorbed Cp2ZrMe+ and for [Cp2ZrMe+][MeB(C6F5)3] (Figure 11).
As can be seen in Figure 11, π-coordination and insertion of ethylene molecule were easier for μ3-O coordinated species even in comparison with [Cp2ZrMe+][MeB(C6F5)3]. Therefore, only a fraction of the surface sites in dehydroxylated γ-Al2O3-supported metallocenes are catalytically significant. However, those catalytically significant surface sites exhibit greater catalytic activities than their homogeneous analogs. In addition, it can be assumed that the chemistry of modified alumina supports (see Section 5) may involve surface modification with a decrease in the content of μ2-O sites.
Ti(III) complex Ti15 with the formula [Ti(nacnac)(CH2tBu)2] (nacnac = [Ar]NC(Me)CHC(Me)N[Ar], Ar = 2,6-(CHMe2)2C6H3) [168] was grafted onto alumina that was partially dehydroxylated at 700 °C. This reaction is accompanied by the release of 0.54 equivalents of neopentane per initial surface OH group, which confirms Al–O–Ti bonding with the support. The obtained complex was found to be active in ethylene polymerization, which expands the understanding of the mechanism of α-olefin polymerization by the notion that d1 Ti alkyls are possible active sites in the heterogeneous Ziegler–Natta polymerization catalysts.

3.3. Silica-Alumina SiO2/Al2O3

Due to their combined Lewis and Brønsted acidities, amorphous SiO2/Al2O3 are widespread supports for multifunctional heterogeneous catalysts. The local environment of the acid sites is essential for activation, and in 2010, Chizallet and Raybaud [169] proposed bridging Si–(OH)Al fragments as an active BAS (Scheme 9a).
During the 13C CP MAS NMR and EXAFS studies of the reaction of Zr25′ with SiO2/Al2O3 [82], different products were detected (Scheme 9b). The main product was the complex with a Si–O–Zr bond, which was observed when studying the reaction of Zr25′ with silica (see Section 3.1), but the minor products can be considered as trinuclear species with more or less degrees of charge separation. However, experiments on ethylene polymerization showed zero catalytic activity.
In the activation of Zr9, iBu3Al-treated SiO2–Al2O3 demonstrated moderate efficiency, which is comparable with the efficiency of Al2O3 if specific surface areas of the supports are considered [167].

3.4. Clay Minerals and Related Systems

In 2002, Weiss and colleagues published the results of the study of activation of Group 4 metallocenes by kaolin (vacuum-dried) and montmorillonite (pre-heated at 200 °C) as inorganic carriers with the use of Me3Al and iBu3Al as activators [170]. In view of the results of preliminary investigations [171], the authors postulated the mechanism of activation, presented in Scheme 10. Note that this mechanism is quite correlated with the results of the activation of metallocenes by silica/R3Al (no activity) [73,74,75].
Using kaolin/Me3Al as an activating support, in the polymerization of ethylene (50 °C, 10 bar), activities of 5.2, 4.3, and 1.5 kg∙mmol−1∙h−1 were achieved for Zr2, Cp2ZrHCl (Zr2″), and Ti2, respectively. On montmorillonite/R3Al support Zr2 demonstrated activities up to 14 kg∙mmol−1∙h−1, and iBu3Al was found to be a more efficient co-catalyst in comparison with Me3Al. In propylene polymerization, Zr1/montmorillonite/iBu3Al ([Al]/[Zr] = 3200) and Zr1/montmorillonite/Me3Al ([Al]/[Zr] = 800) catalysts were even more active than the homogeneous Zr1/MAO system ([Al]/[Zr] = 1000); the activities amounted to 60.3, 6.8, and 5.7 kg∙mmol−1∙h−1, respectively. However, the isotacticity of PP, obtained on supported catalysts, was lower (80 and 83% vs. 89%). Another noteworthy result was obtained when studying the activation of homo- and bimetallic Ti complexes Ti16 and Ti17: at [Al]/[Zr] = 2000, the montmorillonite/iBu3Al activator was much more efficient than MAO (Figure 12).
Montmorillonite/Et3Al system was also studied by Nakano and colleagues in activation of rac-Me2Si(2-Me-4-Ph-1-Ind)2ZrCl2 (Zr26) [172]. After pre-heating at 200 °C and treatment with Et3Al (support 1), 2,6-dimethylpyridine was added in different amounts (supports 2 and 3). Supports 1–3 were reacted with iBu3Al and Zr26. UV-vis spectra of the catalysts showed the formation of cationic Zr species only for support 1, containing highly acidic centers.
Since a specific surface area of the support is of great importance for catalytic activity, Sun and Garcés suggest acid lamellar aerogel prepared by exfoliation of clay in acidic media [173]. Supported catalysts prepared by the reaction of aerogel with R3Al and Zr2′ demonstrated activities of 29–69 kg∙mmol−1∙h−1. Since iBu3Al was the best activator, and Me3Al was the worst, the mechanism of activation can not be interpreted through the formation of alumoxanes. The use of spray-dried silica/clay agglomerates for activation of Zr9 and Ti8 has been patented by W. R. Grace & Co. [174]. When using 2–6 mmol of iBu3Al or Me3Al and 12–120 μmol of pre-catalyst per 1 g of support, activities of 0.3–2 kgPE∙gCat−1∙h−1 were achieved in ethylene polymerization.
Acid-treated commercial montmorillonite clays (K10, K30) and alumina-modified SBA-15 were studied by Cuenca and colleagues in 2010 [175]. After treatment with Me3Al or Et3Al, these supports (2.5 g) were mixed with (η5-C5Me5)(η5-C5H4SiMe2CH2CH=CH2)ZrCl2 (Zr27, 0.007 μmol) and iBu3Al (1 mL) in toluene and studied in polymerization of ethylene (1 bar). Alumina-modified SBA-15 was inactive, whereas all montmorillonite/R3Al samples were proven to be activating supports with the best performance of 19.1 kgPE∙mmolZr−1∙h−1 for K10/Me3Al. The authors proposed that montmorillonite/R3Al supports react with dichloro complex Zr27 in two pathways (Scheme 11). The first one was the ligand exchange and subsequent alkyl/halide abstraction; the support reacts as an organo-Lewis-acidic cocatalyst similar to MAO (Scheme 11a). An alternative activation process involves protonolysis of the M–R bond on Brønsted acidic centers (Scheme 11b). The Lewis and Brønsted acidities of the supports were determined using pyridine as a probe molecule by monitoring the bands for pyridine absorbed on BAS and LAS at 1540 and 1450 cm−1, respectively. It was found LAS comprised the main fraction of the acid sites, detected in montmorillonite and montmorillonite/R3Al. Treatment with R3Al resulted in a decrease in LAS content, and it was notable that BASs were still detected, though in small numbers, after R3Al treatment.
As a final note, the use of clay minerals did not receive further development due to the creation of more efficient activating supports (See Section 4 and Section 5).

3.5. Other Metal Oxides

A comparative study of the activation of Zr9 by different metal oxides (Al2O3, TiO2, AlPO4, TiO2–Al2O3, SiO2–Al2O3, TiO2–SiO2, and H–Y zeolite), calcined at 500 °C and treated with iBu3Al, in the polymerization of ethylene was conducted by Ikenaga et al. [167]. These oxides were also tested in the cracking of cumene to evaluate their acidity. Among the catalysts under study, the AlPO4-supported catalyst showed the highest activity (230 kg∙mol−1∙h−1). The activity of the zeolite-supported catalyst was two times lower. Catalysts supported on other oxides demonstrated even lower activities. No correlation between the activity and the surface area of the oxides was observed. In cumene cracking (in view of the correlation between the cracking rate and Brønsted acidity), the highest average conversion was obtained using H–Y zeolite. Although SiO2–Al2O3 showed the second-highest cracking activity, this oxide was ineffective as an activator for metallocene. Moreover, the most effective activator of polymerization, AlPO4, showed negligible activity in the cracking of cumene (Figure 13).
FT-IR spectral studies of pyridine absorption on Al2O3 and AlPO4 showed more Lewis acidity of the former phase, and the absorption band of pyridinium ions, derived from BAS, was not observed in both cases. Under optimized conditions, the activity of Zr9 reached 820 kg∙mol−1∙h−1 (Mn = 10.0 kDa, ÐM = 2.5). At the same time, activities of Zr2 and Zr1 were 18 and 66 kg∙mol−1∙h−1, respectively, and in the latter case PE with ÐM = 5.3 was obtained. The formation of active sites on the AlPO4 was attributed to the reaction of surface –OH groups with iBu3Al [167]. Nb2O5 was found to be ineffective in the activation of Zr25′; only Nb–O–ZrMe25-C5Me5) species have been detected [82].
Modification of silica by MgO, LiOH, or NaOH resulted in the formation of new phases capable of activation of rac-[SiMe25-4,5,6,7-tetrahydroinden-1-yl)]ZrCl2 (Zr28)/iBu3Al in the polymerization of ethylene and propylene [176]. The highest catalytic activities were detected when using SiO2/MgO support, they comprised 70 kgPE∙gZr−1∙h−1 and 340 kgPP∙gZr−1∙h−1. The acidity of the surfaces was estimated using a set of indicators, and acid strength in the range of 5.6 < pKa < 3.0 was detected for SiO2/MgO. The acid strengths of SiO2 and MgO and Mg(OH)2 were found to be weak [177]; therefore, in SiO2/MgO phase, SiO2 and MgO have undergone a reorganization of the bonding structure with a formation of new acidic sites that are required for high activity of zirconocene in propylene polymerization. The role of the basic anionic sites had been also raised in [176], without mechanistic interpretation.

4. Sulfated Metal Oxides and Related Systems

More than 40 years ago, it was shown that the acidic sulfate treatment of metal oxides such as ZrO2, TiO2, SnO2, and Fe2O3 increases the surface acidity and catalytic activity in the processes involving carbenium ions [178]. The Hammett acidity function (H0) for superacid sites of sulfated metal oxides (SMO) was up to <−16. The common methods of the preparation of SMO are the treatment of precipitated hydroxides with H2SO4 with subsequent calcination, thermal decomposition of metal sulfate hydrates, or treatment of crystalline oxides with H2SO4. The features of the methods, as well as specific reaction and thermal treatment conditions, depend on the type of metal oxide and result in supports with different degrees of sulfation, –OH group content, and porosity.
Most of the studies of Group 4 metal complexes, supported on sulfated metal oxides, have been conducted by Marks and colleagues, and their significant proportion relates specifically to the catalytic hydrogenation of arenes [179,180,181]. Along with that, attention has been paid to the nature of the catalytic species and coordination polymerization of α-olefins, which matters for our review as well.

4.1. Sulfated ZrO2

The preparation of active sulfated ZrO2 (ZrS) is based on the calcination of sulfate-doped zirconium hydroxide or zirconium sulfate hydrates [182,183]. Active SZ is based on tetragonal zirconia (t-ZrO2) as the main phase. ZrS, which is based on the more stable monoclinic ZrO2, is much less active than tetragonal ZrS [184]. High-resolution transmission electron microscopy (HRTEM) studies of ZrS showed that the presence of SO42− groups stabilizes small tetragonal ZrO2 crystallites and induces the formation of well-faceted particles with the (110) plane as the most abundant crystallographic plane [185]. Periodical DFT calculations and FT-IR spectral studies showed a broad variety of structures for sulfated surfaces on t-ZrO2 [184].
In 1998, Ahn and Marks reported the preparation of ZrS by thermal decomposition of Zr(SO4)2∙4H2O with subsequent thermal activation under high vacuum (5 × 10−6 Torr) at 300, 400, and 740 °C [83]. ZrS300 contained strong BAS/LAS and weak BAS, ZrS400 contained both strong BAS and LAS, and ZrS740 contained mostly LAS. Next, Zr2′ and Zr25′ were adsorbed by acid supports from n-pentane solutions. The complex Zr2′ demonstrated high activity in hex-1-ene hydrogenation, and catalytic activity correlated with strong BAS populations. Less sterically hindered Zr25′ showed a dramatic enhancement in hydrogenation activity when supported on ZRS400:, for example, it mediated fast hydrogenation of benzene at 25 °C and 1 bar H2. Zr2′/ZrS400 and Zr25′/ZrS400 also catalyzed ethylene homopolymerization at 25 °C with activities of 1.5 and 40 kg∙mol−1∙bar−1∙h−1, respectively. 13C CP MAS NMR spectra of the Cp2Zr(13CH3)2/ZrS400 sample showed the presence of the signals of cationic Zr–Me+ species, without the signals of the transferred methide group. The authors have proposed that the sulfated zirconia activates Group 4 metallocenes via Zr–C protonolysis with a formation of ion pairs (Scheme 12). This reaction mechanism is qualitatively different from the Zr→Al migratory mechanism that is characteristic for the γ-Al2O3 activating support.
More thorough investigation of the efficiency of ZrS400 supports in activation of different Group 4 metal complexes [84] showed that catalytic activities are reduced in a raw Zr(CH2Ph)4 > Zr(CH2SiMe3)4 > Zr(CH2tBu)4 > Zr25′ for both ethylene and liquid propylene polymerizations. 13C CP MAS NMR spectral studies of Zr2′/ZrS400 and Zr25′/ZrS400 showed the presence of both μ-oxo and cationic species, the signals of transferred methide groups were not reliably detected. These results confirm weaker Lewis acidity of sulfated zirconia than that of dehydroxylated γ-Al2O3.
The further studies of Group 4 polymerization catalysts by Marks’ group [186] were focused on well-known post-metallocene Hf complex Hf3 (Figure 14a), exhibiting excellent catalytic activity in ethylene homopolymerization and ethylene/α-olefin copolymerization [187,188,189,190,191,192]. In the 13C CP MAS NMR spectrum of Hf3/ZrS the downfield-shifted broad signal at δ = 65.5 ppm (Figure 14b, D) can be assigned to a “cation-like” electron-deficient LHf13CH3+ species, as evidenced by the marked downfield shift from δ = 60.2 ppm for Hf3 pre-catalyst (Figure 14b, A). A similar downfield-shifted signal was observed at δ 65.8 ppm in the homogeneous [LHf13CH3]+[MeB(C6F5)3] [193]. DFT modeling of chemisorption of Hf3 on the ZrS surface yielded an average distance d(Hf–O) of 2.14 Å that is slightly elongated in comparison with typical Hf–OR bonds (1.92–2.03 Å).
Such close contact could not fail to affect the catalytic activity. In ethylene homopolymerization, activities of Hf3/ZrS and homogeneous Hf3/[Ph3C]+[B(C6F5)4] were 4.5 and 6260 kg∙mol−1∙bar−1∙h−1, and activities in ethylene/oct-1-ene copolymerization amounted to 3.9 and 25,600 kg∙mol−1∙bar−1∙h−1. And, it is highly significant that oct-1-ene incorporation in the transition from homogeneous to supported catalyst has fallen from 50 to 2.6 mol% [186]. Another important difference between homogeneous and supported catalysts consisted in opposite directions of first ethylene insertion, at Hf–Aryl and Hf–Me bonds, respectively.
In the search for more efficient catalytic systems, pyridylamido complexes Hf4 and Hf5 (Scheme 13) were synthesized and studied [194]. Being supported on ZrS, Hf4 and Hf5 have demonstrated higher activities in comparison with Hf3: 15.7 and 22.7 vs. 4.5 kg∙mol−1∙bar−1∙h−1 in ethylene homopolymerization; 11.5 and 16.6 vs. 3.9 kg∙mol−1∙bar−1∙h−1 on ethylene/oct-1-ene copolymerization.

4.2. Sulfated Al2O3

Sulfated alumina (AlS), prepared by the impregnation of γ-Al2O3 with 1.6 M H2SO4 and subsequent 3 h calcination at 550 °C, was studied as an activating support for (C5Me5)2ZrMe2 (Zr17′) by Marks and colleagues in 2003 [85]. 13C CP MAS NMR spectroscopy revealed that chemisorption proceeds via Zr–Me bond protonolysis with a formation of cationic (C5Me5)2ZrMe+ species. In ethylene polymerization experiments (toluene, 60 °C, 10 bar), catalytic activities of Zr complexes were 2100 (Zr(CH2Ph)4), 1200 (Zr(CH2SiMe3)4), 1100 (Zr25′), 160 (Zr17′), and 80 (Zr2′) kg∙mol−1∙h−1 [85]. Under similar conditions, activities of these complexes, supported on γ-Al2O3, were an order of magnitude lower.
In 2010, Marks’ research group revisited the study of the activation of half-sandwich and sandwich Zr(IV) complexes by AlS. In particular, they prepared nanosized sulfated alumina (n-AlS) by calcination of commercial 25 nm γ-Al2O3 to remove residual surfactant, surface modification with H2SO4, and final calcination; TEM analysis indicated that the nanoparticles were loosely aggregated nanorods (diameter∼10 nm, lengths∼10–30 nm) [195]. When studying the chemisorption of Zr17′ on n-AlS by 13C CP MAS NMR spectroscopy, both (C5Me5)2ZrMe+ and Al–Me species were detected in contrast with AlS. Experiments on ethylene (co)polymerization on Zr(CH2Ph)4/n-AlS and Zr25′/n-AlS showed higher catalytic activities in n-heptane relative to activities in toluene, which is generally not surprising.
Increased understanding of chemisorption of Zr(IV) complexes Zr2′ and Zr25′ on AlS was the result of further research with the use of Zr X-ray absorption spectroscopy (XAS) and DFT modeling [179]. DFT modeling reveals two types of sulfate species, sites SA (formed by double-exchange/condensation with the surface and does not afford an acidic proton) and SB (formed by single exchange/condensation reaction with the surface, preserving one acidic proton, which is then transferred to an Al-O surface site) as well as (Al)nOH hydroxyl groups, with the OH coordination to three or two Al ions. Zr–Me protonolysis is favored at the (Al)3OH site, with a formation of Cp2ZrMe+ or (C5Me5)ZrMe2+ species coordinated at sulfate groups. DFT modeling at site SB and experimental studies with the use of Zr K-edge extended X-ray absorption fine structure (EXAFS) and X-ray absorption near-edge structure (XANES) methods yielded a d(Zr–O) of 2.37 Å, which was significantly longer than in typical covalent Zr–O bonds. Very similar results were obtained when studying (C5Me5)Zr(CH2Ph)2+ species [180] (Figure 15).
A very important and interesting aspect of the use of AlS is associated with the emergence of polymerization stereoselectivity for chemisorbed, initially achiral pre-catalysts such as Zr25′ [196]. At 25 °C, Zr25′/AlS produced polypropylene and poly(hex-1-ene) with 89% and >95% isotacticity, respectively. The requisite Zr25′/AlS C1 symmetry was attributed to a combination of AlS surface topography and sterically imposing groups beyond the active site. DFT confirmed that propylene coordination and insertion are favored at the re enantioface (Figure 16), which results in isotactic polymerization via a ‘back-skip’ mechanism.
Chemisorption on AlS and catalytic activity of post-metallocene constrained-geometry Group 4 metal complexes Zr29′Zr31′ and Ti18 (Scheme 14) in (co)polymerization of ethylene were investigated by Williams and Marks in 2009 [197]. 13C CP MAS NMR spectral studies showed that chemisorption of Zr complexes proceeds with a formation of cationic LZrMe+ species in two reaction pathways, via Zr–Me bond protonolysis and Zr→Al methide transfer: in contrast with activation of Zr17′, surface Al–Me fragments were detected during activation of constrained-geometry complexes.
Experiments on polymerization of ethylene (10 mL of toluene, 5 bar) were conducted with the use of supported catalysts and under homogeneous conditions with the use of perfluoroborate activator [Ph3C]+[B(C6F5)4] (Table 3). In all cases, supported catalysts were significantly less active than the homogeneous catalysts, presumably due to steric hindrance from the ancillary ligand-support interaction blocking the ethylene approach to the active site. However, the use of supported catalysts resulted in the formation of high-MW PE, whereas borate-activated Zr30′ and Zr31′ produced polyethylene waxes. Ti complex Ti15 was more active than Zr catalysts. For ethylene/1-hexene copolymerizations at 25 °C, binuclear complex Ti18/AlS was found to be the most active (16.4 kg∙mol−1∙h−1), followed by Zr31′/AlS, Zr30′/AlS, and Zr29′/AlS (1.24, 1.21, and 0.54 kg∙mol−1∙h−1, respectively). In this way, the cooperative ‘binuclear effect’, observed previously by Marks and colleagues for bimetallic Group 4 constrained-geometry complexes in homogeneous conditions [198,199,200], does not occur when using supported catalysts.
During activation of a highly efficient industrialized pyridylamido-naphthyl Hf pre-catalyst Hf3 (see Figure 17) by AlS, both protonolysis and methide Hf→Al transfer were detected [194]. Hf EXAFS spectrum of Hf3/AlS allowed to determine the ligand environment of Hf as two Hf−N (2.30 and 2.08 Å), one Hf−C (2.26 Å), one Hf−Me (2.21 Å), and one Hf−Osupport (2.06 Å). Comparative studies of supported Hf3 and pyridylamido complexes Hf4 and Hf5 showed that when using AlS these complexes demonstrate close catalytic activities (35.1–46.8 kg∙mol−1∙bar−1∙h−1 in ethylene homopolymerization, 16.6–19.4 kg∙mol−1∙bar−1∙h−1 in ethylene/oct-1-ene copolymerization), which was not observed when using ZrS as an activating support [194]. Oct-1-ene incorporation was also dependent on the pre-catalyst and support used (Figure 17).
A rather high efficiency of AlS as activators for Group 4 metallocenes could not fail to interest chemists and chemical engineers employed in the polyolefin industry. AlS was the subject of a number of patents, filed by Chevron Phillips. In particular, in ethylene polymerization with the use of zirconocene Zr9, AlS, and Et3Al activators, catalytic activities up to 3.23 kg∙gAlS−1∙h−1 have been reached [92]. However, it should be borne in mind that catalytic efficiency of metallocene/AlS/R3Al systems is critically dependent on many factors such as the characteristics of Al2O3, method of sulfation, [SO42−]/malumina ratio, conditions of calcination, nature of Group 4 metal complex and R3Al, and others. Therefore, for example, the highest activity of 3.23 kg∙gAlS−1∙h−1 was achieved when AlS was prepared by impregnation of (NH4)2SO4 solution onto Ketjen grade B alumina (1.5 mmol of sulfate per 1 g of alumina), followed by vacuum-drying and calcination in air at 550 °C [92]. The role of the pre-catalyst can be illustrated by the data from the patent [201] (Scheme 15, Table 4), the maximum productivity of 17.3 kg∙gCat−1∙h−1 was achieved when using Zr36/iBu3Al in homopolymerization of ethylene in the absence of H2. In this patent, only calcining conditions are given (heating to 550 °C with a 240 °C∙h−1 ramp rate in air and 3 h hold at 550 °C).
For the synthesis of PE composites, the chemists of Chevron Phillips proposed the use of binary mixtures of metallocene catalysts, the first being the catalyst of the formation of low-MW PE (Mw 5–50 kDa, for example, (η5-Indenyl)2ZrCl2 (Zr37), and the second being the formation of high-MW PE (Mw 200–1200 kDa, for example, Zr33 [202]. AlS support was prepared by the reaction of bohemite (‘Alumina A’ from W. R. Grace & Co., Columbia, MD, USA) with aq. (NH4)2SO4, followed by vacuum-drying and calcination at 400–600 °C. The combination of binary pre-catalyst, AlS, and iBu3Al allowed to obtain ethylene/hex-1-ene copolymers with unique mechanical and rheological characteristics that may be employed for the production of large diameter pipes. Polymerization experiments were conducted with the use of molecular hydrogen (350 ppm with respect to the ethylene flow) at 92–95 °C. Maximum catalytic activities amounted to 2.84 and 3.16 kgPE∙gCat−1∙h−1 for ethylene homo- and copolymerizations, respectively.
The importance of pre-treatment of AlS by R3Al was the subject of the patent [203]: when using metallocene dichlorides, such pre-treatment resulted in significant increase of the catalytic activity. From the mechanistic viewpoint, pre-treatment of AlS by organoaluminum compounds results in the substitution of all BAS by –OAlR2 fragments, and this modified support retains the ability to absorb LZrCl2 via ligand exchange and Zr→Al alkyl transfer to LAS. However, from the analysis of scientific periodicals, this issue remains unexplored. An efficient, but incomprehensible, method to increase the efficiency of metallocene/AlS/R3Al systems, proposed in the patent [204], is based on using simultaneous mixing of the catalyst components that results in ~30% improvement of catalytic activity. This effect still remains unexplored with the use of modern analytical methods.

4.3. Other Sulfated Metal Oxides

Sulfated tin (IV) oxide microparticles were prepared by sedimentation of Sn(OH)4 gel at pH 8, elimination of residual Cl by aq. NH4OAc, and drying and reaction with 3M H2SO4, followed by calcination at 500 °C for 3 h [86]. 13C CP MAS spectral study of chemisorbed Zr17′ showed the formation of cationic species, SnS–O–Zr(C5Me5)2Me were detected in trace amounts. In ethylene polymerization experiments (toluene, 60 °C, 10 bar), catalytic activities of Zr complexes were 660 (Zr(CH2Ph)4), 98 (Zr25′), 15 (Zr17′) and 8.6 (Zr2′) kg∙mol−1∙h−1.
In 2004, Marks and colleagues undertook a comparative study of sulfated metal oxides ZrS, AlS, SnS, FeS, and TiS as activators for Zr2′ [87]. Solid-state 13C NMR spectroscopic analysis of the Cp2Zr(13CH3)2 adsorbate complexes reveals that each support affords a different ratio of μ-oxo and cationic species. The chemical shift of the signals of Me groups in cationic species (δ = 37.1, 35.7, 37.6, and 33.8 ppm for ZrS, SnS, FeS, and TiS, respectively) correlated with the efficiency of activating supports. In ethylene polymerization experiments (60 °C, 10 bar, using 30 mg of the catalyst), Zr25′ and Zr(CH2Ph)4 have demonstrated wide spectra of catalytic activities depending on the type of support (Table 5); sulfated zirconia was the most efficient activating support.

5. Fluorinated Metal Oxides and Related Systems

As was shown in Section 3, some metal oxides have demonstrated the ability to activate Group 4 metallocenes and post-metallocenes, which is primarily due to the high Lewis acidity of the metal centers on the support surface. The obvious way to increase the Lewis acidity consists in the replacement of the oxygen atom in the ligand environment of the acidic metal center by a more electronegative element that is fluorine, without options. On the other hand, partial chlorination can also be effective.
However, in any case, the starting oxide should already be acidic. And this is the reason why fluorinated silica was not used in catalysis [205] and, in particular, has demonstrated zero efficiency as an activating support for zirconocenes [89,90]. On the contrary, in the case of alumina and especially silica/alumina, surface modifications by introducing halogen (F, Cl) resulted in supports with high acidity and bright prospects for use in the polyolefin industry.

5.1. Fluorinated Al2O3

During the studies of gas-phase disproportionation of fluorochloroalkanes with the use of γ-Al2O3 catalyst, fluorinated alumina Al2O3(F) was formed [206]. γ-Al2O3 alone was inactive and required pre-treatment with HF (or fluorocarbon); in this way, it is Al2O3(F) that is a true catalyst. Although chemical analyses reveal significant amounts of F in the solid after this activation procedure, nearly no structural changes of the γ-Al2O3 were detected by X-ray diffraction (XRD) [206]. Only after a few days, crystalline α-AlF3 becomes detectable. Other approaches to Al2O3(F) are based on ‘wet’ methods, the reaction of γ-Al2O3 with NH4F or with NH4HF2 in aqueous media [207], and the content of F in the reaction product is easily determined by γ-Al2O3/[F] ratio. The reaction of Al2O3 with HF (Equation (1)) is exergonic (ΔG = −329 kJ∙mol−1), and the formation of AlF3 during the ‘fluorination’ of alumina should not be ruled out.
Al2O3 + 6 HF → 2 AlF3 + 3 H2O
In this way, both prolonged gas-phase fluorination or an excess of fluoride when being used in a ‘wet’ process may result in a high fluorine content with a formation of AlF3 after calcination. The problem is that the most stable crystalline phase of AlF3, namely α-AlF3, does not exhibit the properties of strong Lewis acid since the surfaces derived from α-AlF3 are covered completely with basic fluoride anions, irrespective of the plane indexes [208]. It is relevant to remind here that crystalline AlF3 was found to be a low-efficient activating support for Zr1 in the presence of Me3Al in the early work of Soga and Kaminaka [75]. It is quite possible that these unsuccessful experiments were conducted with large-sized crystallites of α-AlF3, whereas AlF3 showed high Lewis acidity only as β-AlF3 crystal modification or in the amorphous form [208,209]. Efficient methods of the synthesis of high-surface amorphous AlF3 have been developed in the early 2000s [210,211]. These methods consist of a fluorolytic sol–gel reaction starting from aluminum isopropoxide, followed by an activation step under the flow of a fluorinating gas. A similar method was also used for the preparation of Al2O3(F) with relatively low [F]/[Al] ratios (0.2–0.5) by the reaction of Al(OCH(Me)Et)3 with HF in IPrOH with subsequent hydrolysis and calcination [212]. Sol–gel methods are too complex, and fluorination of γ-Al2O3 seems preferable for the production of highly acidic Al-based supports.
However, the chemical nature of the surface of Al2O3(F) is not entirely clear. Back in the 2000s, Fischer et al. performed a solid-state NMR study of fluorinated γ-Al2O3 that showed the presence of the species with VIAl(O6−nFn) (n = 1, 2 or 3) environments [213]. With an excess of F, the remaining weakened Al–O–Al bonds anchoring VIAl(O3F3) to the bulk of alumina were eventually broken with the formation of AlF3∙3H2O.
The nature of LAS and BAS formed on the alumina surface during fluorination and the efficiency of fluorinated Al2O3 supports in the activation of zirconocenes Zr2 and Zr2′ were studied by Panchenko et al. [214]. Four samples of Al2O3(F) with different F content, were prepared by treatment of γ-Al2O3 (specific surface area 196 m2∙g−1, pore volume 0.58 cm3∙g−1) with aqueous solutions of NH4F and subsequent calcination in air at 500 °C for 2 h. Then, the supports were treated with a toluene solution of Zr2 or hexane solution of Zr2′ using a ratio of 0.05 g Zr per gram of support. Ethylene polymerization was conducted at 80 °C and an ethylene pressure of 10 bar; in the case of Zr2/γ-Al2O3 and Zr2/Al2O3(F), iBu3Al was added.
Determination of LAS and BAS on the surfaces of γ-Al2O3 and Al2O3(F) was conducted using additionally dehydroxylated samples (500 °C, 0.02 Torr, 1 h) by FT-IR studies of the adsorption of CO (LAS) and pyridine (BAS). The concentrations of LAS and BAS (CS) were estimated from the integrated intensity of the CO absorption bands in the region 2175–2240 cm−1 and pyridinium ion absorption bands with a maximum of 1535–1550 cm−1 using Equation (2).
C S = A A 0 p 10 3
where CS—concentration of acidic sites (μmol∙g−1), A—integrated absorption of the examined band (cm−1), A0—integral adsorption coefficient for CO or pyridine adsorbate concentration 1 μmol∙g−1 (cm∙μmol−1), and p—amount of the sample per 1 cm beam, g∙cm−2. The values of A0 for CO absorption bands at 2215, 2210, 2202, 2198, and 2180 cm−1 are 1.15, 1.10, 1.02, 0.95, and 0.8 cm∙μmol−1, respectively; for pyridinium, the ion absorption band A0 is equal to 3 cm∙μmol−1.
The strength of BAs was characterized by the proton affinity (PA), which was calculated from the IR spectra of adsorbed pyridine by Equation (3). Characteristics of the supports are given in Table 6.
P A = lg ( 3400 ν c g ) 0.0023 51
where 3400 cm−1 is the frequency of N–H stretching vibration for free pyridinium ion, and νcg is the position of the gravity center of N–H stretching vibration for pyridinium ion on the support surface.
Aluminas, calcined at temperatures above 500 °C, contain four types of LAS. In the IR spectra of adsorbed CO, these LAS are characterized by absorption bands at 2178–2182 (L-1), 2185–2195 (L-2), 2203–2210 (L-3), and 2225–2240 cm−1 (L-4). The L-1, L-2, and L-3 LAS correspond most likely to coordinatively unsaturated aluminum atoms in an octahedral environment, while L-4 corresponds to those in a tetrahedral or pentahedral environment. Weak LAS of the L-1 and L-2 types were found on the surface of Al2O3(F) samples. In addition, the spectra of CO, adsorbed on Al2O3(F), contain absorption bands at 2215–2208 cm−1 corresponding to new, stronger sites (L-3) (Table 6). LAS of maximum strength (absorption band 2215 cm−1) were observed in the Al2O3(5%F) sample. As shown by Corma et al. [215], the addition of HF to Al2O3 can affect the strength of LAS via direct introduction of one F (L-2) or two F (L-3) into the coordination sphere of aluminum. As can be seen in Table 6, the introduction of fluorine into Al2O3 sharply decreases the total amount of LAS but increases the content of strong type L-3 LAS. In Al2O3(F) samples with a fluorine content of 3–7 wt%, strong BAS with PA 1190 kJ/mol were detected. The concentration of these sites was 2–7.3 μmol∙g−1 with a maximum in the Al2O3(5%F) sample.
The content and strength of basic sites at the support surfaces were estimated by analyzing FT-IR spectra of adsorbed CDCl3, and it was shown that fluorination of Al2O3 results in formation of weak basic sites.
The catalysts prepared with the use of dimethyl zirconium complex Zr2′ were active in the ethylene polymerization without additional cocatalyst iBu3Al, indicating the formation of the active sites containing a Zr–Me bond by the direct interaction of zirconocene with LAS and BAS of the support. In the polymerization run with iBu3Al, the activity was only slightly higher (620 kgPE∙molZr−1∙h−1) than that without iBu3Al (538 kgPE∙molZr−1∙h−1). The catalyst Zr2′/Al2O3 showed low activity; the activities of Zr2′/Al2O3(F) were noticeably higher with a maximum observed for Zr2′/Al2O3(5%F). In the presence of iBu3Al, the activity of Zr2/Al2O3(F) catalysts was higher than that of Zr2′/Al2O3(F). This result was attributed to the formation of inactive μ-oxo-like zirconium compounds during the reaction of Zr2′ with BAS of the support. The activity of Zr2/Al2O3(F) catalysts also depended on fluorine content in the support; Zr2′/Al2O3(5%F) showed a maximum activity (Table 6) [214]. In this way, Al2O3(F) was a more effective activating support in comparison with γ-Al2O3.
The above-discussed works are of more theoretical than practical value. Al2O3(F) was used for the activation of Cr-based catalysts of ethylene polymerization in 1992 [216]. In the early 2000s, the chemists of Phillips Petroleum (subsequently Chevron Phillips) paid special attention to fluorinated alumina and related phases as activating supports for Group 4 metallocenes. In particular, Al2O3(F) was prepared by the ‘wet’ method from Ketjen B alumina (Akzo Nobel, surface area 340 m2∙g−1, pore volume 1.78 mL∙g−1) with the use of NH4HF2 as an F source at different [F]/[Al] ratios [89]. After calcination at ~500 °C in dry air, these supports were tested in combination with Zr9 and Et3Al in ethylene homopolymerization. Activities of 240–950 gPE∙gCat−1∙h−1 were achieved (while the activity was 6.9 gPE∙gCat−1∙h−1 when using γ-Al2O3 in a comparative experiment), and the relationship between efficiency of the support and fluorine content had a distinct maximum.
Preparation of Al2O3(F) from Ketjen B alumina by ‘wet’ method (optimal calcining temperature 600 °C) and from HPV alumina (W. R. Grace & Co., surface area 500 m2∙g−1, pore volume 2.8 mL∙g−1 after pre-calcining at 600 °C) by heating in fluidizing N2 and n-C6F14 vapor at 600 °C was described in [92]. In the latter case, the support turned black presumably due to carbon deposition. In ethylene polymerization (90 °C, 38 bar) with Zr9 and Et3Al, the highest activities of 1.25 and 1.27 kgPE∙gCat−1∙h−1 were achieved for supports, which were prepared by ‘wet’ and gas-phase fluorination. After recalcining of ‘black’ support in the air at 600 °C, during which the black color turned back to white, catalytic activity increased to 2.18 kgPE∙gCat−1∙h−1.

5.2. Fluorinated SiO2/Al2O3

Silica/alumina appears not to be an efficient activating support per se (see Section 3.3); however, the chemical binding of fluorine proved to be able to fundamentally increase the ability of supports to provide activation of Group 4 metal complexes. When compared with other activating supports, it is this type of support that showed the best efficiency, which puts fluorinated SiO2/Al2O3 on the top of the activating supports for Group 4 metallocenes and post-metallocenes. Deep and comprehensive corporate studies of this type of support have been undertaken (and apparently are ongoing) for Chevron Phillips (previously Phillips Petroleum), Total, and INEOS, which are the leading polyolefin manufacturers.
These studies were aimed at finding the most effective solutions in the development of MAO- and borate-free activating supports, and were based on well-defined and commercially available inorganic phases with known porosity and surface area. However, these studies were conducted in fundamentally different directions, namely, fluorination of the commercial silica/alumina phases and surface treatment of commercial silicas by organoaluminums with subsequent thermal treatment and fluorination (optionally, the latter stage was omitted when alkylaluminum fluorides were used). Consequently, these different approaches (Scheme 16) are discussed separately in our review.
Note that some recent patents protect the somewhat unexpected type of activating supports, namely, fluorinated silica-grafted alumina (Al2O3/g-SiO2(F), see Section 5.2.3).

5.2.1. Fluorination of SiO2/Al2O3 (SiO2/Al2O3(F))

The common ‘wet’ method of the fluorination of silica/alumina, which was studied since the early 2000s by the chemists of Phillips Petroleum, is completely similar to the ‘wet’ method of the preparation of fluorinated γ-Al2O3. In early experiments, the samples of SiO2/Al2O3 MS 13-110 (W. R. Grace & Co., 13 wt% alumina, surface area 400 m2∙g−1, pore volume 1.2 mL∙g−1) were treated by aqueous NH4HF2 at different [F]/[Al] ratios [89]. After calcination at different temperatures, the supports were used for activation of the Zr9/Et3Al catalyst. The results of polymerization experiments have demonstrated that both [F]/[Al] ratio and calcination temperature affect the polymerization activity, and optimal calcination temperatures vary greatly for supports, which are prepared at different [F]/[Al] ratios (Figure 18). In the absence of fluorinated SiO2/Al2O3 support, the Zr9/Et3Al system was inactive.
Fluorinated SiO2/Al2O3 (MS 13-110, W. R. Grace & Co.), prepared under optimized conditions, was studied in the polymerization of ethylene using Zr2/Et3Al with and without pre-contacting of Zr2/Et3Al with hex-1-ene before interaction with an activating support [93]. The reaction of Zr2 with Et3Al and hex-1-ene (toluene/n-heptane, 30 min) prior to charging to the support resulted in a marked increase of activity (2932 vs. 1640 gPE∙gCat−1∙h−1). Under similar conditions, bis(η5-fluorenyl) and (η5-fluorenyl)-Cp ansa-zirconocenes Zr36 and Zr37 (Scheme 17a) have demonstrated lower activities.
An increase in catalytic activity was attributed in [93] to the formation of aluminum metallocyclic species (Scheme 17b). Despite the experimentally proven fact of the formation of similar species during the reaction of Zr2 with Et3Al [28,29], their possible role in further activation and polymerization processes is unclear.
The addition of arylboronic acids to the solutions of Zr2 and Zr2′, followed by mixing with SiO2/Al2O3(F)/iBu3Al activator, resulted in a marked increase of catalytic productivity in the case of Zr2′ [95]. One can assume that Zr2′ reacts with ArB(OH)2 with a formation of Cp2Zr–O–B(OH)Ar species, but the role of similar species in further transformations is unclear too.
Very promising results were obtained when Ketjen Grade B alumina was subjected to ‘wet’ (with final calcining at 500 °C) or gas-phase fluorination (at 600 °C), followed by gas-phase chlorination (CCl4, 600 °C). The activities of the Zr9/Et3Al system on these supports in ethylene polymerization reached 3.1 and 6.3 kgPE∙gCat−1∙h−1, respectively [92].

5.2.2. Treatment of SiO2 by Organoaluminums with Subsequent or Simultaneous Fluorination (SiO2/g-Al2O3(F))

The synthesis of type SiO2/g-Al2O3(F) supports and their use for activation of Group 4 metallocenes and post-metallocenes are described and discussed in a number of patents (owned by Elf Atochem/Total and INEOS) and Total in cooperation with the scientific group of Boisson [99,100,101,102,103,104] and only a few scientific articles [105,106,107].
The method of the preparation of SiO2/g-Al2O3(F) from silica, developed by Elf Atochem/Total, comprises the following stages (Scheme 18):
  • Treatment of silica with organoaluminum compound in the inert organic solvent;
  • Calcination of the product, at the final stage—in the presence of O2;
  • Fluorination using (NH4)2SiF6 and final calcination.
Fluorination is based on the reaction of HF, which was released at a high temperature (typically 450 °C) from the decomposition of the (NH4)2SiF6. An alternative method of fluorination is based on the reaction of silica with Et2AlF, followed by calcination. When using Et2AlF, there is no need for fluorination by (NH4)2SiF6; a sufficient amount of fluorine is introduced at the stage of the reaction between silica and Et2AlF.
An alternative approach to SiO2/g-Al2O3(F) was developed by INEOS [99,100]. The method of INEOS is based on the reaction of SiO2 with fluorinated alkoxides of alkylaluminum, followed by calcination (see below).
In Table 7, we tried to summarize the data on the synthesis of SiO2/g-Al2O3(F) supports and the results of their use for activation of Group 4 metal complexes in α-olefin polymerization. In addition to the zirconocenes Zr2, Zr3, Zr5, Zr9, Zr26, and Zr37 and constrained-geometry complex Ti8, already mentioned in our review, the complexes rac-[Me2Si(η5-inden-1-yl)]ZrCl2 (Zr38) and rac-[Me2Si(η5-2-Me-4,5-benzoinden-1-yl)]ZrCl2 (Zr39) have been studied. The structures of complexes, investigated in combination with SiO2/g-Al2O3(F) activating supports, are presented in Scheme 19.
Polymerization experiments with Zr5 were conducted at 80 °C with ethylene solutions in n-heptane (6 wt%); the [iBu3Al]/[Zr] ratio was ~2000 [105]. Ethylene/hex-1-ene copolymerizations were also performed; the activities of supports have demonstrated similar trends. In both series of experiments, catalytic activities depended significantly on the types of silica used for the preparation of SiO2/g-Al2O3(F). If the average pore diameter was too small (e.g., support based on silica SIL 3), there was no activity. However, no clear correlations between silica specific surface/pore volume and catalytic activity have been found. The comparison of polymerization kinetic curves for Zr5 supported on SiO2/g-Al2O3(F) and on solid MAO (SMAO) showed noticeable deactivation of the catalysts prepared with activating supports [105].
The results of the study of SiO2/g-Al2O3(F) with different fluorine contents (Table 3) as activating supports for Zr5 and Zr39 were reported by Boisson’ group in 2013 [107]. The results of four comparative polymerization experiments are included in Table 7; however, a number of the findings of this comprehensive study deserve a separate discussion. As shown in the Table 8, a significant variation of the Al content was detected for SiO2/g-Al2O3(F), whereas increasing the amount of (NH4)2SiF6 from 5 wt% to 20 wt% led to higher F content in the support. The treatment of silica with a higher amount of (NH4)2SiF6 resulted in a degradation of the silica grain.
As can be seen in Scheme 18, SiO2/g-Al2O3(F) should contain Al–F bonds; however, the formation of Si–F bonds on the surface of SiO2/g-Al2O3(F) is theoretically possible. Figure 19 shows that the spatial distribution of Al and F for the support AS1 (see Table 8) appears homogeneous; therefore, F atoms are mainly bonded to Al atoms. 27Al MAS NMR analysis of AS2, AS3, and AS4 after exposure to wet air showed road resonances at 50, 32, and 3.5 ppm for AS2 and the appearance of the signal at −15 ppm for AS3. This signal can be attributed to AlF3∙3H2O (the reference spectrum of this complex was also recorded). The signal at −15 ppm became the main for AS4. 19F MAS NMR spectrum of AS7 showed the presence of the signal of AlF3∙3H2O (−8 ppm) and the intence signal at 12 ppm. This last resonance is close to the signal assigned to VIAl(O5F) (9 ppm) and VIAl(O4F2) (20 ppm) by Fischer et al. [213].
The support AS1 was studied as an activator for the Zr5/iBu3Al system in ethylene polymerization and ethylene/hex-1-ene copolymerization (~20 mg AS1, 0.5 μmol Zr5, 1 mmol iBu3Al, 4 bar ethylene, 80 °C). High activities (3.4–110 kg∙mmol−1∙h−1, 39–900 g∙gCat−1∙h−1) and a remarkable hex-1-ene activation effect were observed. Tm of polymer decreased from 133 °C (PE) to 104 °C (copolymer with 56 mol% of hex-1-ene). Copolymerization experiments with AS2–AS7 supports and Zr5/iBu3Al showed a high increase in activity as the wt % of F is increased on the carrier. The support AS5 prepared using treatment of the silica with Et2AlF showed comparable properties to support AS2 made using 5 wt % of (NH4)2SiF6, which had a similar content of F. In separate polymerization tests at 10 bar of ethylene (500 mL of heptane, 1 μmol Zr5, 2 mmol iBu3Al, 73 mg of AS8, 80 °C, 6 mL of hex-1-ene), productivity of 3200 g∙gCat−1∙h−1 was achieved without the catalyst’ leaching, as evidenced by perfect morphology of the polymer obtained (Figure 20).
During the comparison of Zr5/AS/iBu3Al and Zr5/MAO catalysts, better incorporation of hexene was observed with the activating support. MWD and TREF analysis data indicated the presence of multiple active sites for the Zr5/AS/iBu3Al catalytic system. Similar comparative studies were also performed for other Group 4 metal complexes Zr9, Zr26, Zr3, and Ti8 (Scheme 19) with the use of supports AS6 and AS7 (Table 9).
The complex Zr9 showed a poor ability to insert hexene into the solution, but its behavior was highly improved when using the activating support. It should also be noted that activities (per mmol) of the iBu3Al/AS-activated complexes Zr3 and Ti8 in ethylene/hex-1-ene copolymerization were higher than activities of Zr3/MAO and Ti8/MAO. Better incorporation of hex-1-ene was also detected in all experiments with SiO2/g-Al2O3(F).
At first glance, SiO2/g-Al2O3(F) has demonstrated promising characteristics of the activating supports for zirconocenes; the catalysts obtained could have competed with conventional SiO2/MAO supports. However, more rapid decay of the polymerization rate when using Zr5 on an activating support in comparison with Zr5/SMAO [105] has encouraged the research group of Boisson to perform a more thorough study of the kinetic behavior of SiO2/g-Al2O3(F)-activated Zr5 and Zr9 in the polymerization of ethylene [106]. In this study, three types of silica (Table 10) were used for the preparation of activating supports by the method described in [107]. The properties of fluorinated support M3703/079A, prepared from Grace Sylopol 952, are presented in the fourth row of Table 10. The complexes were pre-contacted with the supports for 10 min before introduction in the reactor. In all the reactions, around 60 mg of AS was used and the content of zirconocene was 0.4 wt%. Polymerization experiments were conducted at 80 °C and an ethylene pressure of 10 bar, in 0.1M hex-1-ene solution in n-heptane.
The study of the evolution of the Mw as a function of the reaction time using Zr5/AS/iBu3Al showed that the Mw and ÐM decrease as the reaction rate progresses. From the deconvolution results, it was shown that the rate of the formation of high-MW fraction tends to decrease significantly after 15 min of reaction, which corresponds to the beginning of the catalyst deactivation. Replacement of iBu3Al by Et3Al had no effect. However, Zr9 seemed to respond quite differently to iBu3Al and Et3Al. In the absence of a sufficient quantity of alkylaluminum, the catalyst became active, but rapidly reached a maximum rate that was followed by total deactivation of in 20–30 min. In the case of iBu3Al as a co-catalyst, the addition of approximately 500–1000 mol Al/mol Zr was sufficient to stabilize the reaction rate at an intermediate value close to the maximum rate for an hour. These general trends were also observed with Et3Al as a co-catalyst to Zr9, although as with the Zr5, much less Et3Al was required than iBu3Al. Also, in the case of Et3Al, the stable reaction rates that were obtainable were higher than those that were realized with iBu3Al. However, an increase in the concentration of Et3Al has led to a significant decrease in activity due to the formation of heterobimetallic dormant species [106]. The results of the comparative study of two common pre-catalysts Zr5 and Zr9 showed that the influence of the pre-catalyst nature on catalytic properties of the complex system, comprising Group 4 metal complex, SiO2/g-Al2O3(F) and alkylaluminum are still not understood.
Besides the η5-ligand (L) environment in L2ZrX2, the nature of X can also affect the catalytic activity. Therefore, for example, [EBTHI]ZrF2 (Zr1″) demonstrated higher activity than Zr1 and [EBTHI]ZrMe2 (Zr1′) under the same reaction conditions [103].
The high efficiency of SiO2/g-Al2O3(F) activating supports caused competing polyolefin companies to search for alternative synthetic approaches to these phases to circumvent the Elf Atochem/Total patents. Such a search had proved successful during the studies of the chemists of INEOS who proposed the reaction of silica with Et2Al(OR) (R = C6F5, CH2CF3, CH(CF3)2), followed by calcination, as an alternative synthetic pathway to SiO2/g-Al2O3(F) [99] (Scheme 20). The [F]/[Al] ratios in these supports amounted to 1.96–2.26; these values were determined by the temperature of the pre-calcining of SiO2 (CS2050 from PQ Corp.) in the range of 250–450 °C.
These supports were compared with activating support SiO2/g-Al2O3(F), which was prepared by the method of Total with the use of (NH4)2SiF6 ([F]/[Al] = 1.68) [217], in copolymerization of ethylene and hex-1-ene in slurry and gas-phase experiments. Constrained-geometry complex (C5Me4SiMe2NtBu)Ti(η4-1,3-pentadiene) (Ti8″) was selected as a pre-catalyst. In slurry polymerization experiments, Ti8″ was pre-mixed with iBu3Al ([Al]/[Ti] = 50) and added to 0.1 g of SiO2/g-Al2O3(F) (Ti loading of 30 μmol/gAS), and polymerization was conducted in 1.7 L of isobutane containing 0.25 mmol iBu3Al at 90 °C; an H2/ethylene ratio of 0.2 mol% and ethylene pressure of 10 bar were maintained for 1 h. The samples of SiO2/g-Al2O3(F), which were prepared with the use of Et2Al(OC6F5) and (NH4)2SiF6, have demonstrated close activities (2.0–3.0 kgPE∙gCat−1∙h−1). Before gas-phase polymerization experiments, solid catalysts were prepared by the above-described pre-mixing of the catalyst components, followed by washing with n-hexane and drying in vacuo. H2/ethylene and hex-1-ene/ethylene ratios were maintained at 0.13 and 0.55 mol%, respectively, during polymerization at 80 °C and 10 bar. In these experiments, catalytic activities were close to 0.5 kgPE∙gCat−1∙h−1 [99]. Comprehensive rheological studies of copolymers [100] showed that novel supported catalysts provide the formation of polymers with improved rheological properties. In particular, higher viscosities were observed at lower shear rates, providing better bubble stability and lower viscosities at higher shear rates, resulting in better processability in the extruder.
The further development of the activating supports of SiO2/g-Al2O3(F) type by INEOS [101] consisted of the treatment of calcined SiO2 (Sylopol 332, W. R. Grace & Co.) by 2,4,6-trimethylpyridine and POCl3, followed by vacuum-drying (and optionally additional heating at 130 °C) and high-temperature fluorination with (NH4)2SiF6. The comparison of these supports with SiO2/g-Al2O3(F), prepared by the method of Total [217], in activation of Zr5 and Ti8″ for ethylene/hex-1-ene copolymerization showed marked increase in catalytic activities: 927 vs. 340 gPE∙gCat−1∙h−1 for Zr5 and 500 vs. 410 gPE∙gCat−1∙h−1 for Ti8″. As can be seen in Scheme 19, a relatively small number of Group 4 metal complexes have been studied in MAO- and borate-free ethylene (co)polymerization using SiO2/g-Al2O3(F) supports. It can be assumed that the studies of other pre-catalysts can provide a breakthrough in the further development of single-site supported catalysts, which are suitable for use in the polyolefin industry.

5.2.3. Silylation of Alumina with Subsequent Fluorination (Al2O3/g-SiO2(F))

‘Inverted’ phase Al2O3/g-SiO2 can be prepared, for example (so-called ‘SIRAL’ phases), by the hydrolysis of a solution of aluminum hexanolate in hexanol with deionized water at 90 °C, followed by mixing the filtered alumina suspension with a solution of orthosilicic acid, and spray-drying. Hydrothermal treatment (5 h at 180 °C) before spray-drying results in the formation of phases with higher acidity [218]. Modification of alumina through silica addition creates (i) strong LAS through isomorphous substitution of tetrahedral Si4+ by Al3+ ions and (ii) a small quantity of highly acidic BAS, which are thought to be bridged OH species similar to those found in zeolites [218]. The surface structures of SIRALs with different [Al]/[Si] ratios are schematically presented in Figure 21.
In the first patent of Chevron Phillips in this field [94], Al2O3/g-SiO2(F) was prepared by the impregnation of commercial SIRAL 28M (Sasol) Al2O3/g-SiO2 with NH4HF2 in methanol with subsequent calcining. In ethylene polymerization on Zr32/iBu3Al, Al2O3/g-SiO2(F) provided better activity; it was far superior to that of Al2O3/SiO2(F) and almost twice that of sulfated alumina [94]. In ethylene homopolymerization on Zr40 (Scheme 21), activated by iBu3Al, catalytic activity depended on the alumina to silica ratio, varying from 1.69 kgPE∙gCat−1∙h−1 (fluorinated γ-Al2O3) to 7.30 kgPE∙gCat−1∙h−1 (Al2O3/g-SiO2(F) with Al2O3/SiO2 = 1.5:1 by weight) [94]. Undoubtedly of interest are the results of the comparative study of catalytic activity of Zr5/iBu3Al and Zr40/iBu3Al on different activating supports in ethylene/hex-1-ene copolymerization (Table 11).
This table also contains the tan δ (the ratio of loss and storage moduli, determined at 190 °C by small-strain (10%) oscillatory shear measurements, the shear frequency of 0.1 s−1) related to the presence of long-chain branches (LCB) in copolymer (higher tan δ means that the polymer relaxes easily, with little storage of the strain, and that the polymer has relatively lower LCB).
The synthesis of Al2O3/g-SiO2(F) having varied [Al]/[Si] ratios was also described in [94]. Pre-calcined (600 °C) Alumina A (W. R. Grace & Co., surface area ~300 m2∙g−1, pore volume 1.3 mL∙g−1) was treated with different amounts of Si(OEt)4 in MeOH and further with NH4HF2. The product was calcined in N2 at 600 °C. The highest activity for the Zr40/iBu3Al system in ethylene polymerization (6.3 kg∙gCat−1∙h−1) was observed for Al2O3/g-SiO2(F) with Al2O3/SiO2 ratio of 7.3 by weight.
Very high catalytic activities in the polymerization of ethylene were achieved for the Zr40/iBu3Al system when using fluorinated-chlorinated silica-coated alumina as an activating support [96]. This support was prepared by the treatment of Alumina A (W. R. Grace & Co., surface area 300 m2∙g−1, pore volume 1.2 mL∙g−1) with Si(OEt)4/iPrOH, followed by drying and calcining at 500–900 °C. The chlorination stage (4 h) involved injecting and vaporizing CC14 into the gas stream used to fluidize the Al2O3/g-SiO2 during calcination at 500 °C every five minutes. The fluorination step was conducted in a similar way using tetrafluoroethane within 4.5 h. The maximum activity of 17.83 kgPE∙gCat−1∙h−1 was demonstrated by Zr40, which was activated by iBu3Al and Al2O3/g-SiO2(F,Cl) containing 4 wt% of Cl and 7 wt% of F at 95 °C and 28 bar of ethylene [96]. A higher efficiency of Al2O3/g-SiO2(F,Cl) in comparison with other activating supports in ethylene polymerization was also detected for Zr32, Zr35, and Hf9 [98].
Another interesting application of the Al2O3/g-SiO2(F,Cl) support was the activation of the mixture of pre-catalysts Hf6 and Zr41 in ethylene/hex-1-ene copolymerization [96]. The amounts of 1.2 mg of Hf6, 1.4 mg of Zr41, and 150 mg of support were used, and copolymerization was conducted under similar conditions, with the addition of 5 g of hex-1-ene and 175 ppm of the molecular hydrogen. In this experiment, the catalytic activity of 2.19 kgPE∙gCat−1∙h−1 was achieved. In propylene polymerization experiments (70 °C, 31 bar) using the Zr42/iBu3Al/Al2O3/g-SiO2(F,Cl) system, catalytic activities reached up to 5.6 kgPP∙gCat−1∙h−1 [96].
Particularly noteworthy are relatively recent studies of ‘mixed’ catalytic systems, conducted by the chemists of Chevron Phillips [219,220]. The very idea of these studies was to develop a heterogeneous system containing both Ziegler–Natta and zirconocene catalysts. As described in [219,220], Al2O3/g-SiO2(F) was prepared from Alumina A by treatment with Si(OEt)4/iPrOH, followed by drying, calcining at 600 °C, fluorination by NH4HF2/MeOH, and calcination. The Ziegler–Natta catalyst was prepared by the reaction of Al2O3/g-SiO2(F) with Bu2Mg (toluene, 90 °C, 3 h), followed by treatment with TiCl4 (90 °C, 3 h). Activation of zirconocenes Zr33, Zr40, and Zr41 by iBu3Al and this support gave highly efficient catalysts of the copolymerization of ethylene with hex-1-ene [219,220]. When using Zr40/iBu3Al at 90 °C and 27 bar of ethylene, catalytic activities were 49.4 and 44.9 kg∙gCat−1∙h−1 in the absence and in the presence (880 ppm) of the molecular hydrogen. In the absence of zirconocene, Al2O3/g-SiO2(F)-based Ziegler–Natta catalyst was less active (26.5 and 10.8 kg∙gCat−1∙h−1 in the absence and in the presence (1000 ppm) of the H2).
Another method of the preparation of ‘Ziegler–Natta’ Al2O3/g-SiO2(F) catalysts/supports was based on the interaction of Al2O3/g-SiO2(F) with solutions of active metal chlorides and MgCl2 in THF, followed by centrifugation and drying in vacuo [219,220]. When TiCl4 was used as active metal chloride, a very active catalyst was obtained (5.27 kg∙gCat−1∙h−1); however, after absorption of the Zr40/iBu3Al catalytic, activity tripled (~15 kg∙gCat−1∙h−1). Zirconocenes Zr33 and Zr41 were studied in combination with this type of active support with the aim of obtaining polyethylenes having a broad spectrum of characteristics.
The third method of the preparation of Ti/Mg Al2O3/g-SiO2(F), proposed in [219,220], is close to the common method of the preparation of TMZNCs by the reaction of Mg(OEt)2 with TiCl4 [221]: Mg(OEt)2 reacted with Al2O3/g-SiO2(F) in toluene, with subsequent reflux with TiCl4. Zirconocenes Zr33 and Zr41 with this type of Ti/Mg Al2O3/g-SiO2(F) support have demonstrated maximum activities of 22.1 and 11.7 kg∙gCat−1∙h−1, respectively, in copolymerization of ethylene with hex-1-ene in the presence of 100 ppm of H2. In almost all samples of TMZNC/zirconocene catalysts, supported on Al2O3/g-SiO2(F), the [Zr]/[Ti] ratio was ~3:10, from which it follows that the zirconocene component of the binary catalyst is far more active.

5.2.4. Fluorinated SiO2/Al2O3 in Oligomerization of Higher α-Olefins

SiO2/Al2O3(F), prepared by the ‘wet’ method from MS13-110 silica/alumina (W. R. Grace & Co., 13 wt% Al2O3, surface area 400 m2∙g−1, pore volume 1.2 mL∙g−1), was used as an activating support for the series of Group 4 metallocenes (17 examples) in a comparative study of the catalytic activity in oligomerization of oct-1-ene [97]. iBu3Al was used as an organoaluminum activator for L2MCl2 before the addition of the activating supports. The structures of the most active pre-catalysts are presented in Scheme 22; the results of the oligomerization experiments are given in Table 12. As can be seen from this table, TOF of 105 h−1 and higher can be achieved when using SiO2/Al2O3(F), which is comparable to activities of the best oligomerization catalysts under homogeneous conditions with perfluoroborate activation [222,223].
However, this topic has not received further development. It can be assumed that SiO2/Al2O3(F)-based α-olefin oligomerization catalysts have a fundamental flaw resulting from the high acidity of the support. The consequence of such acidity is a possibility of cationic skeletal rearrangements during oligomerization, which leads to the deterioration of the characteristics of α-olefin oligomers in terms of their use for the production of engine oil feedstocks.

5.3. Fluorinated Compositions of SiO2 or Al2O3 with Other Metal Oxides

The preparation of fluorinated supports, based on silica and Group 4 metal oxides, was described in the patent of Phillips Petroleum [90]. The silica/titania, prepared by the cogellation method (8 wt% of TiO2, surface area of ~450 m2∙g−1) was calcined at 600 °C, treated with NH4HF2, and calcined at different temperatures. The temperature of the final calcination had a significant effect on the efficiency of these phases when being used as activating supports for the Zr9/Et3Al catalyzed polymerization of ethylene. Therefore, for example, when using supports calcined at 600 and 450 °C, catalytic activities were 1164 and 2837 gPE∙gCat−1∙h−1. Note that fluorinated TiO2 had demonstrated the lack of activation in Zr9/Et3Al catalyzed polymerization of ethylene. Under the same conditions, the samples of fluorinated silica/zirconia have demonstrated a higher efficiency in comparison with fluorinated silica/titania (up to 5 kgPE∙gCat−1∙h−1).
Impregnation of alumina by Zr(OBu)4 with subsequent calcination (dry N2) and fluorination by perfluorohexane at 600 °C resulted in the formation of fluorinated alumina/zirconia with low own activity in ethylene polymerization after treatment with Et3Al (35 gPE∙gCat−1∙h−1) [91]. When using fluorinated alumina/zirconia as an activating support for Zr9, the catalyst productivity increased to 1.38 kgPE∙gCat−1∙h−1, which was comparable to the efficiency of the fluorinated alumina support [91]. The presence of high-MW ‘Ziegler–Natta’ PE, formed on fluorinated alumina/zirconia catalytic centers, without the participance of metallocene, allows to consider the product of polymerization as a prospective PE composite.

5.4. Chlorinated Metal Oxides

Ketjen Grade B alumina and Davison Grade 952 silica were chlorinated by the reaction with CCl4 vapor at 600 °C. After the addition of a toluene solution of Zr9 and Et3Al, a moderately active ethylene polymerization catalyst was formed by chlorinated alumina (~103 gPE∙gCat−1∙h−1), whereas chlorinated silica was found to be inactive [91]. Impregnation of alumina by Zr(OBu)4 with subsequent calcination (dry N2) and chlorination resulted in a solid phase, which was active in ethylene polymerization after treatment with Et3Al; the addition of Zr9 resulted in a further increase of activity (862 and 1484 gPE∙gCat−1∙h−1, respectively). The efficiency of chlorinated silica/zirconia support in the activation of Zr9 was moderate [91].
The studies of other methods of the chlorination of HPV alumina were reported in the Chevron Phillips patent [92]; SO2Cl2 at 300 °C and AlCl3 at 250 °C were found to be fairly efficient chlorinating agents (activities of 459 and 401 gPE∙gCat−1∙h−1 in ethylene polymerization when using Zr9/Et3Al). Chlorination of alumina by CCl4 under optimized conditions allowed to reach the catalytic activity of 2.8 kgPE∙gCat−1∙h−1 for Zr9/Et3Al [92].

6. Other Inorganic Supports

There are only a few articles that address patents devoted to the study of inorganic activating supports distinct from the supports discussed above. In 1995, Soga and colleagues demonstrated the ability of solid heteropolyacids H3[PMo12O40] and H5[PMo10V2O40] to activate Zr2 and Zr5 in the polymerization of ethylene and propylene, respectively [224]. Catalytic activities were low (up to 23 and 3.7 kg∙molZr−1∙h−1). Dimethyl derivative Zr2′ was inactive, on the basis of which they concluded that activation of zirconocene pre-catalyst L2ZrCl2 proceeds through Cl abstraction.
Impregnation of the silica, alumina, or SiO2/Al2O3 by different metal chlorides with subsequent calcination allowed to obtain solid supports with different activating efficiency [225,226,227]. Active supports were prepared from Ketjen Grade B alumina using Cu(II), Sn(IV), Ag(I), Nb(V), Mn(II), W(VI), La(III), Nd(III), Sb(V), Zn(II) [225], Mo(VI) [226], and Ni [227] salts. The activation of Zr9/Et3Al in comparative experiments on the homopolymerization of ethylene (38 bar, 90 °C) allowed the identification of the most efficient systems, which were based on Sn, Ag, and Zn chlorides. When the loading of ZnCl2 was 20 wt% of alumina, after calcining with CCl4 and impregnation with Zr9/Et3Al, the catalytic activity of 11.8 kg∙gCat−1∙h−1 was achieved [225].
Cation-exchanged fluorotetrasilicic mica (Mn+-mica, Mn+ = Na+, Mg2+, Fe3+, Co2+, Ni2+, Cu2+, Zn2+) were prepared and employed as activating supports in the ethylene polymerization or ethylene/hex-1-ene copolymerization (7 bar, 60 °C) in the presence of L2ZrCl2/iBu3Al [228]. Mn+-mica were calcined at different temperatures, and 100 mg samples of the support were contacted with 20 μmol of Zr2 or Zr9 with subsequent treatment by iBu3Al. Catalytic activities in Zr2 catalyzed homopolymerization of ethylene substantially depended on the nature of Mn+ (Fe3+-mica demonstrated the highest productivity of 174 gPE∙gCat−1∙h−1) and the temperature of calcination (200 °C was the optimal value). The authors proposed that the active sites were probably formed by the reactions of Cp2ZrCl2 and the exchanged cations Mn+ since the activity was significantly increased by replacing Na+ ions with M2+ (M2+ = Mg2+, Co2+, Ni2+, Cu2+, Zn2+) and Fe3+. This work is of particular interest, but it is not applicable in practice due to the extremely low activity of the catalysts.

7. Organochemical Functionalization of Inorganic Surfaces and Organic Polymers

7.1. Organochemical Functionalization of Inorganic Phases

The very idea of low-cost surface modification of SiO2, Al2O3, SiO2/Al2O3, and other inorganic phases by C6F5Br and its analogs with a formation of perfluoroaryl-grafted oxides was patented by W. R. Grace & Co. in 1999 [229]. However, this patent did not include any data on catalytic activity.
In the early 2000s, the activation of post-metallocenes Zr46Zr49 (Scheme 23) by surface-modified heteropolyacids and Et3Al was studied by Fujita’s group [230]. The activating support of the formula (Ph3C)mHn[PMo12O40]∙8H2O (average m/n~2:1) was obtained by the reaction of H3[PMo12O40]∙28H2O with Ph3CCl in acetone with subsequent evaporation, washing by toluene, hexane, and drying at 100 °C under vacuum. After the addition of 10 μmol of Zr complex to the mixture of 5 μmol of the support and 1 mmol of Et3Al, fast ethylene polymerization was observed under atmospheric pressure and 25 °C, and activities of Zr46Zr49 were 1.02–5.64 kg∙molZr−1∙h−1; complex Zr49 demonstrated the highest activity. In the cases of Zr47 and Zr48, bimodal PEs were obtained, whereas complex Zr49 catalyzed the formation of PE with abnormally narrow MWD (ĐM = 1.45). The nature of the supports and activation mechanisms were not studied and discussed in [230]. One can assume that the reaction of (Ph3C)mHn[PMo12O40]∙8H2O with Et3Al results in the formation of soluble dimeric ethylalumoxane and Mo–O–AlEt2 species and does not affect Ph3C+—[PMo12O40]m− interactions. Activation of L′ZrCl2 proceeds through alkylation by Et3Al and subsequent Zr→CPh3 ethyl transfer with a formation of L2ZrEt+ and 1,1,1-triphenylethane. This activation mechanism is very similar to the conventional mechanism of the activation of Group 4 metal complexes by [Ph3C]+[B(C6F5)4] [9].
In 2007 [231], Jones and colleagues proposed the further development of the idea of solid activating supports, which consisted in surface modification of silica (mesoporous SBA-15) by perfluoroalkyl–SO3H groups using fluorinated sultone precursor [232] (Scheme 24a). In the presence of Me3Al, this support activated pre-catalyst Zr17′ in ethylene polymerization (25 °C, 4 bar), yielding productivities up to 1000 kgPE∙molZr−1∙h−1 without reactor fouling. During the study of the activation mechanism, possible ways of MAO formation and direct interaction of –SO3H groups with Zr17′ were ruled out, which led to the structure of the activated catalyst presented in Scheme 24b.
In 2011, Taoufik and colleagues showed that the treatment of silica surface with iBu3Al in Et2O afforded a well-defined bipodal surface species [(≡SiO)2Al(iBu)(Et2O)] [149] (Scheme 25). In further studies, this material was treated with C6F5OH (2.5 eq.) in benzene in the presence of one eq. of N,N-diethylaniline; repeated washing and vacuum-drying resulted in obtaining of activating support AS-1 (Scheme 25) [54]. The structure of the surface species was confirmed by diffuse reflectance infrared Fourier transform (DRIFT) spectroscopy and 1H, 13C, 19F, and 27Al CP-MAS NMR.
The catalytic performance of the support was studied in ethylene polymerization with and without hex-1-ene, using common Zr5 and Zr9 pre-catalysts and iBu3Al. Reference tests were performed with silica-supported MAO (sMAO). The results of the experiments are presented in Table 13 [54].
The Zr5/AS-1 system was found to be more active in comparison with Zr5/sMAO; a major comonomer effect was observed. The initial activity was higher when using the activating support AS-1, but deactivation was observed during the course of polymerization. This deactivation could be ascribed to aryloxide transfer to Zr.
During the search of weakly coordinating surface anions, Conley and colleagues studied the reaction of Al(OC(CF3)3)3∙PhF with calcined silica in perfluorohexane that resulted in the formation of highly acidic [≡SiOH Al(OC(CF3)3)3] species [233] (Scheme 26a). Two years later, they published the results of further studies of this reaction [234]. It turned out that the reaction of Al(OC(CF3)3)3∙PhF with partially dehydroxylated silica in fluorobenzene solvent, which induces proton transfer, results in the formation of [≡SiOAl(OC(CF3)3)2(O(Si≡)2) species (A in Scheme 26b) with extremely high Lewis acidity.
The 27Al{1H} MAS NMR spectrum of A contained a single signal that simulates as an isotropic chemical shift (δiso) of 74 ppm and a quadrupolar coupling constant (CQ) of 18.0 MHz; DFT optimization of A predicted CQ = 18.7 MHz. The calculated fluoride ion affinity (FIA) for A was 528 kJ∙mol−1, which is significantly larger than the calculated FIA for B(C6F5)3 (448 kJ∙mol−1. The reaction of A with Cp2Zr(13CH2)2 was complex. A 13C{1H} CP-MAS NMR spectrum of Cp2Zr(13CH3)2/A contains four signals at 38 (Cp2Zr–13CH3+ in species B, Scheme 26b), 23 (species C), 3 (Si–13CH3 in species D), and –11 (Al–13CH3 in species B and D) ppm. The pressurizing of Cp2Zr(13CH3)2/A in cyclohexane by 20 bar of ethylene resulted in the formation of PE with an activity of 78.6 gPE∙gCat−1∙h−1 [234]. The findings of this study are of great significance for understanding the chemistry of Si/Al-based activating supports and for developing new-generation MAO- and borate-free supported Group 4 metallocene and post-metallocene catalysts.

7.2. Functionalized Organic Polymers

Nafion (Scheme 27) is the first of a class of synthetic polymers with ionic properties; its molecular structure comprises poly(tetrafluoroethylene) backbone and perfluorovinyl ether groups terminated with sulfonate groups as side chains [235]. The presence of highly acidic –SO3H groups and permeation for gases led the chemists of Mitsui Petrochemicals to the suggestion that Nafion can be used as an activating support for L2ZrCl2/R3Al systems [236]. This assumption was confirmed by ethylene polymerization experiments (75 °C, pressure not indicated) using Zr5/iBu3Al; however, substantially cheaper sulfated polymer Amberlist (Scheme 27) showed comparable ability to activate zirconocene. The maximum activity of this catalyst can be roughly estimated as ~2 kg∙gCat−1∙h−1. Note that MgCl2-supported Zr5/iBu3Al under the same conditions was about an order of magnitude less active [236].

8. Ionic Liquids

Ionic liquids (ILs) have long attracted the attention of researchers as carrier phases for different catalysts [237]. If the reaction proceeds in a non-polar medium, the use of ILs simplifies the separation of the products and recycling of the catalysts. In particular, ILs have already found applications in the cationic oligomerization of α-olefins [238,239,240,241,242]. Polymerization of ethylene and α-olefins with the use of coordination catalysts and ILs can be considered to be a heterogeneous process. The term ‘support’ implies belonging to the solid phase, so that the inclusion of ILs in this review is not entirely appropriate. However, ILs represent a convenient reaction medium for the study of activation processes, and we chose to include this short section in the present review.
In 1990 [243], Carlin and Wilkes reported the results of the study of ambient-temperature chloroaluminate molten salt AlCl3∙MEIC (MEIC—1-ethyl-3-methylimidazolium chloride) as applied to activation of Ti2. When bubbling ethylene at 1 atm through 1.1:1.0 AlCl3∙MEIC (6.8 g) containing 22 mg Ti2 and 0.118 g Al2Me3Cl3, 20 mg PE was obtained. Similar Zr (Zr2) and Hf (Hf1) complexes were inactive under similar conditions. The authors proposed the formation of the species of the formula Cp2TiMe-(μ-Cl)-AlCl3 that are able to dissociate with a formation of active Cp2TiMe+ species. The inertness of Zr2 and Hf1 was attributed by the authors to the higher strength of the M–Cl bond for M = Zr, Hf in comparison with Ti, but this explanation contradicts some reports, e.g., [244].
From 2007 to 2013, a number of articles on the subject of metallocene activation using ILs were published by the research group of Ochędzan-Siodłak [245,246,247,248,249]. In [245], the catalytic behavior of Ti2 in ethylene polymerization with 1-n-butyl-3-methylimidazolium tetrachloroaluminate [BMIM]+[AlCl4] when using MAO, Et2AlCl, and Et3Al as alkylating agents was investigated. During polymerization experiments (n-hexane, 30 °C, 1 bar), a marked difference in the distribution of PE between IL and n-hexane phases was observed (Figure 22). At the beginning of the polymerization, the ionic liquid phase becomes white and swells considerably as the polyethylene appears. After 5–10 min, the hexane becomes a white suspension as the polyethylene is progressively shifted from the ionic liquid phase. The best results were obtained using Et2AlCl, where the greatest amount of the polymer is in the hexane phase, whereas the amount of the polymer in the ionic liquid decreases and slowly reaches a constant level. After separation, this [BMIM]+[AlCl4]/Ti2/Et2AlCl phase maintained the ability to catalyze ethylene polymerization, at 30 °C and 1 bar activities, based on the weights of PE collected from the n-hexane phase, which were 9.0 and 8.8 kgPE∙molTi−1∙h−1 for freshly prepared and recycled catalysts, respectively. When using 1-ethyl-3-methylimidazolium tetrachloroaluminate and Ti2, the distribution of PE significantly deteriorated; even in the presence of Et2AlCl, the main fraction of PE remained in the IL phase [246].
The yield and distribution of PE between IL and aliphatic phases substantially depended on the nature of IL and the duration of polymerization. Comparative studies of ILs [C8-mim]+[AlCl4], [C8-β-mpy]+[AlCl4], and [C8-γ-mpy]+[AlCl4] (Scheme 28) as activators for Ti2/EtAlCl2 in ethylene polymerization showed the advantage of pyridinium-based ILs both in activity and in PE distribution at a low Al/Ti ratio of 67 (Table 14) [247]. The best results were obtained with use of [C8-mim]+[AlCl4], for which the optimal Al/Ti molar ratio was 133. In all experiments, the use of the larger amounts of the EtAlCl2 activator had a disadvantageous impact on the polymer transfer to the hexane phase. It was also shown that an active catalyst was formed and immobilized in IL phases, and no catalyst leakage was observed [247].
The replacement of n-octyl substituent in [C8-mim]+[AlCl4] by the PhCH2CH2– fragment was found to be suitable to immobilize Ti2/EtAlCl2, and PE was gathered mainly in the hexane phase [248]. An increase in the reaction time resulted in a considerable increase in the amount of PE and improved the PE transfer from the IL to hexane. Also, PhCH2CH2-substituted IL was more stable at high temperatures in comparison with [C8-mim]+[AlCl4].
In 2014, Ochędzan-Siodłak and Dziubek came up with the idea of the immobilization of Group 4 metal catalysts on IL-modified silica (SIL) [249]. The modification of SiO2 was based on the reaction of surface Si–OH groups with (EtO)3Si-finctionalized imidazolium chloride, followed by treatment with AlCl3 and Et2AlCl (Scheme 29). A comparative study of the pre-catalysts Ti2, bis[N-(salicylideno)anilinato]-titanium(IV) dichloride (Ti19), and N,N′-ethylenebis[5-chloro-salicylideneiminato]titanium(IV) dichloride (Ti20) showed that titanocene dichloride Ti2 catalyzed slurry polymerization of ethylene (30 °C, 5 bar, n-hexane) with a higher efficiency (activity up to 7200 kgPE∙molTi−1∙h−1) in comparison with other pre-catalysts (180 kgPE∙molTi−1∙h−1 for Ti19, 410 kgPE∙molTi−1∙h−1 for Ti20).
The Ti2/Et2AlCl/SIL catalytic system was of spherical morphology (Figure 23a) and provided a formation of PE particles that have the shape of porous granules (Figure 23b). During polymerization experiments using Ti2/Et2AlCl/SIL, a linear PE was obtained (Tm = 138–141 °C, degree of crystallinity 62–82%). Therefore, for example, in the experiment, during which the highest catalytic activity was achieved, the resulting PE had the following characteristics: Mw = 478 kDa, Ðm = 3.1, Tm = 141 °C, and bulk density of 298 g∙dm−3.

9. Conclusions and Outlook

In our review, we summarized scientific periodicals and the patent literature related to the study of the MAO- and borate-free activating supports for Group 4 metallocenes and post-metallocenes and their use in catalytic polymerization of ethylene and α-olefins. In the frameworks of the current views on the mechanism of single-site polymerization catalysis, activation of Group 4 metallocenes and post-metallocenes lies in the formation of alkyl-metal cations, which are capable of α-olefin coordination/insertion. The mechanism of the formation of similar cations depends on the chemical nature of the activating support. If the support contains only strong Lewis acidic sites, activation proceeds via M(IV)→support alkyl migration (MgCl2, γ-Al2O3 calcined at high temperatures). If the support contains mainly Brønsted acidic sites, activation proceeds via protonolysis of M(IV)–alkyl fragment, and catalytic activity depends on the strength of M(IV)–O(support) bonding (sulfated alumina). In the case of weak Brønsted acidic sites such as ≡Si–OH, inactive ≡Si–O–M(IV) species are formed. These three types of processes on the surface activating support are clearly represented in Figure 24 by an example of Cp2ZrMe2 activation [180].
The surface chemistry of the activation of Group 4 metal complexes was studied by different methods, including MAS NMR, EXAFS, FT-IR spectroscopy, and DFT modeling. Among activating supports, γ-Al2O3, sulfated alumina, fluorinated alumina, and fluorinated alumina-grafted silica are the most thoroughly studied in terms of their surface structure and activation chemistry. Experimental and theoretical investigations in this field were carried out by scientific groups led by T. Marks, C. Boisson, V. Zakharov, M. Conley, and others. Among well-studied activating supports, sulfated alumina was in the top position in catalytic activity, and values of more than 17 kg∙gCat−1∙h−1 have been reported [201].
However, recent studies of the chemists of Chevron Phillips led to the development of a new type of activating support, namely, fluorinated silica-grafted alumina, Al2O3-g-SiO2(F). It is noteworthy that the sequential high-temperature chlorination and fluorination of Al2O3-g-SiO2(F,Cl) phases with even higher efficiency. A number of patents describe the use of these phases for the activation of Group 4 metallocenes and hybrid catalytic systems, which contain both single-site and Ziegler–Natta polymerization catalysts. Under optimal conditions and with proper selection of pre-catalysts, activities up to 50 kg∙gCat−1∙h−1 were achieved. Similar activating supports have not been thoroughly studied with respect to understanding mechanisms of activation and the nature of the active species and counterions. The presence of Cl in most active Al2O3-g-SiO2(F,Cl) supports allows us to draw some analogies with other Group 4 metal catalysts with the obvious, still unclear, but undervalued role of M–(μ-Cl)–Al bonding in α-olefin polymerization [250].
It should also be noted that the total number of Group 4 metallocenes and post-metallocenes, studied in ethylene and α-olefin polymerization with the use of activating supports, is relatively small, with only a few dozen pre-catalysts. That is orders of magnitude less than the total number of pre-catalysts studied with the use of conventional MAO and borate activators. It is safe to assume that further search and design of Group 4 metallocenes and post-metallocenes, which are best suited for the use of activating supports and common alkylaluminums, will allow the creation of efficient heterogeneous catalysts for the production of advanced polyolefins. During this search, model reactions of α-olefin oligomerization [251,252,253] with simple end-group analysis and DPn control as well as focusing on metal complexes, capable of homogeneous activation using R3Al [254], deserve particular attention.

Author Contributions

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

Funding

This research was supported by the Russian Science Foundation, Grant No. 21-73-30010.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jubinville, D.; Esmizadeh, E.; Saikrishnan, S.; Tzoganakis, C.; Mekonnen, T. A comprehensive review of global production and recycling methods of polyolefin (PO) based products and their post-recycling applications. Sustain. Mater. Technol. 2020, 25, e00188. [Google Scholar] [CrossRef]
  2. Qiao, J.; Guo, M.; Wang, L.; Liu, D.; Zhang, X.; Yu, L.; Song, W.; Liu, Y. Recent advances in polyolefin technology. Polym. Chem. 2011, 2, 1611–1623. [Google Scholar] [CrossRef]
  3. Chum, P.S.; Swogger, K.W. Olefin polymer technologies—History and recent progress at The Dow Chemical Company. Prog. Polym. Sci. 2008, 33, 797–819. [Google Scholar] [CrossRef]
  4. Sauter, D.W.; Taoufik, M.; Boisson, C. Polyolefins, a Success Story. Polymers 2017, 9, 185. [Google Scholar] [CrossRef] [Green Version]
  5. Geyer, R.; Jambeck, J.R.; Law, K.L. Production, use, and fate of all plastics ever made. Sci. Adv. 2017, 3, e1700782. [Google Scholar] [CrossRef] [Green Version]
  6. McDaniel, M.P. Chapter 3—A Review of the Phillips Supported Chromium Catalyst and Its Commercial Use for Ethylene Polymerization. Adv. Catal. 2010, 53, 123–606. [Google Scholar] [CrossRef]
  7. Shamiri, A.; Chakrabarti, M.H.; Jahan, S.; Hussain, M.A.; Kaminsky, W.; Aravind, P.V.; Yehye, W.A. The Influence of Ziegler-Natta and Metallocene Catalysts on Polyolefin Structure, Properties, and Processing Ability. Materials 2014, 7, 5069–5108. [Google Scholar] [CrossRef] [Green Version]
  8. Chen, C. Designing catalysts for olefin polymerization and copolymerization: Beyond electronic and steric tuning. Nat. Rev. Chem. 2018, 2, 6–14. [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. Sinn, H.; Kaminsky, W.; Vollmer, H.J.; Woldt, R. “Living Polymers” on Polymerization with Extremely Productive Ziegler Catalysts. Angew. Chem. Int. Ed. 1980, 19, 390–392. [Google Scholar] [CrossRef]
  11. Bochmann, M. The Chemistry of Catalyst Activation: The Case of Group 4 Polymerization Catalysts. Organometallics 2010, 29, 4711–4740. [Google Scholar] [CrossRef]
  12. Yang, X.; Stern, C.L.; Marks, T.J. Cationic Metallocene Polymerization Catalysts. Synthesis and Properities of the Cationic Metallocene Polymerization Catalysts. Synthesis and Properities of the First Base-Free Zirconocene Hydride. Angew. Chem. Int. Ed. 1992, 31, 1375–1377. [Google Scholar] [CrossRef]
  13. Parveen, R.; Cundari, T.R.; Younker, J.M.; Rodriguez, G. Computational Assessment of Counterion Effect of Borate Anions on Ethylene Polymerization by Zirconocene and Hafnocene Catalysts. Organometallics 2020, 39, 2068–2079. [Google Scholar] [CrossRef]
  14. Faingol’d, E.E.; Bravaya, N.M.; Panin, A.N.; Babkina, O.N.; Saratovskikh, S.L.; Privalov, V.I. Isobutylaluminum aryloxides as metallocene activators in Homo- and copolymerization of olefins. J. Appl. Polym. Sci. 2016, 133, 43276. [Google Scholar] [CrossRef]
  15. Tritto, I.; Zucchi, D.; Destro, M.; Sacchi, M.C.; Dall’Occo, T.; Galimberti, M. NMR investigations of the reactivity between zirconocenes and β-alkyl-substituted aluminoxanes. J. Mol. Catal. A Chem. 1990, 160, 107–114. [Google Scholar] [CrossRef]
  16. Bravaya, N.M.; Panin, A.N.; Faingol’d, E.E.; Saratovskikh, S.L.; Babkina, O.N.; Zharkov, I.V.; Perepelitsina, E.O. Isobutylalumoxanes as high-performance activators of rac-Et(2-MeInd)2ZrMe2 in copolymerization of ethylene with propylene and ternary copolymerization of ethylene, propylene, and 5-ethylidene-2-norbornene. Polym. Bull. 2016, 73, 473–491. [Google Scholar] [CrossRef]
  17. Tritto, I.; Boggioni, L.; Sacchi, M.C.; Dall’Occo, T. Novel aluminum based cocatalysts for metallocene catalyzed olefin polymerization. J. Mol. Catal. A Chem. 2003, 204–205, 305–314. [Google Scholar] [CrossRef]
  18. Kaminsky, W. Discovery of Methylaluminoxane as Cocatalyst for Olefin Polymerization. Macromolecules 2012, 45, 3289–3297. [Google Scholar] [CrossRef]
  19. Babushkin, D.E.; Semikolenova, N.V.; Zakharov, V.A.; Talsi, E.P. Mechanism of dimethylzirconocene activation with methylaluminoxane: NMR monitoring of intermediates at high Al/Zr ratios. Macromol. Chem. Phys. 2000, 201, 558–567. [Google Scholar] [CrossRef]
  20. Theurkauff, G.; Bader, M.; Marquet, N.; Bondon, A.; Roisnel, T.; Guegan, J.-P.; Amar, A.; Boucekkine, A.; Carpentier, J.-F.; Kirillov, E. Discrete Ionic Complexes of Highly Isoselective Zirconocenes. Solution Dynamics, Trimethylaluminum Adducts, and Implications in Propylene Polymerization. Organometallics 2016, 35, 258–276. [Google Scholar] [CrossRef]
  21. Joshi, A.; Zijlstra, H.S.; Collins, S.; McIndoe, J.S. Catalyst Deactivation Processes during 1-Hexene Polymerization. ACS Catal. 2020, 10, 7195–7206. [Google Scholar] [CrossRef]
  22. Parfenova, L.V.; Khalilov, L.M.; Dzhemilev, U.M. Mechanisms of reactions of organoaluminium compounds with alkenes and alkynes catalyzed by Zr complexes. Russ. Chem. Rev. 2012, 81, 524–548. [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] [PubMed] [Green Version]
  24. 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]
  25. 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]
  26. Baldwin, S.M.; Bercaw, J.E.; Henling, L.M.; Day, M.W.; Brintzinger, H.H. Cationic Alkylaluminum-Complexed Zirconocene Hydrides: NMR-Spectroscopic Identification, Crystallographic Structure Determination, and Interconversion with Other Zirconocene Cations. J. Am. Chem. Soc. 2011, 133, 1805–1813. [Google Scholar] [CrossRef] [Green Version]
  27. 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] [Green Version]
  28. Alt, H.G.; Denner, C.E.; Thewalt, U.; Rausch, M.D. Cp2Zr(C2H4)(PMe3), der erste stabile Ethylenkomplex des Zirkoniums. J. Organomet. Chem. 1988, 356, C83–C85. [Google Scholar] [CrossRef]
  29. Tamotsu, T.; Masato, M.; Masaki, K.; Masahiko, S.; Yasuzo, U.; Kozo, K.; Tokiko, U.; Swanson, D.R.; Negishi, E. Zirconocene–Alkene Complexes. An X-Ray Structure and a Novel Preparative Method. Chem. Lett. 1989, 18, 761–764. [Google Scholar] [CrossRef]
  30. Bochmann, M.; Sarsfield, M.J. Reaction of AlR3 with [CPh3][B(C6F5)4]:  Facile Degradation of [B(C6F5)4] by Transient “[AlR2]+”. Organometallics 1998, 17, 5908–5912. [Google Scholar] [CrossRef]
  31. Bolton, P.D.; Adams, N.; Clot, E.; Cowley, A.R.; Wilson, P.J.; Schroder, M.; Mountford, P. Synthesis and Ethylene Polymerization Capability of Metallocene-like Imido Titanium Dialkyl Compounds and Their Reactions with AliBu3. Organometallics 2006, 25, 5549–5565. [Google Scholar] [CrossRef]
  32. Götz, C.; Rau, A.; Luft, G. Ternary metallocene catalyst systems based on metallocene dichlorides and AlBu3i/[PhNMe2H][B(C6F5)4]: NMR investigations of the influence of Al/Zr ratios on alkylation and on formation of the precursor of the active metallocene species. J. Mol. Catal. A Chem. 2002, 184, 95–110. [Google Scholar] [CrossRef]
  33. Zaccaria, F.; Zuccaccia, C.; Cipullo, R.; Budzelaar, P.H.M.; Vittoria, A.; Macchioni, A.; Busico, V.; Ehm, C. Methylaluminoxane’s Molecular Cousin: A Well-defined and “Complete” Al-Activator for Molecular Olefin Polymerization Catalysts. ACS Catal. 2021, 11, 4464–4475. [Google Scholar] [CrossRef]
  34. Parfenova, L.V.; Balaev, A.V.; Gubaidullin, I.M.; Abzalilova, L.R.; Pechatkina, S.V.; Khalilov, L.M.; Spivak, S.I.; Dzhemilev, U.M. Kinetic model of olefin hydroalumination by HAlBui2 and AlBui3 in the presence of Cp2ZrCl2 catalyst. Int. J. Chem. Kinet. 2007, 39, 333–339. [Google Scholar] [CrossRef]
  35. 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]
  36. Culver, D.B.; Corieri, J.; Lief, G.; Conley, M.P. Reactions of Triisobutylaluminum with Unbridged or Bridged Group IV Metallocene Dichlorides. Organometallics 2022, 41, 892–899. [Google Scholar] [CrossRef]
  37. Negishi, E.; Kondakov, D.Y.; Choueiry, D.; Kasai, K.; Takahashi, T. Multiple Mechanistic Pathways for Zirconium-Catalyzed Carboalumination of Alkynes. Requirements for Cyclic Carbometalation Processes Involving C-H Activation. J. Am. Chem. Soc. 1996, 118, 9577–9588. [Google Scholar] [CrossRef]
  38. Khalilov, L.M.; Parfenova, L.V.; Rusakov, S.V.; Ibragimov, A.G.; Dzhemilev, U.M. Synthesis and conversions of metallocycles. 22. NMR studies of the mechanism of Cp2ZrCl2-catalyzed cycloalumination of olefins with triethylaluminium to form aluminacyclopentanes. Russ. Chem. Bull. 2000, 49, 2051–2058. [Google Scholar] [CrossRef]
  39. Severn, J.R.; Chadwick, J.C.; Duchateau, R.; Friederichs, N. “Bound but Not Gagged” Immobilizing Single-Site α-Olefin Polymerization Catalysts. Chem. Rev. 2005, 105, 4073–4147. [Google Scholar] [CrossRef]
  40. Hlatky, G.G. Heterogeneous Single-Site Catalysts for Olefin Polymerization. Chem. Rev. 2000, 100, 1347–1376. [Google Scholar] [CrossRef]
  41. Copéret, C.; Allouche, F.; Chan, K.W.; Conley, M.P.; Delley, M.F.; Fedorov, A.; Moroz, I.B.; Mougel, V.; Pucino, M.; Searles, K.; et al. Bridging the Gap between Industrial and Well-Defined Supported Catalysts. Angew. Chem. Int. Ed. 2018, 57, 6398–6440. [Google Scholar] [CrossRef] [PubMed]
  42. Copéret, C.; Comas-Vives, A.; Conley, M.P.; Estes, D.P.; Fedorov, A.; Mougel, V.; Nagae, H.; Núñez-Zarur, F.; Zhizhko, P.A. Surface Organometallic and Coordination Chemistry toward Single-Site Heterogeneous Catalysts: Strategies, Methods, Structures, and Activities. Chem. Rev. 2016, 116, 323–421. [Google Scholar] [CrossRef]
  43. Korzyński, M.D.; Copéret, C. Single sites in heterogeneous catalysts: Separating myth from reality. Trends Chem. 2021, 3, 850–862. [Google Scholar] [CrossRef]
  44. Fink, G.; Steinmetz, B.; Zechlin, J.; Przybyla, C.; Tesche, B. Propene Polymerization with Silica-Supported Metallocene/MAO Catalysts. Chem. Rev. 2000, 100, 1377–1390. [Google Scholar] [CrossRef] [PubMed]
  45. Klapper, M.; Joe, D.; Nietzel, S.; Krumpfer, J.W.; Müllen, K. Olefin Polymerization with Supported Catalysts as an Exercise in Nanotechnology. Chem. Mater. 2014, 26, 802–819. [Google Scholar] [CrossRef]
  46. Stalzer, M.M.; Delferro, M.; Marks, T.J. Supported Single-Site Organometallic Catalysts for the Synthesis of High-Performance Polyolefins. Catal. Lett. 2015, 145, 3–14. [Google Scholar] [CrossRef] [Green Version]
  47. Werny, M.J.; Müller, D.; Hendriksen, C.; Chan, R.; Friederichs, N.H.; Fella, C.; Meirer, F.; Weckhuysen, B.M. Elucidating the Sectioning Fragmentation Mechanism in Silica-Supported Olefin Polymerization Catalysts with Laboratory-Based X-Ray and Electron Microscopy. ChemCatChem 2022, 14, e202200067. [Google Scholar] [CrossRef]
  48. Bashir, M.A.; Monteil, V.; Boisson, C.; McKenna, T.F.L. Avoiding leaching of silica supported metallocenes in slurry phase ethylene homopolymerization. React. Chem. Eng. 2017, 2, 521–530. [Google Scholar] [CrossRef]
  49. Bashir, M.A.; Vancompernolle, T.; Gauvin, R.M.; Delevoye, L.; Merle, N.; Monteil, V.; Taoufik, M.; McKenna, T.F.L.; Boisson, C. Silica/MAO/(n-BuCp)2ZrCl2 catalyst: Effect of support dehydroxylation temperature on the grafting of MAO and ethylene polymerization. Catal. Sci. Technol. 2016, 6, 2962–2974. [Google Scholar] [CrossRef]
  50. Velthoen, M.E.Z.; Boereboom, J.M.; Bulo, R.E.; Weckhuysen, B.M. Insights into the activation of silica-supported metallocene olefin polymerization catalysts by methylaluminoxane. Catal. Today 2019, 334, 223–230. [Google Scholar] [CrossRef]
  51. Jeong, S.M.; Park, J.Y.; Hyun, Y.B.; Baek, J.W.; Kim, H.; Yoon, Y.; Chung, S.; Lee, J.; Lee, B.Y. Syntheses of Silylene-Bridged Thiophene-Fused Cyclopentadienyl ansa-Metallocene Complexes for Preparing High-Performance Supported Catalyst. Catalysts 2022, 12, 283. [Google Scholar] [CrossRef]
  52. Charoenchaidet, S.; Chavadej, S.; Gulari, E. Borane-functionalized silica supports: In situ activated heterogeneous zirconocene catalysts for MAO-free ethylene polymerization. J. Mol. Catal. A Chem. 2002, 185, 167–177. [Google Scholar] [CrossRef]
  53. Lamb, J.V.; Buffet, J.-C.; Turner, Z.R.; O’Hare, D. Ethylene Polymerization Using Zirconocenes Supported on Pentafluorophenyl-Modified Solid Polymethylaluminoxane. Macromolecules 2020, 53, 929–935. [Google Scholar] [CrossRef]
  54. Sauter, D.W.; Popoff, N.; Bashir, M.A.; Szeto, K.C.; Gauvin, R.M.; Delevoye, L.; Taoufik, M.; Boisson, C. The design of a bipodal bis(pentafluorophenoxy)aluminate supported on silica as an activator for ethylene polymerization using surface organometallic chemistry. Chem. Commun. 2016, 52, 4776–4779. [Google Scholar] [CrossRef] [PubMed]
  55. Iiskola, E.I.; Timonen, S.; Pakkanen, T.T.; Harkki, O.; Lehmus, P.; Seppala, J.V. Cyclopentadienyl Surface as a Support for Zirconium Polyethylene Catalysts. Macromolecules 1997, 30, 2853–2859. [Google Scholar] [CrossRef]
  56. Leadbeater, N.E.; Marco, M. Preparation of Polymer-Supported Ligands and Metal Complexes for Use in Catalysis. Chem. Rev. 2002, 102, 3217–3274. [Google Scholar] [CrossRef]
  57. Zhang, J.; Jin, C.-X. Self-immobilized and polymerized miscellaneous metallocene catalysts for ethylene polymerization. Inorg. Chem. Commun. 2006, 9, 683–686. [Google Scholar] [CrossRef]
  58. McKittrick, M.W.; Jones, C.W. Toward Single-Site, Immobilized Molecular Catalysts: Site-Isolated Ti Ethylene Polymerization Catalysts Supported on Porous Silica. J. Am. Chem. Soc. 2004, 126, 3052–3053. [Google Scholar] [CrossRef]
  59. Schneider, H.; Puchta, G.T.; Kaul, F.A.R.; Raudaschl-Sieber, G.; Lefebvre, F.; Saggio, G.; Mihalios, D.; Herrmann, W.A.; Basset, J.M. Immobilization of η5-cyclopentadienyltris(dimethylamido)zirconium polymerization catalysts on a chlorosilane- and HMDS-modified mesoporous silica surface: A new concept for supporting metallocene amides towards heterogeneous single-site-catalysts. J. Mol. Catal. A Chem. 2001, 170, 127–141. [Google Scholar] [CrossRef]
  60. Witzke, R.J.; Chapovetsky, A.; Conley, M.P.; Kaphan, D.M.; Delferro, M. Nontraditional Catalyst Supports in Surface Organometallic Chemistry. ACS Catal. 2020, 10, 11822–11840. [Google Scholar] [CrossRef]
  61. Ballard, D.G.H. Transition metal alkyl compounds as polymerization catalysts. J. Polym. Sci. Polym. Chem. Ed. 1975, 13, 2191–2212. [Google Scholar] [CrossRef]
  62. Dudchenko, V.K.; Zakharov, V.A.; Maksimov, N.G.; Yermakov, Y.I. Comparison of properties of supported organozirconium catalysts containing zirconium ions in different valence states. React. Kinet. Catal. Lett. 1977, 7, 419–424. [Google Scholar] [CrossRef]
  63. Zakharov, V.A.; Yermakov, Y.I. Supported Organometallic Catalysts for Olefin Polymerization. Catal. Rev. Sci. Eng. 1979, 19, 67–103. [Google Scholar] [CrossRef]
  64. Martineau, D.; Dumas, P.; Sigwalt, P. Role of cocatalysts in polymerizations initiated with benzyl derivatives of transition metals, 1. Polymerization of ethylene. Makromol. Chem. 1983, 184, 1389–1395. [Google Scholar] [CrossRef]
  65. Ittel, S.D. Elastomeric Polypropylene Using Alumina-Supported Bis(Arene)Titanium and Tetra(Neophyl)Zirconium Catalysts. J. Macromol. Sci. A Chem. 1990, 27, 1133–1146. [Google Scholar] [CrossRef]
  66. Ballard, D.G.H.; Jones, E.; Padget, J.C.; Pioli, A.J.P.; Robinson, P.A.; Walker, J.; Wyatt, R.J. Polymerization Process. U.S. Patent 4,056,669, 1 November 1977. [Google Scholar]
  67. Popoff, N.; Espinas, J.; Pelletier, J.; Macqueron, B.; Szeto, K.C.; Boyron, O.; Boisson, C.; Del Rosal, I.; Maron, L.; De Mallmann, A.; et al. Small Changes Have Consequences: Lessons from Tetrabenzyltitanium and -zirconium Surface Organometallic Chemistry. Chem. Eur. J. 2013, 19, 964–973. [Google Scholar] [CrossRef]
  68. Tullock, C.W.; Mulhaupt, R.; Ittel, S.D. Elastomeric polypropylene (ELPP) from alumina-supported bis(mesitylene)titanium catalysts. Makromol. Chem. Rapid Commun. 1989, 10, 19–23. [Google Scholar] [CrossRef]
  69. Tullock, C.W.; Tebbe, F.N.; Mulhaupt, R.; Ovenall, D.W.; Setterquist, R.A.; Ittel, S.D. Polyethylene and elastomeric polypropylene using alumina-supported bis(arene) titanium, zirconium, and hafnium catalysts. J. Polym. Sci. A Polym. Chem. 1989, 27, 3063–3081. [Google Scholar] [CrossRef]
  70. Toscano, P.J.; Marks, T.J. Supported organoactinides. High-resolution solid-state carbon-13 NMR studies of catalytically active, alumina-bound pentamethylcyclopentadienyl)thorium methyl and hydride complexes. J. Am. Chem. Soc. 1985, 107, 653–659. [Google Scholar] [CrossRef]
  71. Dahmen, K.H.; Hedden, D.; Burwell, R.L., Jr.; Marks, T.J. Organometallic molecule-support interactions. Highly active organozirconium hydrogenation catalysts and the formation of cationic species on alumina surfaces. Langmuir 1988, 4, 1212–1214. [Google Scholar] [CrossRef]
  72. Marks, T.J. Surface-bound metal hydrocarbyls. Organometallic connections between heterogeneous and homogeneous catalysis. Acc. Chem. Res. 1992, 25, 57–65. [Google Scholar] [CrossRef]
  73. Kaminaka, M.; Soga, K. Polymerization of propene with the catalyst systems composed of Al2O3- or MgCl2-supported Et[IndH4]2ZrCl2 and AlR3 (R = CH3, C2H5). Makromol. Chem. Rapid Comm. 1991, 12, 367–372. [Google Scholar] [CrossRef]
  74. Kaminaka, M.; Soga, K. Polymerization of propene with catalyst systems composed of Al2O3 or MgCl2-supported zirconocene and Al(CH3)3. Polymer 1992, 33, 1105–1107. [Google Scholar] [CrossRef]
  75. Soga, K.; Kaminaka, M. Polymerization of propene with zirconocene-containing supported catalysts activated by common trialkylaluminiums. Makromol. Chem. 1993, 194, 1745–1755. [Google Scholar] [CrossRef]
  76. Kaji, E.; Uozumi, T.; Jin, J.; Sano, T.; Soga, K. Ethene–propene copolymerization with the MgCl2-supported dichlorobis(4,4,4-trifluoro-1-phenyl-1,3-butanedionato)titanium catalyst activated by an ordinary trialkylaluminum. J. Polym. Sci. A Polym. Chem. 1998, 36, 2735–2740. [Google Scholar] [CrossRef]
  77. Nakayama, Y.; Bando, H.; Kaneko, H.; Sonobe, Y.; Mitani, M.; Kojoh, S.; Kashiwa, N.; Fujita, T. MAO-Free new olefin polymerization catalyst systems based on FI catalysts. Stud. Surf. Sci. Catal. 2003, 145, 517–518. [Google Scholar] [CrossRef]
  78. Nakayama, Y.; Bando, H.; Sonobe, Y.; Kaneko, H.; Kashiwa, N.; Fujita, T. New olefin polymerization catalyst systems comprised of bis(phenoxy-imine) titanium complexes and MgCl2-based activators. J. Catal. 2003, 215, 171–175. [Google Scholar] [CrossRef]
  79. Nakayama, Y.; Bando, H.; Sonobe, Y.; Fujita, T. Development of Single-Site New Olefin Polymerization Catalyst Systems Using MgCl2-Based Activators: MAO-Free MgCl2-Supported FI Catalyst Systems. Bull. Chem. Soc. Jpn. 2004, 77, 617–625. [Google Scholar] [CrossRef]
  80. Severn, J.R.; Chadwick, J.C. MAO-Free Activation of Metallocenes and other Single-Site Catalysts for Ethylene Polymerization using Spherical Supports based on MgCl2. Macromol. Rapid. Commun. 2004, 25, 1024–1028. [Google Scholar] [CrossRef]
  81. Huang, R.; Duchateau, R.; Koning, C.E.; Chadwick, J.C. Zirconocene Immobilization and Activation on MgCl2-Based Supports: Factors Affecting Ethylene Polymerization Activity. Macromolecules 2008, 41, 579–590. [Google Scholar] [CrossRef]
  82. Jezequel, M.; Dufaud, V.; Ruiz-Garcia, M.J.; Carrillo-Hermosilla, F.; Neugebauer, U.; Niccolai, G.P.; Lefebvre, F.; Bayard, F.; Corker, J.; Fiddy, S.; et al. Supported Metallocene Catalysts by Surface Organometallic Chemistry. Synthesis, Characterization, and Reactivity in Ethylene Polymerization of Oxide-Supported Mono- and Biscyclopentadienyl Zirconium Alkyl Complexes: Establishment of Structure/Reactivity Relationships. J. Am. Chem. Soc. 2001, 123, 3520–3540. [Google Scholar] [CrossRef] [PubMed]
  83. Ahn, H.; Marks, T.J. Supported Organometallics. Highly Electrophilic Cationic Metallocene Hydrogenation and Polymerization Catalysts Formed via Protonolytic Chemisorption on Sulfated Zirconia. J. Am. Chem. Soc. 1998, 120, 13533–13534. [Google Scholar] [CrossRef]
  84. Ahn, H.; Nicholas, C.P.; Marks, T.J. Surface Organozirconium Electrophiles Activated by Chemisorption on “Super Acidic” Sulfated Zirconia as Hydrogenation and Polymerization Catalysts. A Synthetic, Structural, and Mechanistic Catalytic Study. Organometallics 2002, 21, 1788–1806. [Google Scholar] [CrossRef]
  85. Nicholas, C.P.; Ahn, H.; Marks, T.J. Synthesis, Spectroscopy, and Catalytic Properties of Cationic Organozirconium Adsorbates on “Super Acidic” Sulfated Alumina. “Single-Site” Heterogeneous Catalysts with Virtually 100 Active Sites. J. Am. Chem. Soc. 2003, 125, 4325–4331. [Google Scholar] [CrossRef] [PubMed]
  86. Nicholas, C.P.; Marks, T.J. Sulfated Tin Oxide Nanoparticles as Supports for Molecule-Based Olefin Polymerization Catalysts. Nano Lett. 2004, 4, 1557–1559. [Google Scholar] [CrossRef]
  87. Nicholas, C.P.; Marks, T.J. Zirconium Hydrocarbyl Chemisorption on Sulfated Metal Oxides: New Supports, Chemisorption Pathways, and Implications for Catalysis. Langmuir 2004, 20, 9456–9462. [Google Scholar] [CrossRef]
  88. Saudemint, T.; Spitz, R.; Broyer, J.-P.; Malinge, J.; Verdel, N. Activator Solid Support for Metallocene Catalysts in the Polymerization of Olefins, a Process for Preparing Such a Support, and the Corresponding Catalytic System and Polymerization Process. U.S. Patent 6,239,059, 29 May 2001. [Google Scholar]
  89. McDaniel, M.P.; Collins, K.S.; Eaton, A.P.; Benham, E.A.; Jensen, M.D.; Martin, J.L.; Hawley, G.R. Organometal Catalyst Compositions. U.S. Patent 6,355,594, 12 March 2002. [Google Scholar]
  90. McDaniel, M.P.; Collins, K.S.; Smith, J.L.; Benham, E.A.; Johnson, M.M.; Eaton, A.P.; Jensen, M.D.; Martin, J.L.; Hawley, G.R. Organometal Catalyst Compositions. U.S. Patent 6,395,666, 28 May 2002. [Google Scholar]
  91. McDaniel, M.P.; Collins, K.S.; Hawley, G.R.; Jensen, M.D.; Benham, E.A.; Eaton, A.P.; Martin, J.I.; Wittner, C.E. Organometal Compound Catalyst. U.S. Patent 6,548,442, 15 April 2003. [Google Scholar]
  92. McDaniel, M.P.; Benham, E.A.; Martin, S.J.; Collins, K.S.; Smith, J.L.; Hawley, G.R.; Wittner, C.E.; Jensen, M.D. Compositions That Can Produce Polymers. U.S. Patent 7,601,665, 8 September 2005. [Google Scholar]
  93. Jensen, M.D.; Hawley, G.R.; McDaniel, M.P.; Crain, T.; Benham, E.A.; Martin, J.L.; Yang, Q. Acidic Activator-Supports and Catalysts for Olefin Polymerization. U.S. Patent 7,294,599, 29 December 2005. [Google Scholar]
  94. McDaniel, M.P.; Yang, O.; Muninger, R.S.; Benham, E.A.; Collins, K.S. Silica-Coated Aluminia Activator-Supports for Metallocene Catalyst Composition. U.S. Patent 7,884,163, 25 March 2010. [Google Scholar]
  95. Yang, Q.; Jensen, M.D.; Thorn, M.G.; Jayratne, K.C.; Crain, T.R. Catalyst for Olefin Polymerization. U.S. Patent 8,058,200, 20 November 2008. [Google Scholar]
  96. McDaniel, M.P.; Kilgore, U.; Yang, Q.; Clear, K.S. Methods for Producing Fluorided-Chlorided Silica-Coated Alumina Activator-Supports and Catalyst Systems Containing the Same. U.S. Patent 9,365,667, 14 June 2016. [Google Scholar]
  97. Small, B.L.; Hope, K.D.; Masino, A.P.; McDaniel, M.P.; Buck, R.M.; Beaulieu, W.B.; Yang, Q.; Baralt, E.J.; Netemeyer, E.J.; Kreischer, B. Oligomerization of Alpha Olefins Using Metallocene-SSA Catalyst Systems and Use of the Resultant Polyalphaolefins to Prepare Lubricant Blends. U.S. Patent 2010317904, 16 December 2010. [Google Scholar]
  98. McDaniel, M.P.; Yang, Q.; Muninger, R.S.; Benham, E.A.; Clear, K.S. Silica-Coated Alumina Activator-Supports for Metallocene Catalyst Compositions. U.S. Patent 10,239,975, 26 March 2019. [Google Scholar]
  99. Clarembeau, M.; Pannier, G.; Paye, S. Activating Supports. U.S. Patent 9,175,106, 3 November 2015. [Google Scholar]
  100. Pannier, G.; Chai, C.K. Polymers. U.S. Patent 9,175,119, 3 November 2015. [Google Scholar]
  101. Muron, M.; Pannier, G.; Whiteoak, C.J.; Spitz, R.; Boisson, C. Activating Supports. U.S. Patent 9,012,359, 21 April 2015. [Google Scholar]
  102. Prades, F.; Boisson, C.; Spitz, R.; Razavi, A. Activating Supports for Metallocene Catalysis. U.S. Patent 7,759,271, 20 July 2010. [Google Scholar]
  103. Prades, F. Metallocene Catalyst Components Supported on Activating Supports. U.S. Patent 8,298,977, 30 October 2012. [Google Scholar]
  104. Pannier, G.; Boisson, C.; Spitz, R. Activating Supports with Controlled Distribution of OH Groups. U.S. Patent 8,524,627, 3 September 2013. [Google Scholar]
  105. Tisse, V.F.; Prades, F.; Briquel, R.; Boisson, C.; McKenna, T.F.L. Role of Silica Properties in the Polymerisation of Ethylene Using Supported Metallocene Catalysts. Macromol. Chem. Phys. 2010, 211, 91–102. [Google Scholar] [CrossRef]
  106. Tisse, V.F.; Boisson, C.; McKenna, T.F.L. Activation and Deactivation of the Polymerization of Ethylene over rac-EtInd2ZrCl2 and (nBuCp)2ZrCl2 on an Activating Silica Support. Macromol. Chem. Phys. 2014, 215, 1358–1369. [Google Scholar] [CrossRef]
  107. Prades, F.; Broyer, J.-P.; Belaid, I.; Boyron, O.; Miserque, O.; Spitz, R.; Boisson, C. Borate and MAO Free Activating Supports for Metallocene Complexes. ACS Catal. 2013, 3, 2288–2293. [Google Scholar] [CrossRef]
  108. Eisen, M.S.; Marks, T.J. Recent developments in the surface and catalytic chemistry of supported organoactinides. J. Mol. Catal. 1994, 86, 23–50. [Google Scholar] [CrossRef]
  109. Copéret, C.; Chabanas, M.; Saint-Arroman, R.P.; Basset, J.-M. Homogeneous and Heterogeneous Catalysis: Bridging the Gap through Surface Organometallic Chemistry. Angew. Chem. Int. Ed. 2003, 42, 156–181. [Google Scholar] [CrossRef] [PubMed]
  110. Rascón, F.; Wischert, R.; Copéret, C. Molecular nature of support effects in single-site heterogeneous catalysts: Silica vs. alumina. Chem. Sci. 2011, 2, 1449–1456. [Google Scholar] [CrossRef]
  111. Wegener, S.L.; Marks, T.J.; Stair, P.C. Design Strategies for the Molecular Level Synthesis of Supported Catalysts. Acc. Chem. Res. 2012, 45, 206–214. [Google Scholar] [CrossRef]
  112. McDaniel, M.P.; Jensen, M.D.; Jayaratne, K.; Collins, K.S.; Benham, E.A.; McDaniel, N.D.; Das, P.K.; Martin, J.L.; Yang, Q.; Thorn, M.G.; et al. Metallocene Activation by Solid Acids. In Tailor-Made Polymers: Via Immobilization of Alpha-Olefin Polymerization Catalysts; Severn, J.R., Chadwick, J.C., Eds.; Chapter 7; Wiley-VCH: Weinheim, Germany, 2008; pp. 171–210. [Google Scholar] [CrossRef]
  113. Chadwick, J.C. Catalysts Supported on Magnesium Chloride. In Tailor-Made Polymers: Via Immobilization of Alpha-Olefin Polymerization Catalysts; Severn, J.R., Chadwick, J.C., Eds.; Chapter 6; Wiley-VCH: Weinheim, Germany, 2008; pp. 151–169. [Google Scholar] [CrossRef]
  114. Hoff, R. Activator Supports for Metallocene and Related Catalysts. In Handbook of Transition Metal Polymerization Catalysts, 2nd ed.; Chapter 3; Wiley-VCH: Weinheim, Germany, 2018; pp. 57–65. [Google Scholar] [CrossRef]
  115. Kumawat, J.; Gupta, V.K.; Vanka, K. The Nature of the Active Site in Ziegler–Natta Olefin Polymerization Catalysis Systems—A Computational Investigation. Eur. J. Inorg. Chem. 2014, 2014, 5063–5076. [Google Scholar] [CrossRef]
  116. Credendino, R.; Liguori, D.; Fan, Z.; Morini, G.; Cavallo, L. Toward a Unified Model Explaining Heterogeneous Ziegler–Natta Catalysis. ACS Catal. 2015, 5, 5431–5435. [Google Scholar] [CrossRef] [Green Version]
  117. Potapov, A.G. Titanium-magnesium Ziegler-Natta catalysts: New insight on the active sites precursor. Mol. Catal. 2017, 432, 155–159. [Google Scholar] [CrossRef]
  118. Klaue, A.; Kruck, M.; Friederichs, N.; Bertola, F.; Wu, H.; Morbidelli, M. Insight into the Synthesis Process of an Industrial Ziegler–Natta Catalyst. Ind. Eng. Chem. Res. 2019, 58, 886–896. [Google Scholar] [CrossRef] [Green Version]
  119. Zorve, P.; Linnolahti, M. Adsorption of Titanium Tetrachloride on Magnesium Dichloride Clusters. ACS Omega 2018, 3, 9921–9928. [Google Scholar] [CrossRef]
  120. Vittoria, A.; Meppelder, A.; Friederichs, N.; Busico, V.; Cipullo, R. Demystifying Ziegler–Natta Catalysts: The Origin of Stereoselectivity. ACS Catal. 2017, 7, 4509–4518. [Google Scholar] [CrossRef]
  121. Correa, A.; Credendino, R.; Pater, J.T.M.; Morini, G.; Cavallo, L. Theoretical Investigation of Active Sites at the Corners of MgCl2 Crystallites in Supported Ziegler–Natta Catalysts. Macromolecules 2012, 45, 3695–3701. [Google Scholar] [CrossRef]
  122. Potapov, A.G.; Terskikh, V.V.; Bukatov, G.D.; Zakharov, V.A. 27Al NMR MAS study of AlEt3-nClnMgCl2 systems. J. Mol. Catal. A Chem. 1997, 122, 61–65. [Google Scholar] [CrossRef]
  123. Linnolahti, M.; Pakkanen, T.A.; Bazhenova, A.S.; Denifl, P.; Leinonen, T.; Pakkanen, A. Alkylation of titanium tetrachloride on magnesium dichloride in the presence of Lewis bases. J. Catal. 2017, 353, 89–98. [Google Scholar] [CrossRef]
  124. Yu, Y.; McKenna, T.F.L.; Boisson, C.; Lacerda Miranda, M.S.; Martins, O., Jr. Engineering Poly(ethylene-co-1-butene) through Modulating the Active Species by Alkylaluminum. ACS Catal. 2020, 10, 7216–7229. [Google Scholar] [CrossRef]
  125. Ochędzan-Siodłak, W.; Nowakowska, M. Magnesium chloride modified with organoaluminium compounds as a support of the zirconocene catalyst for ethylene polymerisation. Eur. Polym. J. 2004, 40, 839–846. [Google Scholar] [CrossRef]
  126. Soga, K.; Uozumi, T.; Saito, M.; Shiono, T. Structure of polypropene and poly(ethylene-co-propene) produced with an alumina-supported CpTiCl3/common alkylaluminium catalyst system. Macromol. Chem. Phys. 1994, 195, 1503–1515. [Google Scholar] [CrossRef]
  127. Kang, K.K.; Oh, J.K.; Jeong, Y.-T.; Shiono, T.; Ikeda, T. Highly active MgCl2-supported CpMCl3 (M = Ti, Zr) catalysts for ethylene polymerization. Macromol. Rapid Commun. 1999, 20, 308–311. [Google Scholar] [CrossRef]
  128. Goto, K.; Taniike, T.; Terano, M. Dual-Active-Site Nature of Magnesium Dichloride—Supported Cyclopentadienyl Titanium Chloride Catalysts Switched by an Activator in Propylene Polymerization. Macromol. Chem. Phys. 2013, 214, 1011–1018. [Google Scholar] [CrossRef]
  129. Soga, K.; Kaji, E.; Uozumi, T. Polymerization of propene with the MgCl2-supported dichlorobis (β-diketonato) titanium catalysts combined with methylaluminoxane or an ordinary alkylaluminum. J. Polym. Sci. A Polym. Chem. 1997, 35, 823–826. [Google Scholar] [CrossRef]
  130. Matsui, S.; Mitani, M.; Saito, J.; Tohi, Y.; Makio, H.; Matsukawa, N.; Takagi, Y.; Tsuru, K.; Nitabaru, M.; Nakano, T.; et al. A Family of Zirconium Complexes Having Two Phenoxy-Imine Chelate Ligands for Olefin Polymerization. J. Am. Chem. Soc. 2001, 123, 6847–6856. [Google Scholar] [CrossRef]
  131. Mitani, M.; Mohri, J.; Yoshida, Y.; Saito, J.; Ishii, S.; Tsuru, K.; Matsui, S.; Furuyama, R.; Nakano, T.; Tanaka, H.; et al. Living Polymerization of Ethylene Catalyzed by Titanium Complexes Having Fluorine-Containing Phenoxy-Imine Chelate Ligands. J. Am. Chem. Soc. 2002, 124, 3327–3336. [Google Scholar] [CrossRef]
  132. Mitani, M.; Furuyama, R.; Mohri, J.; Saito, J.; Ishii, S.; Terao, H.; Kashiwa, N.; Fujita, T. Fluorine- and Trimethylsilyl-Containing Phenoxy-Imine Ti Complex for Highly Syndiotactic Living Polypropylenes with Extremely High Melting Temperatures. J. Am. Chem. Soc. 2002, 124, 7888–7889. [Google Scholar] [CrossRef] [PubMed]
  133. Mitani, M.; Furuyama, R.; Mohri, J.; Saito, J.; Ishii, S.; Terao, H.; Nakano, T.; Tanaka, H.; Fujita, T. Syndiospecific Living Propylene Polymerization Catalyzed by Titanium Complexes Having Fluorine-Containing Phenoxy-Imine Chelate Ligands. J. Am. Chem. Soc. 2003, 125, 4293–4305. [Google Scholar] [CrossRef] [PubMed]
  134. Makio, H.; Kashiwa, N.; Fujita, T. FI Catalysts: A New Family of High Performance Catalysts for Olefin Polymerization. Adv. Synth. Catal. 2002, 344, 477–493. [Google Scholar] [CrossRef]
  135. Makio, H.; Terao, H.; Iwashita, A.; Fujita, T. FI catalysts for olefin polymerization—A comprehensive treatment. Chem. Rev. 2011, 111, 2363–2449. [Google Scholar] [CrossRef] [PubMed]
  136. Satyanarayana, G.; Sivaram, S. An unusually stable supported bis(cyclopentadienyl)titanium dichloride-trialkylaluminum catalyst system for ethylene polymerization. Macromolecules 1993, 26, 4712–4714. [Google Scholar] [CrossRef]
  137. Nowakowska, M.; Ochędzan-Siodłak, W.; Kordowska, M. Metallocene catalysts on complex magnesium support modified by organoaluminium compounds. Polimery 2000, 45, 323–327. [Google Scholar] [CrossRef]
  138. Cho, H.S.; Lee, W.Y. Synthesis of inorganic MgCl2–alcohol adduct via recrystallization method and its application in supported organometallic catalysts for the polymerization of ethylene with 1-hexene. J. Mol. Catal. A Chem. 2003, 191, 155–165. [Google Scholar] [CrossRef]
  139. Sobota, P.; Utko, J.; Lis, T.; John, Ł.; Petrus, R.; Drąg-Jarząbek, A. Unexpected Reactions between Ziegler–Natta Catalyst Components and Structural Characterization of Resulting Intermediates. Inorg. Chem. 2016, 55, 4636–4642. [Google Scholar] [CrossRef]
  140. Nakayama, Y.; Saito, J.; Bando, H.; Fujita, T. MgCl2/RnAl(OR)3−n: An Excellent Activator/Support for Transition-Metal Complexes for Olefin Polymerization. Chem. Eur. J. 2006, 12, 7546–7556. [Google Scholar] [CrossRef]
  141. Chung, J.S.; Cho, H.S.; Ko, Y.G.; Lee, W.Y. Preparation of the Ziegler–Natta/metallocene hybrid catalysts on SiO2/MgCl2 bisupport and ethylene polymerization. J. Mol. Catal. A Chem. 1999, 144, 61–69. [Google Scholar] [CrossRef]
  142. Kissin, Y.V.; Nowlin, T.E.; Mink, R.I.; Brandolini, A.J. A New Cocatalyst for Metallocene Complexes in Olefin Polymerization. Macromolecules 2000, 33, 4599–4601. [Google Scholar] [CrossRef]
  143. Carino, A.C.; Park, S.J.; Ko, Y.S. Preparation of (n-BuCp)2ZrCl2 Catalyst Supported on SiO2/MgCl2 Binary Support and its Ethylene-1-hexene Copolymerization. Appl. Chem. Eng. 2018, 29, 461–467. [Google Scholar] [CrossRef]
  144. Scott, S.L.; Church, T.L.; Nguyen, D.H.; Mader, E.A.; Moran, J. An investigation of catalyst/cocatalyst/support interactions in silica-supported olefin polymerization catalysts based on Cp*TiMe3. Top. Catal. 2005, 34, 109–120. [Google Scholar] [CrossRef]
  145. Kumar, D.; Schumacher, K.; du Fresne von Hohenesche, C.; Grün, M.; Unger, K.K. MCM-41, MCM-48 and related mesoporous adsorbents: Their synthesis and characterisation. Coll. Surf. A Physicochem. Eng. Asp. 2001, 187–188, 109–116. [Google Scholar] [CrossRef]
  146. Anwander, R.; Palm, C.; Groeger, O.; Engelhardt, G. Formation of Lewis Acidic Support Materials via Chemisorption of Trimethylaluminum on Mesoporous Silicate MCM-41. Organometallics 1998, 17, 2027–2036. [Google Scholar] [CrossRef]
  147. Sano, T.; Niimi, T.; Miyazaki, T.; Tsubaki, S.; Oumi, Y.; Uozumi, T. Effective activation of metallocene catalyst with AlMCM-41 in propylene polymerization. Catal. Lett. 2001, 71, 105–110. [Google Scholar] [CrossRef]
  148. Li, J.; DiVerdi, J.A.; Maciel, G.E. Chemistry of the Silica Surface:  Liquid−Solid Reactions of Silica Gel with Trimethylaluminum. J. Am. Chem. Soc. 2006, 128, 17093–17101. [Google Scholar] [CrossRef]
  149. Pelletier, J.; Espinas, J.; Vu, N.; Norsic, S.; Baudouin, A.; Delevoye, L.; Trébosc, J.; Le Roux, E.; Santini, C.; Basset, J.-M.; et al. A well-defined silica-supported aluminium alkyl through an unprecedented, consecutive two-step protonolysis–alkyl transfer mechanism. Chem. Commun. 2011, 47, 2979–2981. [Google Scholar] [CrossRef]
  150. Kerber, R.N.; Kermagoret, A.; Callens, E.; Florian, P.; Massiot, D.; Lesage, A.; Copéret, C.; Delbecq, F.; Rozanska, X.; Sautet, P. Nature and Structure of Aluminum Surface Sites Grafted on Silica from a Combination of High-Field Aluminum-27 Solid-State NMR Spectroscopy and First-Principles Calculations. J. Am. Chem. Soc. 2012, 134, 6767–6775. [Google Scholar] [CrossRef]
  151. Kermagoret, A.; Kerber, R.N.; Conley, M.P.; Callens, E.; Florian, P.; Massiot, D.; Copéret, C.; Delbecq, F.; Rozanska, X.; Sautet, P. Triisobutylaluminum: Bulkier and yet more reactive towards silica surfaces than triethyl or trimethylaluminum. Dalton Trans. 2013, 42, 12681–12687. [Google Scholar] [CrossRef]
  152. Werghi, B.; Bendjeriou-Sedjerari, A.; Sofack-Kreutzer, J.; Jedidi, A.; Abou-Hamad, E.; Cavallo, L.; Basset, J. Well-defined silica supported aluminum hydride: Another step towards the utopian single site dream? Chem. Sci. 2015, 6, 5456–5465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Kermagoret, A.; Kerber, R.N.; Conley, M.P.; Callens, E.; Florian, P.; Massiot, D.; Delbecq, F.; Rozanska, X.; Copéret, C.; Sautet, P. Chlorodiethylaluminum supported on silica: A dinuclear aluminum surface species with bridging μ2-Cl-ligand as a highly efficient co-catalyst for the Ni-catalyzed dimerization of ethene. J. Catal. 2014, 313, 46–54. [Google Scholar] [CrossRef]
  154. Trueba, M.; Trasatti, S.P. γ-Alumina as a Support for Catalysts: A Review of Fundamental Aspects. Eur. J. Inorg. Chem. 2005, 2005, 3393–3403. [Google Scholar] [CrossRef]
  155. Motta, A.; Fragala, I.L.; Marks, T.J. Links Between Single-Site Heterogeneous and Homogeneous Catalysis. DFT Analysis of Pathways for Organozirconium Catalyst Chemisorptive Activation and Olefin Polymerization on γ-Alumina. J. Am. Chem. Soc. 2008, 130, 16533–16546. [Google Scholar] [CrossRef]
  156. Stuart, N.M.; Sohlberg, K. The Microstructure of γ-Alumina. Energies 2021, 14, 6472. [Google Scholar] [CrossRef]
  157. Luo, Z. Structure of boehmite-derived γ-alumina and its transformation mechanism revealed by electron crystallography. Acta Cryst. B 2021, 77, 772–784. [Google Scholar] [CrossRef]
  158. Joubert, J.; Delbecq, F.; Thieuleux, C.; Taoufik, M.; Blanc, F.; Copéret, C.; Thivolle-Cazat, J.; Basset, J.-M.; Sautet, P. Synthesis, Characterization, and Catalytic Properties of γ-Al2O3-Supported Zirconium Hydrides through a Combined Use of Surface Organometallic Chemistry and Periodic Calculations. Organometallics 2007, 26, 3329–3335. [Google Scholar] [CrossRef]
  159. Delgado, M.; Delbecq, F.; Santini, C.C.; Lefebvre, F.; Norsic, S.; Putaj, P.; Sautet, P.; Basset, J.-M. Evolution of Structure and of Grafting Properties of γ-Alumina with Pretreatment Temperature. J. Phys. Chem. C 2012, 116, 834–843. [Google Scholar] [CrossRef]
  160. Wischert, R.; Laurent, P.; Copéret, C.; Delbecq, F.; Sautet, P. γ-Alumina: The Essential and Unexpected Role of Water for the Structure, Stability, and Reactivity of “Defect” Sites. J. Am. Chem. Soc. 2012, 134, 14430–14449. [Google Scholar] [CrossRef]
  161. Wischert, R.; Florian, P.; Copéret, C.; Massiot, D.; Sautet, P. Visibility of Al Surface Sites of γ-Alumina: A Combined Computational and Experimental Point of View. J. Phys. Chem. C 2014, 118, 15292–15299. [Google Scholar] [CrossRef]
  162. Joubert, J.; Delbecq, F.; Sautet, P.; Le Roux, E.; Taoufik, M.; Thieuleux, C.; Blanc, F.; Copéret, C.; Thivolle-Cazat, J.; Basset, J.-M. Molecular Understanding of Alumina Supported Single-Site Catalysts by a Combination of Experiment and Theory. J. Am. Chem. Soc. 2006, 128, 9157–9169. [Google Scholar] [CrossRef] [PubMed]
  163. Joubert, J.; Delbecq, F.; Copéret, C.; Basset, J.-M.; Sautet, P. Gamma-alumina: An Active Support to Obtain Immobilized Electron Poor Zr Complexes. Top. Catal. 2008, 48, 114–119. [Google Scholar] [CrossRef]
  164. Mason, A.H.; Motta, A.; Das, A.; Ma, Q.; Bedzyk, M.J.; Kratish, Y.; Marks, T.J. Rapid atom-efficient polyolefin plastics hydrogenolysis mediated by a well-defined single-site electrophilic/cationic organo-zirconium catalyst. Nat. Commun. 2022, 13, 7187. [Google Scholar] [CrossRef] [PubMed]
  165. Delgado, M.; Santini, C.C.; Delbecq, F.; Wischert, R.; Le Guennic, B.; Tosin, G.; Spitz, R.; Basset, J.-M.; Sautet, P. Alumina as a Simultaneous Support and Co Catalyst: Cationic Hafnium Complex Evidenced by Experimental and DFT Analyses. J. Phys. Chem. C 2010, 114, 18516–18528. [Google Scholar] [CrossRef]
  166. Mazoyer, E.; Trébosc, J.; Baudouin, A.; Boyron, O.; Pelletier, J.; Basset, J.-M.; Vitorino, M.J.; Nicholas, C.P.; Gauvin, R.M.; Taoufik, M.; et al. Heteronuclear NMR Correlations To Probe the Local Structure of Catalytically Active Surface Aluminum Hydride Species on γ-Alumina. Angew. Chem. 2010, 49, 9854–9858. [Google Scholar] [CrossRef]
  167. Ikenaga, K.; Chen, S.; Ohshima, M.; Kurokawa, H.; Miura, H. Ethylene polymerization over zirconocenes supported on alumina- and titania-based acidic oxides. Catal. Commun. 2007, 8, 36–38. [Google Scholar] [CrossRef]
  168. Ashuiev, A.; Allouche, F.; Wili, N.; Searles, K.; Klose, D.; Copéret, C.; Jeschke, G. Molecular and supported Ti(III)-alkyls: Efficient ethylene polymerization driven by the π-character of metal–carbon bonds and back donation from a singly occupied molecular orbital. Chem. Sci. 2021, 12, 780–792. [Google Scholar] [CrossRef]
  169. Chizallet, C.; Raybaud, P. Acidity of Amorphous Silica–Alumina: From Coordination Promotion of Lewis Sites to Proton Transfer. ChemPhysChem 2010, 11, 105–108. [Google Scholar] [CrossRef]
  170. Weiss, K.; Wirth-Pfeifer, C.; Hofmann, M.; Botzenhardt, S.; Lang, H.; Bruning, K.; Meichel, E. Polymerisation of ethylene or propylene with heterogeneous metallocene catalysts on clay minerals. J. Mol. Catal. A Chem. 2002, 182–183, 143–149. [Google Scholar] [CrossRef]
  171. Weiss, K.; Botzenhardt, S.; Hofmann, M. Ziegler-Natta and Metallocene Polymerisation of Olefins with Heterogeneous Catalysts. In Metalorganic Catalysts for Synthesis and Polymerization; Kaminsky, W., Ed.; Springer: Berlin/Heidelberg, Germany, 2008; pp. 97–101. [Google Scholar] [CrossRef]
  172. Nakano, H.; Takahashi, T.; Uchino, H.; Tayano, T.; Sugano, T. 4-Polymerization Behavior with Metallocene Catalyst Supported by Clay Mineral Activator. Stud. Surf. Sci. Catal. 2006, 161, 19–24. [Google Scholar] [CrossRef]
  173. Sun, T.; Garcés, J.M. Acidic lamellar aerogel nanoplate activated olefin polymerization with metallocene catalysts. Catal. Commun. 2003, 4, 97–100. [Google Scholar] [CrossRef]
  174. Shih, K.-Y.; Denton, D.A.; Carney, M.J. Metallocene and Constrained Geometry Catalyst Systems Employing Agglomerated Metal Oxide/Clay Support-Activator and Method of Their Preparation. U.S. Patent 6,559,090, 6 May 2003. [Google Scholar]
  175. Tabernero, V.; Camejo, C.; Terreros, P.; Alba, M.D.; Cuenca, T. Silicoaluminates as “Support Activator” Systems in Olefin Polymerization Processes. Materials 2010, 3, 1015–1030. [Google Scholar] [CrossRef] [Green Version]
  176. Yamamoto, K. Olefin polymerizations with zirconocene supported on SiO2 modified by MgO, NaOH and LiOH. Appl. Catal. A Gen. 2009, 368, 65–70. [Google Scholar] [CrossRef]
  177. Higuchi, K.; Onaka, M.; Izumi, Y. Solid Acid and Base-Catalyzed Cyanosilylation of Carbonyl Compounds with Cyanotrimethylsilane. Bull. Chem. Soc. Jpn. 1993, 66, 2016–2032. [Google Scholar] [CrossRef]
  178. Corma, A. Inorganic Solid Acids and Their Use in Acid-Catalyzed Hydrocarbon Reactions. Chem. Rev. 1995, 95, 559–614. [Google Scholar] [CrossRef]
  179. Williams, L.A.; Guo, N.; Motta, A.; Delferro, M.; Fragalà, I.L.; Miller, J.T.; Marks, T.J. Surface structural-chemical characterization of a single-site d0 heterogeneous arene hydrogenation catalyst having 100% active sites. Proc. Natl. Acad. Sci. USA 2012, 110, 413–418. [Google Scholar] [CrossRef]
  180. Gu, W.; Stalzer, M.M.; Nicholas, C.P.; Bhattacharyya, A.; Motta, A.; Gallagher, J.R.; Zhang, G.; Miller, J.T.; Kobayashi, T.; Pruski, M.; et al. Benzene Selectivity in Competitive Arene Hydrogenation: Effects of Single-Site Catalyst···Acidic Oxide Surface Binding Geometry. J. Am. Chem. Soc. 2015, 137, 6770–6780. [Google Scholar] [CrossRef] [Green Version]
  181. Stalzer, M.M.; Nicholas, C.P.; Bhattacharyya, A.; Motta, A.; Delferro, M.; Marks, T.J. Single-Face/All-cis Arene Hydrogenation by a Supported Single-Site d0 Organozirconium Catalyst. Angew. Chem. Int. Ed. 2016, 55, 5263–5267. [Google Scholar] [CrossRef]
  182. Li, X.; Nagaoka, K.; Simon, L.J.; Olindo, R.; Lercher, J.A.; Hofmann, A.; Sauer, J. Oxidative Activation of n-Butane on Sulfated Zirconia. J. Am. Chem. Soc. 2005, 127, 16159–16166. [Google Scholar] [CrossRef]
  183. Liu, J.; Liu, N.; Ren, K.; Shi, L.; Meng, X. Sulfated Zirconia Synthesized in a One Step Solvent-Free Method for Removal of Olefins from Aromatics. Ind. Eng. Chem. Res. 2017, 56, 7693–7699. [Google Scholar] [CrossRef]
  184. Hofmann, A.; Sauer, J. Surface Structure of Hydroxylated and Sulfated Zirconia. A Periodic Density-Functional Study. J. Phys. Chem. B 2004, 108, 14652–14662. [Google Scholar] [CrossRef]
  185. Benaïssa, M.; Santiesteban, J.G.; Díaz, G.; Chang, C.D.; José-Yacamán, M. Interaction of Sulfate Groups with the Surface of Zirconia: An HRTEM Characterization Study. J. Catal. 1996, 161, 694–703. [Google Scholar] [CrossRef]
  186. Zhang, J.; Motta, A.; Gao, Y.; Stalzer, M.M.; Delferro, M.; Liu, B.; Lohr, T.L.; Marks, T.J. Cationic Pyridylamido Adsorbate on Brønsted Acidic Sulfated Zirconia: A Molecular Supported Organohafnium Catalyst for Olefin Homo- and Co-Polymerization. ACS Catal. 2018, 8, 4893–4901. [Google Scholar] [CrossRef]
  187. Domski, G.J.; Lobkovsky, E.B.; Coates, G.W. Polymerization of α-Olefins with Pyridylamidohafnium Catalysts: Living Behavior and Unexpected Isoselectivity from a Cs-Symmetric Catalyst Precursor. Macromolecules 2007, 40, 3510–3513. [Google Scholar] [CrossRef]
  188. Wei, J.; Zhang, W.; Wickham, R.; Sita, L.R. Programmable Modulation of Co-monomer Relative Reactivities for Living Coordination Polymerization through Reversible Chain Transfer between “Tight” and “Loose” Ion Pairs. Angew. Chem. Int. Ed. 2010, 49, 9140–9144. [Google Scholar] [CrossRef]
  189. Froese, R.D.J.; Jazdzewski, B.A.; Klosin, J.; Kuhlman, R.L.; Theriault, C.N.; Welsh, D.M.; Abboud, K.A. Imino-Amido Hf and Zr Complexes: Synthesis, Isomerization, and Olefin Polymerization. Organometallics 2011, 30, 251–262. [Google Scholar] [CrossRef]
  190. Figueroa, R.; Froese, R.D.; He, Y.; Klosin, J.; Theriault, C.N.; Abboud, K.A. Synthesis of Imino-Enamido Hafnium and Zirconium Complexes: A New Family of Olefin Polymerization Catalysts with Ultrahigh-Molecular-Weight Capabilities. Organometallics 2011, 30, 1695–1709. [Google Scholar] [CrossRef]
  191. Klosin, J.; Fontaine, P.P.; Figueroa, R. Development of Group IV Molecular Catalysts for High Temperature Ethylene-α-Olefin Copolymerization Reactions. Acc. Chem. Res. 2015, 48, 2004–2016. [Google Scholar] [CrossRef]
  192. Baier, M.C.; Zuideveld, M.A.; Mecking, S. Post-Metallocenes in the Industrial Production of Polyolefins. Angew. Chem. Int. Ed. 2014, 53, 9722–9744. [Google Scholar] [CrossRef] [Green Version]
  193. Zuccaccia, C.; Macchioni, A.; Busico, V.; Cipullo, R.; Talarico, G.; Alfano, F.; Boone, H.W.; Frazier, K.A.; Hustad, P.D.; Stevens, J.C.; et al. Intra- and Intermolecular NMR Studies on the Activation of Arylcyclometallated Hafnium Pyridyl-Amido Olefin Polymerization Precatalysts. J. Am. Chem. Soc. 2008, 130, 10354–10368. [Google Scholar] [CrossRef]
  194. Zhang, J.; Mason, A.H.; Motta, A.; Cesar, L.G.; Kratish, Y.; Lohr, T.L.; Miller, J.T.; Gao, Y.; Marks, T.J. Surface vs Homogeneous Organo-Hafnium Catalyst Ion-Pairing and Ligand Effects on Ethylene Homo- and Copolymerizations. ACS Catal. 2021, 11, 3239–3250. [Google Scholar] [CrossRef]
  195. Williams, L.A.; Marks, T.J. Synthesis, Characterization, and Heterogeneous Catalytic Implementation of Sulfated Alumina Nanoparticles. Arene Hydrogenation and Olefin Polymerization Properties of Supported Organozirconium Complexes. ACS Catal. 2011, 1, 238–245. [Google Scholar] [CrossRef]
  196. Zhang, J.; Mason, A.H.; Wang, Y.; Motta, A.; Kobayashi, T.; Pruski, M.; Gao, Y.; Marks, T.J. Beyond the Active Site. Cp*ZrMe3/Sulfated Alumina-Catalyzed Olefin Polymerization Tacticity via Catalyst Surface Ion-Pairing. ChemCatChem 2021, 13, 2564–2569. [Google Scholar] [CrossRef]
  197. Williams, L.A.; Marks, T.J. Chemisorption Pathways and Catalytic Olefin Polymerization Properties of Group 4 Mono- and Binuclear Constrained Geometry Complexes on Highly Acidic Sulfated Alumina. Organometallics 2009, 28, 2053–2061. [Google Scholar] [CrossRef]
  198. McInnis, J.P.; Delferro, M.; Marks, T.J. Multinuclear Group 4 Catalysis: Olefin Polymerization Pathways Modified by Strong Metal-Metal Cooperative Effects. Acc. Chem. Res. 2014, 47, 2545–2557. [Google Scholar] [CrossRef]
  199. Delferro, M.; Marks, T.J. Multinuclear Olefin Polymerization Catalysts. Chem. Rev. 2011, 111, 2450–2485. [Google Scholar] [CrossRef]
  200. Liu, S.; Invergo, A.M.; McInns, J.P.; Mouat, A.R.; Motta, A.; Lohr, T.L.; Delferro, M.; Marks, T.J. Distinctive Stereochemically Linked Cooperative Effects in Bimetallic Titanium Olefin Polymerization Catalysts. Organometallics 2017, 36, 4403–4421. [Google Scholar] [CrossRef]
  201. Yang, Q.; Jensen, M.D.; Martin, J.L.; Thorn, M.G.; McDaniel, M.P.; Yu, Y.; Rohlfing, D.C. Polymerization Catalysts for Producing High Molecular Weight Polymers with Low Levels of Long Chain Branching. U.S. Patent 7,517,939, 2 August 2007. [Google Scholar]
  202. Yang, Q.; McDaniel, M.P.; Martin, J.L.; Ding, E.; Rohlfing, D.C.; Crain, T.R. Catalyst Systems and Methods of Making and Using Same. U.S. Patent 8,932,975, 8 March 2012. [Google Scholar]
  203. McDaniel, M.P.; Yang, Q.; Crain, T.R. Processes for Preparing Metallocene-Based Catalyst Systems. U.S. Patent 9,163,098, 20 October 2015. [Google Scholar]
  204. Clark, K.M.; Yang, Q.; Glass, G.L. Processes for Preparing Solid Metallocene-Based Catalyst Systems. U.S. Patent 9,303,106, 5 April 2016. [Google Scholar]
  205. Iacono, S.T.; Jennings, A.R. Recent Studies on Fluorinated Silica Nanometer-Sized Particles. Nanomaterials 2019, 9, 684. [Google Scholar] [CrossRef] [Green Version]
  206. Kolditz, L.; Calov, U.; Kauschka, G.; Schmidt, W. Herstellung von fluorreichen Verbindungen der Reihe C2ClnF6−n in heterogener Katalyse. Z. Anorg. Allg. Chem. 1977, 434, 55–62. [Google Scholar] [CrossRef]
  207. Scokart, P.O.; Selim, S.A.; Damon, J.P.; Rouxhet, P.G. The chemistry and surface chemistry of fluorinated alumina. J. Colloid Interface Sci. 1979, 70, 209–222. [Google Scholar] [CrossRef]
  208. Krahl, T.; Kemnitz, E. Aluminium fluoride—The strongest solid Lewis acid: Structure and reactivity. Catal. Sci. Technol. 2017, 7, 773–796. [Google Scholar] [CrossRef]
  209. Becker, C.; Braun, T.; Paulus, B. Theoretical Study on the Lewis Acidity of the Pristine AlF3 and Cl-Doped α-AlF3 Surfaces. Catalysts 2021, 11, 565. [Google Scholar] [CrossRef]
  210. Kemnitz, E.; Groß, U.; Rüdiger, S.; Shekar, C.S. Amorphous Metal Fluorides with Extraordinary High Surface Areas. Angew. Chem. Int. Ed. 2003, 42, 4251–4254. [Google Scholar] [CrossRef] [PubMed]
  211. Calvo, B.; Marshall, C.P.; Krahl, T.; Kröhnert, J.; Trunschke, A.; Scholz, G.; Braun, T.; Kemnitz, E. Comparative study of the strongest solid Lewis acids known: ACF and HS-AlF3. Dalton Trans. 2018, 47, 16461–16473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Ahrem, L.; Wolf, J.; Scholz, G.; Kemnitz, E. A novel fluoride-doped aluminium oxide catalyst with tunable Brønsted and Lewis acidity. Catal. Sci. Technol. 2018, 8, 1404–1413. [Google Scholar] [CrossRef]
  213. Fischer, L.; Harlé, V.; Kasztelan, S.; d’Espinose de la Caillerie, J.-B. Identification of fluorine sites at the surface of fluorinated γ-alumina by two-dimensional MAS NMR. Solid State Nucl. Magn. Reson. 2000, 16, 85–91. [Google Scholar] [CrossRef] [PubMed]
  214. Panchenko, V.N.; Danilova, G.; Zakharov, V.A.; Semikolenova, N.V.; Paukshtis, E.A. Effect of the acid–base properties of the support on the catalytic activity of ethylene polymerization using supported catalysts composed of Cp2ZrX2 (X = Cl, Me) and Al2O3. Reac. Kinet. Mech. Catal. 2017, 122, 275–287. [Google Scholar] [CrossRef]
  215. Corma, A.; Fornés, V.; Ortega, E. The nature of acid sites on fluorinated γ-Al2O3. J. Catal. 1985, 92, 284–290. [Google Scholar] [CrossRef]
  216. McDaniel, M.P.; Klendworth, D.D.; Johnson, M.M. Fluorided Aluminas, Catalysts, and Polymerization Processes. U.S. Patent 5,171,798, 15 December 1992. [Google Scholar]
  217. Prades, F.; Spitz, R.; Boisson, C.; Sirol, S.; Razavi, A. Transition Metal Complexes Supported on Activating Support. U.S. Patent 8,426,539, 23 April 2013. [Google Scholar]
  218. Daniell, W.; Schubert, U.; Glöckler, R.; Meyer, A.; Noweck, K.; Knözinger, H. Enhanced surface acidity in mixed alumina–silicas: A low-temperature FTIR study. Appl. Catal. A Gen. 2000, 196, 247–260. [Google Scholar] [CrossRef]
  219. Ding, E.; Yang, Q.; Guatney, L.W.; Greco, J.F. Ziegler-Natta—Metallocene Dual Catalyst Systems with Activator-Supports. U.S. Patent 9,540,457, 10 January 2017. [Google Scholar]
  220. Ding, E.; Yang, Q.; Guatney, L.W.; Greco, J.F. Heterogeneous Ziegler-Natta Catalysts with Fluorided Silica-Coated Alumina. U.S. Patent 9,758,599, 12 September 2017. [Google Scholar]
  221. Terano, M.; Soga, H.; Kimura, K. Catalyst for Polymerization of Olefins. U.S. Patent 4,829,037, 9 May 1989. [Google Scholar]
  222. 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]
  223. Nifant’ev, I.E.; Vinogradov, A.A.; Vinogradov, A.A.; Bagrov, V.V.; Kiselev, A.V.; Minyaev, M.E.; Samurganova, T.I.; Ivchenko, P.V. Heterocene Catalysts and Reaction Temperature Gradient in Dec-1-ene Oligomerization for the Production of Low Viscosity PAO Base Stocks. Ind. Eng. Chem. Res. 2023, 62, 6347–6353. [Google Scholar] [CrossRef]
  224. Soga, K.; Uozumi, T.; Kishi, N. A metallocene catalyst system for olefin polymerization using a heteropolyacid as the counter-anion. Macromol. Rapid Commun. 1995, 16, 793–798. [Google Scholar] [CrossRef]
  225. McDaniel, M.P.; Collins, K.S.; Johnson, M.M.; Smith, J.L.; Benham, E.A.; Hawley, G.R.; Wittner, C.E.; Jensen, M.D. Compositions That Can Produce Polymers. U.S. Patent 6,107,230, 22 August 2000. [Google Scholar]
  226. McDaniel, M.P.; Kimble, J.B.; Collins, K.S.; Benham, E.A.; Jensen, M.D.; Hawley, G.R.; Martin, J.L. Organometal Catalyst Compositions. U.S. Patent 6,376,415, 23 April 2002. [Google Scholar]
  227. McDaniel, M.P.; Collins, K.S.; Eaton, A.P.; Benham, E.A.; Jensen, M.D.; Martin, J.L.; Hawley, G.R. Organometal Catalyst Compositions. U.S. Patent 6,548,441, 15 April 2003. [Google Scholar]
  228. Kurokawa, H.; Morita, S.; Matsuda, M.; Suzuki, H.; Ohshima, M.; Miura, H. Polymerization of ethylene using zirconocenes supported on swellable cation-exchanged fluorotetrasilicic mica. Appl. Catal. A Gen. 2009, 360, 192–198. [Google Scholar] [CrossRef]
  229. Ward, D.G. Halogenated Supports and Supported Activators. U.S. Patent 5,885,924, 23 March 1999. [Google Scholar]
  230. Bando, H.; Nakayama, Y.; Sonobe, Y.; Fujita, T. Bis(Phenoxy-Imine) Zr Complexes/Et3Al/Heteropoly Compound Catalyst Systems for Ethylene Polymerization. Macromol. Rapid. Commun. 2003, 24, 732–736. [Google Scholar] [CrossRef]
  231. Hicks, J.C.; Mullis, B.A.; Jones, C.W. Sulfonic Acid Functionalized SBA-15 Silica as a Methylaluminoxane-Free Cocatalyst/Support for Ethylene Polymerization. J. Am. Chem. Soc. 2007, 129, 8426–8427. [Google Scholar] [CrossRef]
  232. Alvaro, M.; Corma, A.; Das, D.; Fornésa, V.; García, H. Single-step preparation and catalytic activity of mesoporous MCM-41 and SBA-15 silicas functionalized with perfluoroalkylsulfonic acid groups analogous to Nafion. Chem. Commun. 2004, 8, 956–957. [Google Scholar] [CrossRef]
  233. Culver, D.B.; Venkatesh, A.; Huynh, W.; Rossini, A.J.; Conley, M.P. Al(ORF)3 (RF = C(CF3)3) activated silica: A well-defined weakly coordinating surface anion. Chem. Sci. 2020, 11, 1510–1517. [Google Scholar] [CrossRef] [Green Version]
  234. Samudrala, K.K.; Huynh, W.; Dorn, R.W.; Rossini, A.J.; Conley, M.P. Formation of a Strong Heterogeneous Aluminum Lewis Acid on Silica. Angew. Chem. Int. Ed. 2022, 61, e202205745. [Google Scholar] [CrossRef]
  235. Kusoglu, A.; Weber, A.Z. New Insights into Perfluorinated Sulfonic-Acid Ionomers. Chem. Rev. 2017, 117, 987–1104. [Google Scholar] [CrossRef]
  236. Tasgiro, T.; Ueda, T. Catalyst Olefin Polymerization and Process for Olefin Polymerization Using the Same. U.S. Patent 5,703,181, 30 December 1997. [Google Scholar]
  237. Pârvulescu, V.I.; Hardacre, C. Catalysis in Ionic Liquids. Chem. Rev. 2007, 107, 2615–2665. [Google Scholar] [CrossRef]
  238. Hogg, J.M.; Ferrer-Ugalde, A.; Coleman, F.; Swadźba-Kwaśny, M. Borenium Ionic Liquids as Alternative to BF3 in Polyalphaolefins (PAOs) Synthesis. ACS Sustain. Chem. Eng. 2019, 7, 15044–15052. [Google Scholar] [CrossRef]
  239. Fehér, C.; Tomasek, S.; Hancsók, J.; Skoda-Földes, R. Oligomerization of light olefins in the presence of a supported Brønsted acidic ionic liquid catalyst. Appl. Catal. B Environ. 2018, 239, 52–60. [Google Scholar] [CrossRef]
  240. Mashayekhi, M.; Talebi, S.; Sadjadi, S.; Bahri-Laleh, N. Production of polyalfaolefin-based lubricants using new (poly)ionic liquid/AlCl3 catalysts as environmentally friendly alternatives to commercial AlCl3 route. Appl. Catal. A Gen. 2021, 623, 118274. [Google Scholar] [CrossRef]
  241. Tabrizi, M.; Bahri-Laleh, N.; Sadjadi, S.; Nekoomanesh-Haghighi, M. The effect of ionic liquid containing AlCl3 catalytic systems on the microstructure and properties of polyalphaolefin based lubricants. J. Mol. Liq. 2021, 335, 116299. [Google Scholar] [CrossRef]
  242. Karimi, S.; Sadjadi, S.; Bahri-Laleh, N.; Nekoomanesh-Haghighi, M. New less-toxic halloysite-supported ionic liquid/AlCl3 oligomerization catalysts: A comparative study on the effects of various ionic liquids on the properties of polyalphaolefins. Mol. Catal. 2021, 509, 111648. [Google Scholar] [CrossRef]
  243. Carlin, R.T.; Wilkes, J.S. Complexation of Cp2MCl2 in a chloroaluminate molten salt: Relevance to homogeneous Ziegler-Natta catalysis. J. Mol. Catal. 1990, 63, 125–129. [Google Scholar] [CrossRef]
  244. Deeth, R.J.; Jenkins, H.D.B. A Density Functional and Thermochemical Study of M−X Bond Lengths and Energies in [MX6]2− Complexes:  LDA versus Becke88/Perdew86 Gradient-Corrected Functionals. J. Phys. Chem. A 1997, 101, 4793–4798. [Google Scholar] [CrossRef]
  245. Ochędzan-Siodłak, W.; Sacher-Majewska, B. Biphasic ethylene polymerisation using ionic liquid over a titanocene catalyst activated by an alkyl aluminium compound. Eur. Polym. J. 2007, 43, 3688–3694. [Google Scholar] [CrossRef]
  246. Ochędzan-Siodłak, W.; Pawelska, P. Chloroaluminate ionic liquids as a medium of titanocene catalyst activated by alkyl aluminum compounds for ethylene polymerization. Polymery 2008, 53, 371–376. [Google Scholar] [CrossRef]
  247. Ochędzan-Siodłak, W.; Dziubek, K.; Czaja, K. Comparison of imidazolium and pyridinium ionic liquids as the media for biphasic ethylene polymerization in the presence of titanocene catalyst. Polymery 2009, 54, 501–506. [Google Scholar] [CrossRef] [Green Version]
  248. Ochędzan-Siodłak, W.; Dziubek, K.; Czaja, K. Effect of immobilization of titanocene catalyst in aralkyl imidazolium chloroaluminate media on performance of biphasic ethylene polymerization and polyethylene properties. Polym. Bull. 2013, 70, 1–21. [Google Scholar] [CrossRef] [Green Version]
  249. Ochędzan-Siodłak, W.; Dziubek, K. Metallocenes and post-metallocenes immobilized on ionic liquid-modified silica as catalysts for polymerization of ethylene. Appl. Catal. A Gen. 2014, 484, 134–141. [Google Scholar] [CrossRef]
  250. Nifant’ev, I.E.; Salakhov, I.I.; Ivchenko, P.V. Transition Metal–(μ-Cl)–Aluminum Bonding in α-Olefin and Diene Chemistry. Molecules 2022, 27, 7164. [Google Scholar] [CrossRef] [PubMed]
  251. Jalali, A.; Nekoomanehs-Haghighi, M.; Dehghani, S.; Bahri-Laleh, N. Effect of metal type on the metallocene-catalyzed oligomerization of 1-hexene and 1-octene to produce polyα-olefin-based synthetic lubricants. Appl. Organomet. Chem. 2020, 34, e5338. [Google Scholar] [CrossRef]
  252. Hanifpour, A.; Bahri-Laleh, N.; Nekoomanesh-Haghighi, M.; Poater, A. 1-Decene oligomerization by new complexes bearing diamine-diphenolates ligands: Effect of ligand structure. Appl. Organomet. Chem. 2021, 35, e6227. [Google Scholar] [CrossRef]
  253. Hanifpour, A.; Gargari, M.H.; Darounkola, M.R.R.; Kalantari, Z.; Bahri-Laleh, N. Kinetic and microstructural studies of Cp2ZrCl2 and Cp2HfCl2-catalyzed oligomerization of higher α-olefins in mPAO oil base stocks production. Polyolefins J. 2021, 8, 31–40. [Google Scholar] [CrossRef]
  254. Nifant’ev, I.E.; Vinogradov, A.A.; Vinogradov, A.A.; Minyaev, M.E.; Bagrov, V.V.; Salakhov, I.I.; Shaidullin, N.M.; Chalykh, A.E.; Shapagin, A.V.; Ivchenko, P.V. Heterocene-catalyzed ethylene/oct-1-ene copolymerization under MAO-free and low-MAO conditions: The synthesis of highly statistical copolymers and their use in blending with HDPE. Polymer 2023, 272, 125836. [Google Scholar] [CrossRef]
Scheme 1. (a) Polymerization of α-olefins on Group 4 metal single-site catalysts; (b) activation of pre-catalysts; (c) neutral and cationic species formed during two-stage activation of zirconocenes by R3Al and perfluoroaryl borates. L2—bis(η5-ligand), ≠—transition state.
Scheme 1. (a) Polymerization of α-olefins on Group 4 metal single-site catalysts; (b) activation of pre-catalysts; (c) neutral and cationic species formed during two-stage activation of zirconocenes by R3Al and perfluoroaryl borates. L2—bis(η5-ligand), ≠—transition state.
Polymers 15 03095 sch001
Scheme 2. Heterogenization of Group 4 single-site catalysts: (a) using silica/MAO (or perfluoroaryl borate) supports; (b) via immobilization of the ligand precursors; and (c) using activating supports.
Scheme 2. Heterogenization of Group 4 single-site catalysts: (a) using silica/MAO (or perfluoroaryl borate) supports; (b) via immobilization of the ligand precursors; and (c) using activating supports.
Polymers 15 03095 sch002
Scheme 3. Early studies: activation of Group 4 metal alkyls (a) and bis(arene) complexes (b) by silica and alumina.
Scheme 3. Early studies: activation of Group 4 metal alkyls (a) and bis(arene) complexes (b) by silica and alumina.
Polymers 15 03095 sch003
Figure 1. Schematic representation of a MgCl2 monolayer fragment developed by Cavallo et al. The (104) and (110) lateral faces are indicated. The Mg atoms are colored orange. The Cl atoms above and below the Mg plane are dark and light green colored, respectively. Reprinted with permission from [116]. Copyright (2012) American Chemical Society.
Figure 1. Schematic representation of a MgCl2 monolayer fragment developed by Cavallo et al. The (104) and (110) lateral faces are indicated. The Mg atoms are colored orange. The Cl atoms above and below the Mg plane are dark and light green colored, respectively. Reprinted with permission from [116]. Copyright (2012) American Chemical Society.
Polymers 15 03095 g001
Scheme 4. (a) β-Diketonate Ti(IV) complexes studied in propylene polymerization after activation by R3Al/MgCl2 [129]; (b) FI catalysts studied in ethylene polymerization using Et3Al/MgCl2 activator [79].
Scheme 4. (a) β-Diketonate Ti(IV) complexes studied in propylene polymerization after activation by R3Al/MgCl2 [129]; (b) FI catalysts studied in ethylene polymerization using Et3Al/MgCl2 activator [79].
Polymers 15 03095 sch004
Scheme 5. Proposed mechanism for the impregnation of the metallocene catalysts over the recrystallized MgCl2 [138].
Scheme 5. Proposed mechanism for the impregnation of the metallocene catalysts over the recrystallized MgCl2 [138].
Polymers 15 03095 sch005
Figure 2. Scanning electron micrograph of polyethylene prepared using Cp2TiCl2 (Ti2)/MgCl2 1.1EtOH/Et3Al catalyst. Reprinted with permission from [80]. Copyright (2004) Wiley-VCH Verlag GmbH and Co.
Figure 2. Scanning electron micrograph of polyethylene prepared using Cp2TiCl2 (Ti2)/MgCl2 1.1EtOH/Et3Al catalyst. Reprinted with permission from [80]. Copyright (2004) Wiley-VCH Verlag GmbH and Co.
Polymers 15 03095 g002
Figure 3. Possible modes of coordination of a zirconocene on (a) a (110) and (b) a (100) face of MgCl2. Reprinted with permission from [81]. Copyright (2008) American Chemical Society.
Figure 3. Possible modes of coordination of a zirconocene on (a) a (110) and (b) a (100) face of MgCl2. Reprinted with permission from [81]. Copyright (2008) American Chemical Society.
Polymers 15 03095 g003
Scheme 6. Bis(phenoxy-imine) titanium complexes, studied in ethylene polymerization using MgCl2/iBumAl(OR)n support [77,78,79].
Scheme 6. Bis(phenoxy-imine) titanium complexes, studied in ethylene polymerization using MgCl2/iBumAl(OR)n support [77,78,79].
Polymers 15 03095 sch006
Figure 4. Photographs of the polyethylenes formed with complex Ti7 using (a) MgCl2/iBumAl(OR)n or (b) MAO as an activator. Reprinted with permission from [78]. Copyright (2003) Elsevier B. V.
Figure 4. Photographs of the polyethylenes formed with complex Ti7 using (a) MgCl2/iBumAl(OR)n or (b) MAO as an activator. Reprinted with permission from [78]. Copyright (2003) Elsevier B. V.
Polymers 15 03095 g004
Scheme 7. Bis(phenoxy-imine) titanium (a) and zirconium (b) complexes, studied in polymerization using MgCl2/iBumAl(OR)n supports [79,140].
Scheme 7. Bis(phenoxy-imine) titanium (a) and zirconium (b) complexes, studied in polymerization using MgCl2/iBumAl(OR)n supports [79,140].
Polymers 15 03095 sch007
Figure 5. SEM image of the spherical particle PEs produced with MgCl2-supported Zr-FI catalyst Zr23. Reprinted with permission from [140]. Copyright (2006) Wiley-VCH Verlag GmbH and Co.
Figure 5. SEM image of the spherical particle PEs produced with MgCl2-supported Zr-FI catalyst Zr23. Reprinted with permission from [140]. Copyright (2006) Wiley-VCH Verlag GmbH and Co.
Polymers 15 03095 g005
Figure 6. (a) Optimized structure of γ-alumina. Reprinted with permission from [155]. Copyright (2008) American Chemical Society. (b) Crystal structure of boehmite. Reprinted with permission from [157]. Copyright (2021) International Union of Crystallography.
Figure 6. (a) Optimized structure of γ-alumina. Reprinted with permission from [155]. Copyright (2008) American Chemical Society. (b) Crystal structure of boehmite. Reprinted with permission from [157]. Copyright (2021) International Union of Crystallography.
Polymers 15 03095 g006
Figure 7. Top views of (a) fully dehydrated (100), (b) fully dehydrated (110), (c) hydrated (110), and (d) hydrated (111) terminations of γ-Al2O3, showing different types of surface Al atoms. X(μ-OH)w(OH)y represents an Al atom with a total coordination number of X, including w bridging μ-OH and y terminal OH groups. Only atoms in the top surface layers are shown for clarity. Reprinted with permission from [161]. Copyright (2014) American Chemical Society.
Figure 7. Top views of (a) fully dehydrated (100), (b) fully dehydrated (110), (c) hydrated (110), and (d) hydrated (111) terminations of γ-Al2O3, showing different types of surface Al atoms. X(μ-OH)w(OH)y represents an Al atom with a total coordination number of X, including w bridging μ-OH and y terminal OH groups. Only atoms in the top surface layers are shown for clarity. Reprinted with permission from [161]. Copyright (2014) American Chemical Society.
Polymers 15 03095 g007
Figure 8. {[(AlSO)2Zr(CH2tBu)+][AlS(CH2tBu)]} structure optimized on the model surface. Reprinted with permission from [162]. Copyright (2006) American Chemical Society.
Figure 8. {[(AlSO)2Zr(CH2tBu)+][AlS(CH2tBu)]} structure optimized on the model surface. Reprinted with permission from [162]. Copyright (2006) American Chemical Society.
Polymers 15 03095 g008
Scheme 8. Grafting reaction of Hf(CH2tBu)4 on γ-Al2O3 at room temperature [159].
Scheme 8. Grafting reaction of Hf(CH2tBu)4 on γ-Al2O3 at room temperature [159].
Polymers 15 03095 sch008
Figure 9. Structure–activity relationship for Zr25′ and Zr2′, supported on γ-Al2O3. Reprinted with permission from [82]. Copyright (2001) American Chemical Society. The number of “+” symbols reflects catalytic activity according to the assessment of the authors of [82], from very low (+) to very high (+++++).
Figure 9. Structure–activity relationship for Zr25′ and Zr2′, supported on γ-Al2O3. Reprinted with permission from [82]. Copyright (2001) American Chemical Society. The number of “+” symbols reflects catalytic activity according to the assessment of the authors of [82], from very low (+) to very high (+++++).
Polymers 15 03095 g009
Figure 10. (a) Aluminum and oxide sites exposed on the optimized γ-Al2O3 (110) surface; (b) dioxo- and oxo-bridged zirconocenium coordination on μ3-O and μ2-O surface sites. Reprinted with permission from [155]. Copyright (2008) American Chemical Society.
Figure 10. (a) Aluminum and oxide sites exposed on the optimized γ-Al2O3 (110) surface; (b) dioxo- and oxo-bridged zirconocenium coordination on μ3-O and μ2-O surface sites. Reprinted with permission from [155]. Copyright (2008) American Chemical Society.
Polymers 15 03095 g010
Figure 11. Calculated ethylene enchainment energy profiles for μ2-O coordinated (a) and μ3-O coordinated (b) Cp2ZrMe+ species; the energy profile for [Cp2ZrMe+][MeB(C6F5)3] is given for comparison (c). Reprinted with permission from [155]. Copyright (2008) American Chemical Society.
Figure 11. Calculated ethylene enchainment energy profiles for μ2-O coordinated (a) and μ3-O coordinated (b) Cp2ZrMe+ species; the energy profile for [Cp2ZrMe+][MeB(C6F5)3] is given for comparison (c). Reprinted with permission from [155]. Copyright (2008) American Chemical Society.
Polymers 15 03095 g011
Scheme 9. (a) Model for Brønsted acidic site (BAS) in silica-alumina [169]; (b) the structures of the products of the reaction of Zr25′ with SiO2/Al2O3 [82].
Scheme 9. (a) Model for Brønsted acidic site (BAS) in silica-alumina [169]; (b) the structures of the products of the reaction of Zr25′ with SiO2/Al2O3 [82].
Polymers 15 03095 sch009
Scheme 10. Formation of heterogeneous Ti and Zr metallocene catalysts [170,171].
Scheme 10. Formation of heterogeneous Ti and Zr metallocene catalysts [170,171].
Polymers 15 03095 sch010
Figure 12. Activities of homogeneous and heterogeneous ethylene polymerization of catalysts Ti16 and Ti17. Reprinted with permission from [170]. Copyright (2002) Elsevier B. V.
Figure 12. Activities of homogeneous and heterogeneous ethylene polymerization of catalysts Ti16 and Ti17. Reprinted with permission from [170]. Copyright (2002) Elsevier B. V.
Polymers 15 03095 g012
Scheme 11. Different activation pathways of a pre-catalyst complex considering the different nature of the acidic sites in the activating support: (a) via double alkylation; (b) via single alkylation and proton abstraction [175].
Scheme 11. Different activation pathways of a pre-catalyst complex considering the different nature of the acidic sites in the activating support: (a) via double alkylation; (b) via single alkylation and proton abstraction [175].
Polymers 15 03095 sch011
Figure 13. Ethylene polymerization activity using acidic oxide-supported (n-BuCp)2ZrCl2 (Zr9), with acid properties as determined by cumene cracking: (1) polymerization (bar): Zr9 = 4.6 μmol, acidic oxide = 20 mg, TIBA = 0.52 mmol (Al/Zr = 115), temperature = 60 °C, and ethylene pressure = 0.70 MPa; (2) cracking (bent line): reaction temperature = 400 °C, catalyst = 0.10 g, carrier (He) = 30 mL∙min−1, and pulse size = 0.5 μL; and (3) BET surface area of each oxide presented in parenthesis. Reprinted with permission from [167]. Copyright (2007) Elsevier B. V.
Figure 13. Ethylene polymerization activity using acidic oxide-supported (n-BuCp)2ZrCl2 (Zr9), with acid properties as determined by cumene cracking: (1) polymerization (bar): Zr9 = 4.6 μmol, acidic oxide = 20 mg, TIBA = 0.52 mmol (Al/Zr = 115), temperature = 60 °C, and ethylene pressure = 0.70 MPa; (2) cracking (bent line): reaction temperature = 400 °C, catalyst = 0.10 g, carrier (He) = 30 mL∙min−1, and pulse size = 0.5 μL; and (3) BET surface area of each oxide presented in parenthesis. Reprinted with permission from [167]. Copyright (2007) Elsevier B. V.
Polymers 15 03095 g013
Scheme 12. Activation of Cp2ZrMe2 by sulfated zirconia (ZrS) [83].
Scheme 12. Activation of Cp2ZrMe2 by sulfated zirconia (ZrS) [83].
Polymers 15 03095 sch012
Figure 14. (a) Immobilization of Hf3 (LHfMe2) on ZrS support; (b) 13C CPMAS NMR spectra (100 MHz; repetition time, 5 s; contact time, 2 ms; 10 kHz spinning speed) of (A) LHfMe2, 876 scans; (B) LHfMe+/ZrS, 10,000 scans; (C) 13C labeled LHf(13CH3)2 (60 scans); and (D) LHf13CH3+/ZrS, 1600 scans. *—Rotational sidebands, #—impurities. Reprinted with permission from [186]. Copyright (2018) American Chemical Society.
Figure 14. (a) Immobilization of Hf3 (LHfMe2) on ZrS support; (b) 13C CPMAS NMR spectra (100 MHz; repetition time, 5 s; contact time, 2 ms; 10 kHz spinning speed) of (A) LHfMe2, 876 scans; (B) LHfMe+/ZrS, 10,000 scans; (C) 13C labeled LHf(13CH3)2 (60 scans); and (D) LHf13CH3+/ZrS, 1600 scans. *—Rotational sidebands, #—impurities. Reprinted with permission from [186]. Copyright (2018) American Chemical Society.
Polymers 15 03095 g014
Scheme 13. Pyridylamido complexes of Hf, studied in activation and polymerization on sulfated zirconia [194].
Scheme 13. Pyridylamido complexes of Hf, studied in activation and polymerization on sulfated zirconia [194].
Polymers 15 03095 sch013
Figure 15. Energy-minimized computed chemisorbed Cp*Zr(CH2Ph)2+ catalyst structures on sulfated alumina surface between site SA (a) and SB (b) (Cp*—η5-C5Me5). Reprinted and corrected with permission from [180]. Copyright (2015) American Chemical Society.
Figure 15. Energy-minimized computed chemisorbed Cp*Zr(CH2Ph)2+ catalyst structures on sulfated alumina surface between site SA (a) and SB (b) (Cp*—η5-C5Me5). Reprinted and corrected with permission from [180]. Copyright (2015) American Chemical Society.
Polymers 15 03095 g015
Figure 16. (a) re and si face propylene insertion transition states with Zr25′/AlS; (b) computed Gibbs free energy profiles for si vs. re face propylene activation/insertion at the Zr25′/AlS catalytic center. Reprinted with permission from [196]. Copyright (2021) Wiley-VCH Verlag GmbH and Co.
Figure 16. (a) re and si face propylene insertion transition states with Zr25′/AlS; (b) computed Gibbs free energy profiles for si vs. re face propylene activation/insertion at the Zr25′/AlS catalytic center. Reprinted with permission from [196]. Copyright (2021) Wiley-VCH Verlag GmbH and Co.
Polymers 15 03095 g016
Scheme 14. Constrained-geometry Group 4 metal complexes, selected for activation with sulfated alumina [197].
Scheme 14. Constrained-geometry Group 4 metal complexes, selected for activation with sulfated alumina [197].
Polymers 15 03095 sch014
Figure 17. Effects of pre-catalyst structure and support on (a) activity and (b) oct-1-ene incorporation in ethylene copolymerization using the indicated surface molecular catalysts. Reprinted with permission from [194]. Copyright (2021) American Chemical Society.
Figure 17. Effects of pre-catalyst structure and support on (a) activity and (b) oct-1-ene incorporation in ethylene copolymerization using the indicated surface molecular catalysts. Reprinted with permission from [194]. Copyright (2021) American Chemical Society.
Polymers 15 03095 g017
Scheme 15. Metallocenes studied in ethylene polymerization using AlS/R3Al support [201].
Scheme 15. Metallocenes studied in ethylene polymerization using AlS/R3Al support [201].
Polymers 15 03095 sch015
Scheme 16. Two distinct approaches to fluorinated SiO2/Al2O3. (a) Fluorination of surface –OH groups in SiO2/Al2O3; (b) Surface modification of SiO2 by aluminum alkyls, followed by oxidation and fluorination. The content of Si–OH and Al–OH surface fragments depends on calcination temperature and duration.
Scheme 16. Two distinct approaches to fluorinated SiO2/Al2O3. (a) Fluorination of surface –OH groups in SiO2/Al2O3; (b) Surface modification of SiO2 by aluminum alkyls, followed by oxidation and fluorination. The content of Si–OH and Al–OH surface fragments depends on calcination temperature and duration.
Polymers 15 03095 sch016
Figure 18. The function of catalyst activity on calcination temperature for Zr9/Et3Al system and SiO2/Al2O3(F) activating supports, prepared with different NH4HF2 loadings. Designed based on the data presented in [89].
Figure 18. The function of catalyst activity on calcination temperature for Zr9/Et3Al system and SiO2/Al2O3(F) activating supports, prepared with different NH4HF2 loadings. Designed based on the data presented in [89].
Polymers 15 03095 g018
Scheme 17. (a) Structures of zirconocene pre-catalysts and catalytic activities in ethylene polymerization (in brackets, gPE∙gCat−1∙h−1) of the corresponding heterogeneous supported catalysts, prepared with the use of fluorinated SiO2/Al2O3 and Et3Al; (b) formation of cyclic organoaluminum species during ‘pre-contacting’ of Zr2/Et3Al with hex-1-ene [93].
Scheme 17. (a) Structures of zirconocene pre-catalysts and catalytic activities in ethylene polymerization (in brackets, gPE∙gCat−1∙h−1) of the corresponding heterogeneous supported catalysts, prepared with the use of fluorinated SiO2/Al2O3 and Et3Al; (b) formation of cyclic organoaluminum species during ‘pre-contacting’ of Zr2/Et3Al with hex-1-ene [93].
Polymers 15 03095 sch017
Scheme 18. (a) Thermal decomposition of (NH4)2SiF6; (b) preparation of SiO2/g-Al2O3(F) [107].
Scheme 18. (a) Thermal decomposition of (NH4)2SiF6; (b) preparation of SiO2/g-Al2O3(F) [107].
Polymers 15 03095 sch018
Scheme 19. Group 4 metal complexes, investigated with SiO2/g-Al2O3(F) activating supports.
Scheme 19. Group 4 metal complexes, investigated with SiO2/g-Al2O3(F) activating supports.
Polymers 15 03095 sch019
Figure 19. EDX analysis of the support AS1. Reprinted with permission from [107]. Copyright (2013) American Chemical Society.
Figure 19. EDX analysis of the support AS1. Reprinted with permission from [107]. Copyright (2013) American Chemical Society.
Polymers 15 03095 g019
Figure 20. Polyethylene particles obtained using the catalyst Zr5/iBu3Al/AS8 (heptane, 10 bar, 6 mL of hexene, 80 °C). Reprinted with permission from [107]. Copyright (2013) American Chemical Society.
Figure 20. Polyethylene particles obtained using the catalyst Zr5/iBu3Al/AS8 (heptane, 10 bar, 6 mL of hexene, 80 °C). Reprinted with permission from [107]. Copyright (2013) American Chemical Society.
Polymers 15 03095 g020
Scheme 20. Preparation of SiO2/g-Al2O3(F) activating supports by the method of INEOS [99]. (a) Preparation of fluorinated alkoxide or aryloxide; (b) Treatment of the silica surface followed by calcination in an inert atmosphere and in the air.
Scheme 20. Preparation of SiO2/g-Al2O3(F) activating supports by the method of INEOS [99]. (a) Preparation of fluorinated alkoxide or aryloxide; (b) Treatment of the silica surface followed by calcination in an inert atmosphere and in the air.
Polymers 15 03095 sch020
Figure 21. Model depicting the surface composition of PURAL SB (A), SIRALs 1.5–5 (B), SIRALs 10–20 (C), SIRALs 30–40 (D), SIRALs 60–80 (E), and SIRALs 90–100 (F). Reprinted with permission from [218]. Copyright (2000) Elsevier B. V.
Figure 21. Model depicting the surface composition of PURAL SB (A), SIRALs 1.5–5 (B), SIRALs 10–20 (C), SIRALs 30–40 (D), SIRALs 60–80 (E), and SIRALs 90–100 (F). Reprinted with permission from [218]. Copyright (2000) Elsevier B. V.
Polymers 15 03095 g021
Scheme 21. Group 4 metallocene pre-catalysts studied in olefin polymerization with the use of Al2O3/g-SiO2(F) activating supports.
Scheme 21. Group 4 metallocene pre-catalysts studied in olefin polymerization with the use of Al2O3/g-SiO2(F) activating supports.
Polymers 15 03095 sch021
Scheme 22. Group 4 metallocene pre-catalysts that have demonstrated high efficiency in the oligomerization of oct-1-ene when activated by SiO2/Al2O3(F) and iBu3Al [97].
Scheme 22. Group 4 metallocene pre-catalysts that have demonstrated high efficiency in the oligomerization of oct-1-ene when activated by SiO2/Al2O3(F) and iBu3Al [97].
Polymers 15 03095 sch022
Scheme 23. Zirconium complexes, studied in ethylene polymerization using (Ph3C)mHn[PMo12O40] as an activator [230].
Scheme 23. Zirconium complexes, studied in ethylene polymerization using (Ph3C)mHn[PMo12O40] as an activator [230].
Polymers 15 03095 sch023
Scheme 24. (a) Synthesis of tethered Brønsted acid sites by the reaction of silica with perfluorinated sultone; (b) hypothetical activated catalyst [231].
Scheme 24. (a) Synthesis of tethered Brønsted acid sites by the reaction of silica with perfluorinated sultone; (b) hypothetical activated catalyst [231].
Polymers 15 03095 sch024
Scheme 25. Preparation of a bis(pentafluorophenoxy)aluminate activator [54].
Scheme 25. Preparation of a bis(pentafluorophenoxy)aluminate activator [54].
Polymers 15 03095 sch025
Scheme 26. (a) Formation of Brønsted (a) and Lewis (b) acidic sites during the reaction of Al(OC(CF3)3)3∙PhF with silica; (c) activation of Cp2Zr(13CH3)2 [234]. The key intermediates A, B, C and D are discussed below.
Scheme 26. (a) Formation of Brønsted (a) and Lewis (b) acidic sites during the reaction of Al(OC(CF3)3)3∙PhF with silica; (c) activation of Cp2Zr(13CH3)2 [234]. The key intermediates A, B, C and D are discussed below.
Polymers 15 03095 sch026
Scheme 27. Structures of sulfated polymers used as activating supports for the Zr5/iBu3Al system [236].
Scheme 27. Structures of sulfated polymers used as activating supports for the Zr5/iBu3Al system [236].
Polymers 15 03095 sch027
Figure 22. Influence of the kind of alkylaluminum compound and the polymerization time on the PE yield obtained using the Cp2TiCl2 catalyst in biphasic [BMIM]+[AlCl4]/hexane polymerization. The PE yield is presented separately for both phases. The alkylaluminum compound: (a) MAO, (b) Et2AlCl (ionic liquid)/MAO (hexane), (c) AlEt2Cl. Polymerization condition: Ti = 0.06 μmol, Al = 4 mmol, 1 bar. Reprinted with permission from [245]. Copyright (2007) Springer Nature.
Figure 22. Influence of the kind of alkylaluminum compound and the polymerization time on the PE yield obtained using the Cp2TiCl2 catalyst in biphasic [BMIM]+[AlCl4]/hexane polymerization. The PE yield is presented separately for both phases. The alkylaluminum compound: (a) MAO, (b) Et2AlCl (ionic liquid)/MAO (hexane), (c) AlEt2Cl. Polymerization condition: Ti = 0.06 μmol, Al = 4 mmol, 1 bar. Reprinted with permission from [245]. Copyright (2007) Springer Nature.
Polymers 15 03095 g022
Scheme 28. Structures of ionic liquids studied as activators for Ti2/EtAlCl2 in the polymerization of ethylene [247].
Scheme 28. Structures of ionic liquids studied as activators for Ti2/EtAlCl2 in the polymerization of ethylene [247].
Polymers 15 03095 sch028
Scheme 29. Structures of ionic liquids studied as activators for Ti2/EtAlCl2 in the polymerization of ethylene [247].
Scheme 29. Structures of ionic liquids studied as activators for Ti2/EtAlCl2 in the polymerization of ethylene [247].
Polymers 15 03095 sch029
Figure 23. SEM images of the SIL-metallocene catalytic system (a) and the obtained polyethylene (b). Reprinted with permission from [247]. Copyright (2014) Elsevier B. V.
Figure 23. SEM images of the SIL-metallocene catalytic system (a) and the obtained polyethylene (b). Reprinted with permission from [247]. Copyright (2014) Elsevier B. V.
Polymers 15 03095 g023
Figure 24. Proposed structures of chemisorbed Cp2ZrMe2 on the surfaces of dehydroxylated Lewis acidic metal oxides (a), highly Brønsted acidic sulfated metal oxides (b), and weakly Brønsted acidic hydroxylated metal oxides (c). Reprinted with permission from [180]. Copyright (2015) American Chemical Society.
Figure 24. Proposed structures of chemisorbed Cp2ZrMe2 on the surfaces of dehydroxylated Lewis acidic metal oxides (a), highly Brønsted acidic sulfated metal oxides (b), and weakly Brønsted acidic hydroxylated metal oxides (c). Reprinted with permission from [180]. Copyright (2015) American Chemical Society.
Polymers 15 03095 g024
Table 1. Ethylene polymerization using (RCp)2ZrCl2 and (C5HnR5–n)2ZrCl2 immobilized on an MgCl2-based support [81] 1.
Table 1. Ethylene polymerization using (RCp)2ZrCl2 and (C5HnR5–n)2ZrCl2 immobilized on an MgCl2-based support [81] 1.
Zr Comp.Cp Ring
Substitution
Activity, kg∙mol−1
bar−1∙h−1
Tm, °CDegree of
Crystallinity
χc
Mn, kDaMw, kDaĐM
Zr2H160134.553.5883524.0
Zr6Et260135.155.61133943.5
Zr7n-Pr2760135.458.5952422.5
Zr8iPr360135.651.71944852.5
Zr9n-Bu3680135.356.91202712.3
Zr10tBu120135.153.71264253.4
Zr11n-Pentyl1120134.557.9983023.1
Zr12n-Dodecyl700135.653.81223442.8
Zr131,3-Me2140135.247.11724742.8
Zr141-Me-3-n-Bu880135.954.32174672.2
Zr151,2,4-Me3120136.150.11844302.3
Zr16Me4140139.247.31193923.3
Zr17Me5180139.247.31483912.6
1 Polymerization conditions: 500 mL of light petroleum, 100 mg of immobilized catalyst (1 μmol Zr), 1 mmol of AlEt3, 50 °C, ethylene pressure of 5 bar, 1 h.
Table 2. Calculated ion pair formation enthalpies ΔHform and heterolytic ion pair separation enthalpies ΔHips (kcal∙mol−1) for activation of Cp2ZrMe2 on the Al2O3 (110) surface and with the use of B(C6F5)3 [155].
Table 2. Calculated ion pair formation enthalpies ΔHform and heterolytic ion pair separation enthalpies ΔHips (kcal∙mol−1) for activation of Cp2ZrMe2 on the Al2O3 (110) surface and with the use of B(C6F5)3 [155].
Complex of Cp2ZrMe+d(Zr–O), Å 1ΔHform, kcal∙mol−1ΔHips, kcal∙mol−1
μ3-O dioxo-bridged2.82−16.777.2
μ3-O oxo-bridged2.58−21.782.1
μ2-O dioxo-bridged2.34−51.7112.2
μ2-O oxo-bridged2.13−73.3127.5
[MeB(C6F5)3] adduct2.56 2−1.777.0
1 In the case of dioxo-bridged structures, a mean distance is given. 2 The value of d(Zr–C) for Zr–(μ-Me)–B fragment.
Table 3. Ethylene polymerization on supported (AlS) and nonsupported constrained-geometry catalysts [197].
Table 3. Ethylene polymerization on supported (AlS) and nonsupported constrained-geometry catalysts [197].
CatalystT, °C[M], μmolReaction
Time, h
Activity, kg∙mol−1∙h−1Tm of PE,
°C
Zr29′/AlS254.4631.21138
Zr30′/AlS254.0233.51138
Zr30′/AlS602.8534.93137
Zr31′/AlS255.6031.66139
Zr31′/AlS604.4631.19137
Ti18′/AlS252.36314.4136
Ti18′/AlS602.19320.6137
Zr30′/BA 12510.00.75127n. d. 2
Zr31′/2 BA255.01.2587n. d.
1 Homogeneous polymerization, BA—[Ph3C]+[B(C6F5)4]. 2 No data, low-MW PE obtained (Mn = 610 and 760 Da, respectively).
Table 4. Ethylene polymerization activities of Zr complexes, supported on sulfated Al2O3 (30 bar, isobutane) [201].
Table 4. Ethylene polymerization activities of Zr complexes, supported on sulfated Al2O3 (30 bar, isobutane) [201].
Zr
Complex
T, °CTime, minAlS Weight, mgZr Comp. Weight, mgR3Al/
mmol
PE, gActivity, kg∙mmol−1∙h−1Mn, kDaMw, kDa
Hf690601003.0iBu3Al/0.529480.8296911
Hf6105351003.0iBu3Al/0.220395.6239730
Hf790601003.0iBu3Al/0.525268.1315972
Hf7105331003.0iBu3Al/0.218691.3318843
Zr3280301001.0iBu3Al/0.25315455.5305754
Zr3290301051.0iBu3Al/0.5295426.5263639
Zr3290301041.0Bu3Al/0.5320462.7278708
Zr3390301011.0iBu3Al/0.5272400.9223591
Zr3390301081.0Bu3Al/0.5211311.0314750
Zr349060502.0Bu3Al/0.515860.4311772
Zr3490301002.0iBu3Al/0.25255195.1205637
Zr3590601002.0iBu3Al/0.254214.0ins.
Zr35105601002.0iBu3Al/0.256321.1ins.
Zr369016501.0iBu3Al/0.5232655.270183
Hf890601003.0iBu3Al/0.529482.4108375
Hf8100601003.0iBu3Al/0.5369103.497267
Table 5. Ethylene polymerization activities of Zr complexes, supported on sulfated metal oxides [87].
Table 5. Ethylene polymerization activities of Zr complexes, supported on sulfated metal oxides [87].
SupportZr Complex[Zr],
molZr/mg
Time, minActivity, kg∙mol−1∙h−1Tm of PE,
°C
Active Site % 1
ZrSZr25′1.43 × 10−72770133.1~65
ZrSZr(CH2Ph)41.37 × 10−70.752500134.0
AlSZr25′3.9 × 10−8101100133.687 ± 3
AlSZr(CH2Ph)43.9 × 10−882100134.2
SnSZr25′4.4 × 10−86098136.161 ± 5
SnSZr(CH2Ph)44.4 × 10−830660135.1
FeSZr25′9.0 × 10−83024134.222 ± 2
FeSZr(CH2Ph)49.0 × 10−87510135.2
TiSZr25′1.85 × 10−71201.2138.463 ± 9
TiSZr(CH2Ph)41.85 × 10−76063138.0
1 Determined by poisoning Zr25′/SMO ethylene polymerization with neopentyl alcohol.
Table 6. Characteristics of the supports and ethylene polymerization results [214].
Table 6. Characteristics of the supports and ethylene polymerization results [214].
SupportCS of Type L1 and L2 LAS, μmol∙g−1CS of Type L3 LAS, μmol∙g−1CS of BAS, μmol∙g−1PA, kJ∙mol−1Zr2–Based CatalystZr2′–Based Catalyst
Zr wt% (μmol∙g−1)PE Yield, gActivity, kg∙mol−1∙h−1Zr wt% (μmol∙g−1)PE Yield, gActivity, kg∙mol−1∙h−1
γ-Al2O360, 6603513801.21 (133)3.3254.07 (447)4.59
Al2O3(1%F)38020412800.62 (68)111621.93 (212)6.430
Al2O3(3%F)18050211900.60 (66)203032.53 (278)46166
Al2O3(5%F)120507.311900.62 (68)618952.53 (278)149.5538
Al2O3(7%F)100506.111900.66 (73)304142.53 (278)51183
Table 7. The types and characteristics of SiO2; reagents and conditions of the preparation of SiO2/g-Al2O3(F) activating supports and their catalytic applications.
Table 7. The types and characteristics of SiO2; reagents and conditions of the preparation of SiO2/g-Al2O3(F) activating supports and their catalytic applications.
Silica/Specific Surface Area, m2∙g−1 /
Pore Volume, mL∙g−1/
Particle Size, μm
Organo-
Aluminum Compound
Stages 2, 3; Calcination RegimeStage 3: Fluorinating Reagent/3rd Calcination RegimePre-catalyst/wt% or (mmol∙g−1)R3AlPolymer/
Activity, g∙gCat−1∙h−1 (kg∙mol−1∙h−1)
Ref.
Grace 332/300/1.65/70(BuO)2AlOSi
(OEt)3
20→130 °C, 1 h
130→450 °C, 1 h
450 °C, 1 h
6 wt% (NH4)2SiF6 /
20→130 °C, 1 h
130→450 °C, 1 h
450 °C, 1 h
Zr2/(53)iBu3AlPE/—(900)[88]
Zr2/(53)Et3AlPE/—(490)[88]
Zr37/(53)iBu3AlPE/—(6720)[88]
Zr38/(53)iBu3AlPE/—(2020)[88] 1
Grace 332/300/1.65/70Et3Al20→130 °C, 1 h
130 °C, 1 h
130→450 °C, 1 h
450 °C, 4 h
5 wt% (NH4)2SiF6/
30→450 °C, 2 h
450 °C, 2 h
Zr5/n. d.iBu3AlPE/39 (3400)[107]
Grace 332/300/1.65/7010 wt% (NH4)2SiF6/
30→450 °C, 2 h
450 °C, 2 h
Zr5/n. d.iBu3Alethylene—hex-1-ene (20 mol%)/550 (75,000)[107]
Crossfield ES70X/276/1.54/53Zr5/n. d.iBu3Alethylene—hex-1-ene (20 mol%)/725 (91,000)[107]
Grace 332/300/1.65/70Et2AlF20→130 °C, 1 h
130 °C, 1 h
130→450 °C, 1 h
450 °C, 1 h
Zr38iBu3AlPE/820 (–)[102]
Grace 332/300/1.65/70Zr39/1.0iBu3AlPP/700 (–) 2[102,107]
Grace 332/300/1.65/70Zr5/iBu3Alethylene—hex-1-ene (20 mol%)/110 (19,000)[107]
MS-3030/320/3/85Zr5/0.4iBu3AlPE/160 (–)[105]
MS-3040/420/3/80Zr5/0.4iBu3AlPE/320 (–)–”–
SIL 3/800/1/15Zr5/0.4iBu3AlPE/0 (–)–”–
SIL 19/700/2.2/15Zr5/0.4iBu3AlPE/280 (–)–”–
SIL 20/630/3.2/100Zr5/0.4iBu3AlPE/400 (–)–”–
SIL 30/300/3/100Zr5/0.4iBu3AlPE/250 (–)–”–
SP9 446/520/1.9/53Zr5/0.4iBu3AlPE/1080 (–)–”–
Grace 948/290/1.7/58Zr5/0.4iBu3AlPE/700 (–)–”–
1 In [88], ethylene/hex-1-ene copolymerization results for Zr5/iBu3Al/AS are also mentioned without meaningful results. 2 The percentage of isotactic pentads mmmm = 96%.
Table 8. Preparation and elemental analysis of activating supports [107].
Table 8. Preparation and elemental analysis of activating supports [107].
SupportSilica 1Alkylaluminum(NH4)2SiF6 wt%Al wt%F wt%
AS1Grace 332Et3Al53.721.18
AS2Grace 332Et3Al54.751.74
AS3Grace 332Et3Al103.673.31
AS4Grace 332Et3Al203.815.30
AS5Grace 332Et2AlF4.562.21
AS6Grace 332Et3Al104.983.60
AS7ES70XEt3Al103.013.98
AS8 2ES70XEt3Al103.202.20
AS9 3ES70XEt3Al53.262.34
1 Properties of silicas are given in Table 2. Before treatment with alkylaluminum, silicas were calcined at 0.01 mbar and 450 °C, [Si–OH] = 1.3 mmol∙g−1. 2 Pre-calcining temperature was 200 °C. 3 Pre-calcining temperature was 300 °C.
Table 9. Activation of a range of Group 4 metal complexes using AS6, AS7, and MAO [107] 1.
Table 9. Activation of a range of Group 4 metal complexes using AS6, AS7, and MAO [107] 1.
CompActivatorHex-1-ene mol%Activity, kg∙mmol−1∙h−1Activity, g∙gCat−1∙h−1Mw, kDaĐMTm, °C
Zr5AS6 (20 mg)2075550973.7111
Zr5AS7 (20 mg)20917251253.6112
Zr9MAO (2000 eq)0150 2062.9134
Zr9MAO (2000 eq)2076 1603.0134
Zr9AS6 (21 mg)209.1711023.0124
Zr9AS7 (23 mg)0281421462.3131
Zr9AS7 (21 mg)2022462943.1119
Zr26 2MAO (2000 eq)0140 7045.1140
Zr26 2MAO (2000 eq)2027 1542.190
Zr26 2AS6 (20 mg)20241963465.281
Zr26 3AS7 (43 mg)04.2295573.6131
Zr26 2AS7 (26 mg)20321891933.384
Zr3MAO (2000 eq)02.6 1613.7131
Zr3 2MAO (2000 eq)209.5 512.0121
Zr3 2AS6 (48 mg)20332241444.0116
Zr3AS7 (41 mg)02.9233424.1135
Zr3 2AS7 (42 mg)20433381223.5111
Ti8MAO (2000 eq)01.1 1733.3138
Ti8 2MAO (2000 eq)203.9 151.9100
Ti8 2AS6 (108 mg)20382871863.3103
Ti8AS7 (112 mg)03.32410225.7133
Ti8 2AS7 (112 mg)2012953565.5101
1 300 mL heptane, 4 bar ethylene, 80 °C, 60 min, MAO 10 wt %. For Zr5, Zr9, and Zr26, [Zr] = 0.5 μmol, [Al] = 1 mmol; Zr3 [Zr] = 2 μmol, [Al] = 4 mmol; and for Ti8 [Ti] = 2.5 μmol, [Al] = 5 mmol. 2 15 min. 3 [Zr] = 1 μM.
Table 10. Characteristics of the silica used for the synthesis of the SiO2/g-Al2O3(F) supports [106].
Table 10. Characteristics of the silica used for the synthesis of the SiO2/g-Al2O3(F) supports [106].
SilicaPore Volume, mL∙g−1Specific Surface Area, m2∙g−1Particle Size, μmMean Pore
Diameter, Å
Grace Sylopol 9521.829058246
Grace Sylopol 9481.729058232
SP9_4461.952053131
M3703/079A1.924053240
Table 11. Catalytic activities of zirconocenes Zr5 and Zr40 on different activating supports in ethylene/hex-1-ene copolymerization and the values of tan δ for the copolymers obtained [94].
Table 11. Catalytic activities of zirconocenes Zr5 and Zr40 on different activating supports in ethylene/hex-1-ene copolymerization and the values of tan δ for the copolymers obtained [94].
Activating Support
(Al2O3/SiO2 by Weight)
Zr5Zr40
Activity, kg∙gCat−1∙h−1Tan δActivity, kg∙gCat−1∙h−1Tan δ
Sulfated Al2O3 (–)0.182.891.2714.43
Fluorinated Al2O3 (–)0.021.343.114.12
Chlorinated Al2O3 (–)0.071.880.361.53
SiO2/Al2O3 (0.15:1)0.151.171.617.05
Al2O3/g-SiO2(F) (2.6:1)9.105.786.355.44
Sulfated Al2O3/g-SiO2(F) (2.6:1) 16.693.694.949.99
Phosphated Al2O3/g-SiO2(F) (P) 26.415.005.138.79
1 Al2O3/g-SiO2 was treated by MeOH solution of NH4HF2 and H2SO4 with subsequent calcining at 600 °C for 3 h. 2 H3PO4 was used instead of H2SO4.
Table 12. Oligomerization of 1-octene, catalyzed by zirconium complexes, and activated by TIBA and fluorinated aluminosilicate [97].
Table 12. Oligomerization of 1-octene, catalyzed by zirconium complexes, and activated by TIBA and fluorinated aluminosilicate [97].
Cat.[Mon]/[Zr] RatioT, °CTOF, h−1KV100VI
Zr21.1·105903.1·1038.9211
Zr54.2·1051154.3·104136210
Zr61.5·106701.3·105130222
Zr83.1·1051103.4·10462186
Zr95.6·1051054.1·10445175
Zr351.1·105909.0·10323169
Zr434.3·1051203.1·104159214
Zr444.4·1051203.6·104132200
Zr451.0·1051005.2·10310.3194
Hf105.5·105902.5·1048.3157
Table 13. Ethylene (co)-polymerization results from the combination of AS-1 and sMAO solid activators with zirconium metallocene complexes Zr5 and Zr9 (80 °C, 4 bar, [iBu3Al] = 1 mmol, heptane, 30 min) [54].
Table 13. Ethylene (co)-polymerization results from the combination of AS-1 and sMAO solid activators with zirconium metallocene complexes Zr5 and Zr9 (80 °C, 4 bar, [iBu3Al] = 1 mmol, heptane, 30 min) [54].
CatalystMol% hex-1-eneAlsurface/
Zr Ratio
Wt% 1[Zr], μmolActivity,
g∙gCat−1∙h−1
Activity,
kg∙molZr−1∙h−1
Zr5/AS-19.61.182.02508800
Zr5/AS-119.59.31.222.0565622,500
Zr9/AS-19.01.212.01304300
Zr9/AS-119.59.01.212.01334400
Zr5/sMAO1201.562.12005370
Zr5/sMAO19.51201.562.162016,600
Zr9/sMAO1481.322.032611,400
Zr9/sMAO19.51481.552.02308600
1 Weight percent of metallocene in the supported metallocene catalysts.
Table 14. Influence of the kind of the ionic liquid (IL) and the [AlEtCl2]/[Ti2] molar ratio on PE yield in the biphasic IL/hexane polymerization of ethylene (30 °C, 5 bar).
Table 14. Influence of the kind of the ionic liquid (IL) and the [AlEtCl2]/[Ti2] molar ratio on PE yield in the biphasic IL/hexane polymerization of ethylene (30 °C, 5 bar).
IL[AlEtCl2]/
[Ti2] Ratio
Reaction Time, hPE Yield, kgPE∙molTi−1
IL PhaseHexane PhaseTotal
[C8-mim]+
[AlCl4]
67110.746.757.3
100114.462.076.4
133117.0101.0118.0
167178.085.7163.7
100213.7146.7160.3
133248.3147.0195.3
[C8-β-mpy]+
[AlCl4]
6718.068.076.0
100176.736.7113.3
1331101.741.0142.7
67265.737.6103.3
1332103.3105.0208.3
[C8-γ-mpy]+
[AlCl4]
6715.359.865.1
100194.342.0136.3
1331171.537.5209.0
6728.981.190.1
133242.0236.7278.7
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

Nifant’ev, I.E.; Komarov, P.D.; Kostomarova, O.D.; Kolosov, N.A.; Ivchenko, P.V. MAO- and Borate-Free Activating Supports for Group 4 Metallocene and Post-Metallocene Catalysts of α-Olefin Polymerization and Oligomerization. Polymers 2023, 15, 3095. https://doi.org/10.3390/polym15143095

AMA Style

Nifant’ev IE, Komarov PD, Kostomarova OD, Kolosov NA, Ivchenko PV. MAO- and Borate-Free Activating Supports for Group 4 Metallocene and Post-Metallocene Catalysts of α-Olefin Polymerization and Oligomerization. Polymers. 2023; 15(14):3095. https://doi.org/10.3390/polym15143095

Chicago/Turabian Style

Nifant’ev, Ilya E., Pavel D. Komarov, Oksana D. Kostomarova, Nikolay A. Kolosov, and Pavel V. Ivchenko. 2023. "MAO- and Borate-Free Activating Supports for Group 4 Metallocene and Post-Metallocene Catalysts of α-Olefin Polymerization and Oligomerization" Polymers 15, no. 14: 3095. https://doi.org/10.3390/polym15143095

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop