Next Article in Journal
Controlled Polymerization
Next Article in Special Issue
Novel 1,2-Bismethacrylate-3-Eugenyl Propane for Resin Composites: Synthesis, Characterization, Rheological, and Degree of Conversion
Previous Article in Journal
Crazing Initiation and Growth in Polymethyl Methacrylate under Effects of Alcohol and Stress
Previous Article in Special Issue
Copper II Complexes Based on Benzimidazole Ligands as a Novel Photoredox Catalysis for Free Radical Polymerization Embedded Gold and Silver Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Hydrocarbon Soluble, Molecular and “Complete” Al-Cocatalyst for High Temperature Olefin Polymerization

by
Gaia Urciuoli
1,2,3,
Francesco Zaccaria
1,3,*,
Cristiano Zuccaccia
2,3,*,
Roberta Cipullo
1,3,*,
Peter H. M. Budzelaar
1,3,
Antonio Vittoria
1,3,
Christian Ehm
1,3,
Alceo Macchioni
2,3 and
Vincenzo Busico
1,3
1
Department of Chemical Sciences, Federico II University of Naples, 80126 Napoli, Italy
2
Department of Chemistry, Biology and Biotechnology and CIRCC, University of Perugia, 06123 Perugia, Italy
3
Dutch Polymer Institute (DPI), P.O. Box 902, 5600 AX Eindhoven, The Netherlands
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(6), 1378; https://doi.org/10.3390/polym15061378
Submission received: 3 February 2023 / Accepted: 1 March 2023 / Published: 9 March 2023
(This article belongs to the Special Issue Polymers Synthesis and Characterization II)

Abstract

:
The dinuclear aluminum salt {[iBu2(DMA)Al]2(μ-H)}+[B(C6F5)4] (AlHAl; DMA = N,N-dimethylaniline) is the prototype of a new class of molecular cocatalysts for catalytic olefin polymerization, its modular nature offering easy avenues for tailoring the activator to specific needs. We report here, as proof of concept, a first variant (s-AlHAl) bearing p-hexadecyl-N,N-dimethylaniline (DMAC16) units, which enhances solubility in aliphatic hydrocarbons. The novel s-AlHAl was used successfully as an activator/scavenger in ethylene/1-hexene copolymerization in a high-temperature solution process.

Graphical Abstract

1. Introduction

Several high-performance polyolefin resins are produced with molecular catalysts in solution [1,2,3]. The active species are typically ion pairs of a cationic group 4 metal-alkyl complex and a weakly coordinating anion [4]. Said ion pairs are generated in situ by the reaction between a neutral precatalyst and a cocatalyst (activator), both playing key roles; in fact, the success story of industrial molecular olefin polymerization catalysis must be traced to the identification of high-performing combinations of (pre)catalysts [2,3,4,5,6,7,8] and cocatalysts [9,10,11,12,13]. This notwithstanding, precatalyst diversification attracted great attention, whereas reports on cocatalysts are more limited [14,15,16,17,18,19,20,21,22,23,24,25].
In the framework of a more general study of the aforementioned activation process [26], we have recently identified a novel Al-based cocatalyst, namely {[iBu2(PhNMe2)Al]2(μ-H)}+[B(C6F5)4] (AlHAl, Figure 1, left) [27]. This unusual homodinuclear cation has a distinctive bridging hydride between two Al atoms. While the two Al centers are coordinatively saturated, AlHAl possesses “latent” Lewis acidity; indeed, we reported indications that it can release the unsaturated [AliBu2(PhNMe2)]+ fragment [27], resembling [AlMe2]+ release from methylalumoxane (MAO) [28,29,30,31,32,33]. AlHAl can be viewed as a “tamed” version of the highly unstable [AliBu2]+ species [34,35], and yet it is able to competently activate group 4 metal precatalysts (and on top of that to scavenge impurities). Activation of a prototypical ansa-zirconocene precatalyst turned out to occur at [Al]/[Zr] and [Zr] values as low as 50 and ~10−7 M, respectively [27]; catalyst performance in propene polymerization was comparable to that with MAO activation, even though the latter required 100-fold to 1000-fold larger [Al]/[Zr]. Moreover and at odds with other well-known molecular activators (e.g., borate salts of organic Lewis and Brønsted acids) [4,10], AlHAl is able to activate dichloride precatalysts with no need for a separate alkylating and scavenging agent, like e.g., tri-iso-butylaluminum (TIBAL). Last but not least, the ease of synthesis and the modular nature of AlHAl (comprised of Al-alkyl groups, neutral N-ligands, a bridging hydride and a borate anion) makes it the prototype of a new cocatalyst class, potentially opening the door to tailored applications.
On the other hand, like all ionic activators (and MAO too) AlHAl is poorly soluble in aliphatic hydrocarbons, which represents a major drawback for use in homogeneous phase [16,19,20,24,36,37,38]. To overcome this problem, we modified the N-donor fragments of AlHAl by installing p-hexadecyl residues on the phenyl rings of aniline. In this paper, we report on the synthesis and characterization of the thus obtained s-AlHAl species (Figure 1, right), its solubility in pentane, and the performance in the activation of a representative ansa-zirconocene. The productivity of the resulting catalyst system in hydrocarbon solution at high temperature was evaluated in ethene/1-hexene copolymerization, comparatively with conventional activators.

2. Materials and Methods

All manipulations were performed under rigorous exclusion of oxygen and moisture in flame-dried Schlenk-type glassware interfaced to a high-vacuum line (<10−5 Torr), or in a nitrogen-filled MBraun glovebox (<0.5 ppm O2). Molecular sieves (4Å, MS) were activated for 24 h at ca. 200–230 °C under dynamic vacuum. Benzene-d6 and toluene-d8 were freeze–pump–thaw degassed on a high-vacuum line, dried over Na/K alloy, vacuum-transferred to a dry storage tube with a PTFE valve and stored over activated MS in the glovebox. TTB (Boulder Scientific Co., Longmont, Colorado), AB (Boulder Scientific Co., Longmont, CO, USA), TIBAL (Merck, Milan, Italy), DIBAL-H (Merck, Milan, Italy), 2M HCl in ether (Merck, Milan, Italy), 2,6-di-tert-butyl-4-methylphenol (BHT; Merck, Milan, Italy), Li[B(C6F5)4] etherate complex (abcr), DMAC16 (Ricci Chimica, Perugia, Italy), and 1,2-dichlorobenzene (>99.8% isomeric purity; Romil, Cambridge, UK) were purchased and used as received. Ethene (99.95%; Linde, Arluno—MI, Italy) was purified by flowing through a column containing activated 4 Å molecular sieves and an activated Cu catalyst (BASF R0-11G). 1-hexene (99%; Merck, Milan, Italy) was purified by passing it through a mixed-bed column of activated Cu catalyst and 4 Å molecular sieves. Toluene (Romil, Cambridge, UK), pentane (Romil, Cambridge, UK), and ISOPAR-G were dried by using an MBraun SPS-5 solvent purification unit. Dichloromethane (Carlo Erba, Cornaredo, Italy) was dried over calcium hydride, distilled and freeze–pump–thaw degassed on a high-vacuum line. Cat-Zr was synthesized according to reported procedures [39] and kindly donated by the group of Prof. A. Voskoboynikov from Lomonosov Moscow State University, Russia.

2.1. NMR Spectroscopy Experiments

NMR experiments were performed by using a Bruker Avance III HD 400 instrument equipped with a smart probe (400 MHz for 1H). 1H NMR spectra were referenced to the residual protons of the deuterated solvent used; 13C NMR spectra were referenced internally to the D-coupled 13C resonances of the NMR solvent. To describe the multiplicity of the signals, the following abbreviations are used: s, singlet; bs, broad singlet; d, doublet; bd, broad doublet; dd, doublet of doublets; t, triplet; and m, multiplet.
DOSY NMR measurements were performed at 298 K in pentane by using a coaxial capillary containing acetone-d6. The standard double-stimulated echo pulse sequence without spinning was used. The shape of the gradients was rectangular, their duration (d) was 4 ms, and their strength (G) was varied during the experiments. All the spectra were acquired by using 16 K data points, 32 increments, between 16 and 128 scans, a spectral width of 6000 Hz, an acquisition time of 1.3 s, and a relaxation delay of 2 s per transient, and processed by using the inverse Laplace transform (ILT) routine of the Bruker Dynamics Center software suite. Pentane was used as internal standard to account for temperature and viscosity fluctuations; its Dt was calibrated by using an external sample of HDO in D2O under the same conditions [40]. The Dt data were treated as described in the literature to derive the hydrodynamic dimensions [41,42]. The uncertainties were estimated by assuming an error of approximately 5% in the hydrodynamic radius, and 15% in hydrodynamic volume and aggregation number [43,44]; error intervals in Kaggr are given at 95% confidence.
The average aggregation number was calculated as the ratio between the average hydrodynamic volume of the aggregate (VH) and that of the single ion pair (VH0) [45]. The latter was estimated as the sum of the hydrodynamic volume of the cation (VH+0) and the anion (VH−0) by assuming that
VH0 = VH+0 + VH−0 = 2 × (VHDMAC16 + 2/3 × VHTIBAL) + VH−0,
where the hydrodynamic volume of the free DMAC16 molecule (VHDMAC16 = 659 ± 99 Å3) and that of TIBAL (VHTIBAL = 381 ± 57 Å3) were estimated by performing ad-hoc DOSY NMR experiments, whereas that of the borate anion (VH−0 = 605 ± 91 Å3) was obtained from the literature [44], resulting in VH0 = 2432 ± 365 Å3.
Experimental trends of N vs. C (see Section 3.1) were analyzed as reported in the literature [44] by using three different models for indefinite self-aggregation data: EK (equal K), IK (incremental K), and AK (attenuated K) [46]. The IK gave a better fit than the EK model (as determined by the Euclidean norm of the residuals with respect to the experimental N values, being 0.594 and 9.510, respectively), whereas the AK model was unable to reach convergence.

2.2. Synthetic Procedures

[(p-C16H33)PhN(Me)2H]+[B(C6F5)4]. Taking inspiration from literature procedures [47], a 2 M solution of HCl in diethyl ether (0.44 mL, 0.88 mmol) was added dropwise to a stirred solution of DMAC16 (304 mg, 0.88 mmol) in pentane (15 mL). A fluffy white solid rapidly precipitated, and the resulting suspension was stirred at room temperature for 30 min. The volatiles were removed under vacuum. The white residue, corresponding to [(p-C16H33)PhN(Me)2H]Cl, was suspended in fresh pentane and isolated by filtration (275 mg). An aliquot of [(p-C16H33)PhN(Me)2H]Cl (150 mg, 0.39 mmol) was then transferred in a Schlenk flask along with Li[B(C6F5)4] etherate complex (352 mg, 0.39 mmol) and dissolved in dichloromethane (10 mL). The resulting turbid mixture was stirred at room temperature for 3 h, observing formation of a white precipitate over time. The volatiles were removed under vacuum and the residue was extracted with toluene, filtered, dried under vacuum, and washed with pentane to obtain [(p-C16H33)PhN(Me)2H]+[B(C6F5)4] as a white solid (277 mg, 66% yield).
1H NMR (400 MHz, benzene-d6, 298 K): 6.83 (d, 2H, 3J = 8.8 Hz, aromatic m-CH), 6.15 (d, 2H, aromatic o-CH), 5.13 (bs, 1H, N-H), 2.34 (t, 2H, 3J = 7.6 Hz, -CH2-CH2-(CH2)12-CH2-CH3), 1.66 (s, 6H, N-CH3), 1.43 (m, 2H, -CH2-CH2-(CH2)12-CH2-CH3), 1.37–1.27 (m, 24H, -CH2-CH2-(CH2)12-CH2-CH3), 1.30 (m, 2H, -CH2-CH2-(CH2)12-CH2-CH3), 0.91 (t, 3H, 3J = 7.0 Hz, -CH2-CH2-(CH2)12-CH2-CH3) ppm. 13C NMR (100 MHz, benzene-d6, 298 K): 147.3 (quaternary aromatic C), 137.1 (quaternary aromatic C-N), 130.5 (aromatic m-CH), 118.1 (aromatic o-CH), 46.5 (N-CH3), 35.0 (-CH2-CH2-(CH2)12-CH2-CH3), 31.0 (-CH2-CH2-(CH2)12-CH2-CH3), 29.9–29.1 (-CH2-CH2-(CH2)12-CH2-CH3), 22.7 (-CH2-CH2-(CH2)12-CH2-CH3), 14.0 (-CH2-CH2-(CH2)12-CH2-CH3) ppm.
s-Al and s-AlHAl. TIBAL (42 µL, 166 mmol) was added to a stirred suspension of [(p-C16H33)PhN(Me)2H]+[B(C6F5)4] (150 mg, 141 mmol) and DMAC16 (54 mg, 156 mmol) in toluene (2 mL) at room temperature. The mixture rapidly turns from deep to pale orange, and effervescence is observed. After stirring for 16h, the volatiles were removed under vacuum, and the residue washed with pentane to give s-Al as a pale orange ionic liquid (170 mg, 80% yield). An aliquot of s-Al (30 mg, 20 mmol) was then dissolved in a solution of DIBAL-H (20 mmol) in ISOPAR-G (8 mL, for polymerization experiments) or pentane (8 mL, for synthetic purposes) to obtain s-AlHAl, which can be isolated in quantitative yield as a pale orange oil upon removal of volatiles under vacuum.
s-Al. 1H NMR (400 MHz, benzene-d6, 298 K): 7.00 (d, 4H, 3J = 8.8 Hz, aromatic m-CH), 6.65 (d, 4H, aromatic o-CH), 2.50 (t, 4H, 3J = 7.8 Hz, -CH2-CH2-(CH2)12-CH2-CH3), 1.99 (s, 12H, N-CH3), 1.55 (m, 4H, -CH2-CH2-(CH2)12-CH2-CH3), 1.51 (m, 2H, Al-CH2-CH(CH3)2), 1.40–1.29 (m, 48H, -CH2-CH2-(CH2)12-CH2-CH3), 1.30 (m, 4H, -CH2-CH2-(CH2)12-CH2-CH3), 0.95 (d, 12H, 3J = 6.4 Hz, Al-CH2-CH(CH3)2), 0.92 (t, 6H, 3J = 7.0 Hz, -CH2-CH2-(CH2)12-CH2-CH3), 0.05 (d, 4H, 3J = 6.8 Hz, Al-CH2-CH(CH3)2) ppm. 13C NMR (100 MHz, benzene-d6, 298 K): 144.1 (quaternary aromatic C), 142.4 (quaternary aromatic C-N), 129.7 (aromatic m-CH), 120.4 (aromatic o-CH), 46.7 (N-CH3), 34.8 (-CH2-CH2-(CH2)12-CH2-CH3), 31.2 (-CH2-CH2-(CH2)12-CH2-CH3), 29.9–29.1 (-CH2-CH2-(CH2)12-CH2-CH3), 27.8 (Al-CH2-CH(CH3)2), 25.8 (Al-CH2-CH(CH3)2), 22.8 (-CH2-CH2-(CH2)12-CH2-CH3), 20.5 (Al-CH2-CH(CH3)2), 14.0 (-CH2-CH2-(CH2)12-CH2-CH3) ppm.
s-AlHAl. 1H NMR (400 MHz, benzene-d6, 298 K): 7.04 (d, 4H, 3J = 8.8 Hz, aromatic m-CH), 6.82 (d, 4H, aromatic o-CH), 2.91 (bs, 1H, AlHAl), 2.48 (t, 4H, 3J = 7.8 Hz, -CH2-CH2-(CH2)12-CH2-CH3), 2.43 (s, 12H, N-CH3), 1.53 (m, 4H, -CH2-CH2-(CH2)12-CH2-CH3), 1.62 (m, 4H, Al-CH2-CH(CH3)2), 1.41–1.23 (m, 48H, -CH2-CH2-(CH2)12-CH2-CH3), 1.31 (m, 4H, -CH2-CH2-(CH2)12-CH2-CH3), 0.94 (t, 6H, 3J = 7.0 Hz, -CH2-CH2-(CH2)12-CH2-CH3), 0.92 (d, 12H, 3J = 6.6 Hz, Al-CH2-CH(CH3)2), 0.87 (d, 12H, 3J = 6.6 Hz, Al-CH2-CH(CH3)2), 0.10 (d, 8H, 3J = 7.0 Hz, Al-CH2-CH(CH3)2) ppm. 13C NMR (100 MHz, benzene-d6, 298 K): 143.8 (quaternary aromatic C), 142.1 (quaternary aromatic C-N), 130.0 (aromatic m-CH), 119.9 (aromatic o-CH), 46.8 (N-CH3), 34.8 (-CH2-CH2-(CH2)12-CH2-CH3), 31.2 (-CH2-CH2-(CH2)12-CH2-CH3), 29.9–29.1 (-CH2-CH2-(CH2)12-CH2-CH3), 27.7 (Al-CH2-CH(CH3)2), 25.5 (Al-CH2-CH(CH3)2), 22.7 (-CH2-CH2-(CH2)12-CH2-CH3), 21.9 (Al-CH2-CH(CH3)2), 14.0 (-CH2-CH2-(CH2)12-CH2-CH3) ppm.

2.3. Polymerization Experiments

Ethene/1-hexene copolymerization experiments were performed in a Freeslate Parallel Pressure Reactor setup with 48 reaction cells (PPR48), fully contained in a triple MBraun glovebox operating under nitrogen. The cells, each with a liquid working volume of 6.0 mL, featured an 800 rpm magnetically coupled stirring, and individual online reading/control of temperature, pressure, monomer uptake, and monomer uptake rate. Experiments were carried out according to established experimental protocols [27,39,48,49,50,51], under the following experimental conditions: ISOPAR-G solvent; Tp = 100 °C; p(C2H4) = 12 bar (175 psi; monomer fed on demand in a semi-batch fashion); 1-hexene initial concentration in reaction solution of 3.7 v/v% was used (see Section 3.2). Conversion of 1-hexene was kept below 15% to ensure a constant comonomer concentration. All experiments were performed at least in duplicate.
Prior to the execution of a polymerization library, the PPR modules undergo “bake-and-purge” cycles overnight (8 h at 90–140 °C with intermittent dry N2 flow), to remove any contaminants and leftovers from previous experiments. After cooling to glovebox temperature, the module stir tops are taken off, and the 48 cells are fitted with disposable 10-mL glass inserts (preweighed in a Bohdan Balance Automator, Mettler–Toledo, Milan, Italy) and polyether ether ketone (PEEK) stir paddles. The stir tops are then set back in place, and N2 in the reactors is replaced with ethene (ambient pressure).
The cells are then loaded with the appropriate amounts of a mixed alkane diluent (ISOPAR-G), containing TIBAL as a scavenger where appropriate, and 1-hexene. The system is then thermostated at the reaction temperature and brought to 12 bar of pressure with ethene. At this point, the catalyst injection sequence is started; aliquots of (a) an ISOPAR-G “chaser”, (b) a toluene solution of catalyst, (c) a toluene solution of the proper activator, and (d) a ISOPAR-G “buffer” are uploaded into the needle and subsequently injected into the cell of destination in reverse order, thus starting the reaction. This is left to proceed under stirring (800 rpm) at constant temperature and pressure with feed of ethene on demand until the desired monomer consumption has been reached (for reaction time; see Table S4), and quenched by overpressurizing the cell with 50 psi (3.4 bar) of dry air (preferred over other possible catalyst quenchers because in case of cell or quench line leakage oxygen is promptly detected by the dedicated glovebox sensor).
Once all cells have been quenched, the modules are cooled down to glovebox temperature and vented, the stir tops are removed, and the glass inserts containing the reaction phases are taken out and transferred to a centrifugal evaporator (RVC 2-33 CDplus, Martin Christ, Osterode am Harz, Germany), where all volatiles are removed, and the polymers are thoroughly dried overnight. The dry polymers were robotically weighed while still in the reaction vials, subtracting the prerecorded tare. Reaction yields are double-checked against online monomer conversion measurements. Polymer aliquots are then sent to the characterizations.
All polymers were characterized by means of high-temperature gel permeation chromatography (GPC) and 13C NMR spectroscopy. GPC curves were recorded with a Freeslate Rapid GPC setup, equipped with a set of two mixed-bed Agilent PLgel 10-μm columns and a Polymer Char IR4 detector. Calibration was performed with the universal method by using 10 monodisperse polystyrene samples (Mn between 1.3 and 3700 kDa).
Quantitative 13C NMR spectra were recorded by using a Bruker Avance III 400 spectrometer equipped with a high-temperature cryoprobe for 5-mm OD tubes, on 45-mg mL−1 polymer solutions in tetrachloroethane-d2 (with BHT added as a stabilizer, [BHT] = 0.4 mg mL−1). Acquisition conditions were: 45° pulse; acquisition time, 2.7 s; relaxation delay, 5.0 s; 2 K transients. Broadband proton decoupling was achieved with a modified WALTZ16 sequence (BI_WALTZ16_32 by Bruker, Billerica, MA, USA).

3. Results and Discussion

3.1. Synthesis and Characterization

s-AlHAl was synthesized by adapting the previously optimized procedure for AlHAl and its mononuclear precursor Al [27], replacing N,N-dimethylaniline (DMA) with its p-hexadecyl substituted analogue (DMAC16). Reaction of equimolar amounts of TIBAL, substituted aniline and the corresponding anilinium borate provided access to the mononuclear species s-Al [52], which was then converted into the dinuclear product by addition of one equivalent of di-iso-butylaluminum hydride (DIBAL-H; Scheme 1).
The alkyl substituent on the phenyl ring does not significantly affect the overall structure. The NMR spectra of s-AlHAl (see Supplementary Materials) are analogous to those of AlHAl, except for the additional set of signals of the p-hexadecyl residues. Likewise, the equilibrium constant (Keq) for the reaction
s-Al + [AliBu2H]3s-AlHAl + [AliBu2H]2
in toluene-d8 at 298 K was estimated to be approximately 24, corresponding to a ΔG ≈ −1.9 kcal/mol, implying that s-AlHAl is the largely dominant species in solution (Figure S3 in Supplementary Materials). These values are comparable within experimental uncertainty with those reported for Al and AlHAl under the same conditions (Keq = 12; ΔG ≈ −1.5 kcal/mol) [27], indicating only marginal effects of the para-alkyl substituents on the relative stability of the mono- and dinuclear species.
At the same time, solubility of s-AlHAl in saturated hydrocarbons turned out to be significantly enhanced. Pentane solutions with concentration of (at least) 7 × 10−2 M can be easily prepared. This value is well adequate, considering that typical catalyst concentrations in industrial reactors are as low as 10−7 M, and that (as noted before) less than 102 equivalents of this novel activator are sufficient for efficient catalyst activation and impurity scavenging [27] (with MAO, which is the preferred cocatalyst in commercial processes, [Al]/[M] ratios between 103 and 105 are often necessary [9,53]).
Importantly, the mononuclear species s-Al is far less soluble than s-AlHAl; this is convenient from the synthetic point of view because it allows for easy purification of the mononuclear species by washing with pentane (see Materials and Methods). The difference can likely be ascribed to the larger nature of the cation and the more delocalized charge of the dinuclear species [27], facilitating solvation by nonpolar solvents.
The high solubility of the alkyl-substituted cocatalyst also allowed recording NMR spectra in pure pentane at 298 K (Figure 2). Although the concentrations of s-Al and DIBAL-H could not be reliably estimated under these conditions, dinuclear s-AlHAl appears to be the dominant species in solution, similar to what is observed in toluene-d8. Furthermore, diffusion-ordered NMR spectroscopy (DOSY) experiments [54,55] could be carried out to investigate the tendency of s-AlHAl to aggregate in this representative low-polarity solvent [45].
The trend of average aggregation number (N) vs. concentration (C; Figure 2c) was estimated in the range 0.6–33 mM at 298 K. Experimental datapoints were fitted with the incremental Kaggr (IK) model for indefinite self-association [43,44,46,56] to extrapolate the reference equilibrium constant Kaggr, IK = 3343 ± 1032 M−1 and the corresponding ΔGaggr, IK = −4.8 ± 0.2 kcal/mol (see also Materials and Methods and Supplementary Materials). These values are comparable to those reported for the conceptually related zirconaziridine [Cp2Zr(η2-CH2N(C18H37)2)]+[B(C6F5)4] having long N-alkyl substituents (Kaggr = 15,589 ± 595 M−1; ΔGaggr = −6.0 ± 0.1 kcal/mol in cyclohexane-d12, 298 K) [43], and are consistent with the formation of outer sphere ion pairs [4,43,57].

3.2. Polymerization Results

The cocatalytic properties of s-AlHAl were tested in ethene/1-hexene copolymerization at 100 °C in a high-boiling alkane solvent by using rac-Me2Si(2,6-dimethyl-4-phenyl-1-indenyl)2ZrCl2 (Figure 3, Cat-Zr) [58] as a representative ansa-zirconocene dichloride precatalyst (see Materials and Methods and Supporting Information for further details).
A well-established hydrocarbon-soluble activator, namely [HMeN(C16H33)2]+[B(C6F5)4] (RIBS), was used as a benchmark [36,37]. For catalytic activity RIBS alone is not sufficient because it can only serve as alkyl abstractor; therefore, TIBAL was added to the reaction system, thus acting as alkylating agent and scavenger [4,10]. s-AlHAl in principle does not require any additive [27], and it was therefore tested with and without added TIBAL. All polymerization experiments in the presence of TIBAL were performed by using [B]/[Zr] and [Al]/[Zr] ratios of 5 and 5.0 × 103, respectively, which is fairly typical [26].
The results obtained under optimized conditions for each cocatalyst formulation are summarized in Table 1. The productivity obtained with s-AlHAl/TIBAL was appreciably higher than that with RIBS/TIBAL (Table 1, entries 1–2 vs. 3–5). Moreover, s-AlHAl performed well as a standalone cocatalyst, provided that its concentration was high enough to scavenge the reaction system. Optimum catalyst productivities without adding TIBAL were achieved at [s-AlHAl]/[Zr]~50 (Table 1, entries 6–8), instead of ~5 with TIBAL. In all cases, these Al levels are much lower than when using RIBS/TIBAL ([Al]/[Zr]~5 × 103; Table 1, entries 6–8 vs. 1–2), and even more so compared with MAO (for which ([Al]/[Zr] up to 104–105 may be necessary for optimum performance [9,53]). No significant effects of activator nature and level on average copolymer molar mass and comonomer incorporation were observed.

4. Conclusions

In this paper, we introduced a novel molecular activator of the AlHAl family bearing long alkyl fragments on the aniline donor so as to enhance solubility in aliphatic hydrocarbons. This s-AlHAl species proved to be a competent cocatalyst in ethene/1-hexene copolymerizations in alkane diluent at high temperature. In combination with a representative ansa-zirconocene, catalyst productivities comparable to those with the better-known RIBS/TIBAL were obtained, with the advantage of working with a single-component activator at much lower [Al]/[Zr] ratios. These results represent a first demonstration that the AlHAl cocatalyst frame can be structurally amplified and tailored for specific applications; further variants are being prepared and screened in our laboratories, as will be reported in due course.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym15061378/s1, Figure S1: 1H NMR spectrum of [(p-C16H33)PhN(Me)2H]+[B(C6F5)4]; Figure S2: 1H and 13C NMR spectra of s-Al; Figure S3: 1H and 13C NMR spectra of s-AlHAl; Figure S4: 1H NOESY spectrum of s-AlHAl; Figure S5: DOSY NMR maps of s-AlHAl in pentane at different concentrations; Figure S6: Ln(I/I0) vs. G2 plots relative to the DOSY maps in Figure S5; Table S1: DOSY NMR measured self-diffusion coefficient (Dt), and estimated hydrodynamic radius (rH), hydrodynamic volume (VH) and average aggregation number (N) for TIBAL, DMAC16 and s-AlHAl at different concentrations in pentane at 298 K; Determination of Keq.

Author Contributions

Conceptualization, F.Z., C.Z. and R.C.; Methodology, P.H.M.B.; Investigation, G.U., A.V. and C.E.; Data curation, G.U., P.H.M.B., A.V. and C.E.; Writing—original draft, G.U., F.Z., C.Z. and R.C.; Writing—review & editing, P.H.M.B., A.V., C.E., A.M. and V.B.; Supervision, F.Z., C.Z. and R.C.; Funding acquisition and Project Administration, A.M. and V.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research forms part of the research program of DPI (project 857). This work was financially supported by the Dutch Polymer Institute (Project #857), PON—Ricerca e Innovazione (DM 1062), the University of Perugia and MUR (AMIS, “Dipartimenti di Eccellenza 2018−2022” program).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the Supplementary Materials.

Acknowledgments

The authors wish to thank the group of A. Z. Voskoboynikov (Moskow State University, Russian Federation) for the synthesis of the zirconocene precatalyst.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stürzel, M.; Mihan, S.; Mülhaupt, R. From Multisite Polymerization Catalysis to Sustainable Materials and All-Polyolefin Composites. Chem. Rev. 2016, 116, 1398–1433. [Google Scholar] [CrossRef]
  2. 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] [PubMed]
  3. 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]
  4. Zaccaria, F.; Sian, L.; Zuccaccia, C.; Macchioni, A. Ion Pairing in Transition Metal Catalyzed Olefin Polymerization. In Advances in Organometallic Chemistry; Perez, P.J., Ed.; Elsevier: Amsterdam, The Netherlands, 2020; Volume 73, pp. 1–78. [Google Scholar]
  5. Wild, F.R.W.P.; Zsolnai, L.; Huttner, G.; Brintzinger, H.H. Ansa-Metallocene Derivatives. IV. Synthesis and Molecular Structures of Chiral Ansa-Titanocene Derivatives with Bridged Tetrahydroindenyl Ligands. J. Organomet. Chem. 1982, 232, 233–247. [Google Scholar] [CrossRef]
  6. Ewen, J.A. Mechanisms of Stereochemical Control in Propylene Polymerizations with Soluble Group 4B Metallocene/Methylalumoxane Catalysts. J. Am. Chem. Soc. 1984, 106, 6355–6364. [Google Scholar] [CrossRef]
  7. Stevens, J.C.; Timmers, F.J.; Wilson, D.R.; Schmidt, G.F.; Nickias, P.N.; Rosen, R.K.; Knight, G.W.; Lai, S.-Y. Constrained Geometry Addition Polymerization Catalysts, Processes for Their Preparation, Precursors Therefor, Methods of Use, and Novel Polymers Formed Therewith. Eur. Patent Application No. 0,416,815 A2, 13 March 1991. [Google Scholar]
  8. Canich, J.A.M.; Licciardi, G.F. Mono-Cp Heteroatom Containing Group IVB Transition Metal Complexes with MAO: Supported Catalyst for Olefin Polymerization. U.S. Patent 5,057,475, 15 October 1991. [Google Scholar]
  9. Kaminsky, W. Discovery of Methylaluminoxane as Cocatalyst for Olefin Polymerization. Macromolecules 2012, 45, 3289–3297. [Google Scholar] [CrossRef]
  10. 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]
  11. Andresen, A.; Cordes, H.-G.; Herwig, J.; Kaminsky, W.; Merck, A.; Mottweiler, R.; Pein, J.; Sinn, H.; Vollmer, H.-J. Halogen-Free Soluble Ziegler Catalysts for the Polymerization of Ethylene. Control of Molecular Weight by Choice of Temperature. Angew. Chem. Int. Ed. Eng. 1976, 15, 630–632. [Google Scholar] [CrossRef]
  12. Yang, X.; Stern, C.L.; Marks, T.J. “Cation-like” Homogeneous Olefin Polymerization Catalysts Based upon Zirconocene Alkyls and Tris(Pentafluorophenyl)Borane. J. Am. Chem. Soc. 1991, 113, 3623–3625. [Google Scholar] [CrossRef]
  13. Ewen, J.A.; Elder, M.J. Metallocene Catalysts with Lewis Acids and Aluminum Alkyls. Eur. Patent Application No. 0,427,697, 15 May 1991. [Google Scholar]
  14. Gunther, S.O.; Lai, Q.; Senecal, T.; Huacuja, R.; Bremer, S.; Pearson, D.M.; DeMott, J.C.; Bhuvanesh, N.; Ozerov, O.V.; Klosin, J. Highly Efficient Carborane-Based Activators for Molecular Olefin Polymerization Catalysts. ACS Catal. 2021, 11, 3335–3342. [Google Scholar] [CrossRef]
  15. Nakashima, T.; Nakayama, Y.; Shiono, T.; Tanaka, R. Neutral, Noncoordinating, and Hydrocarbon-Soluble Protic Cocatalyst for Olefin Polymerization. ACS Catal. 2021, 11, 865–870. [Google Scholar] [CrossRef]
  16. Gao, Y.; Chen, J.; Wang, Y.; Pickens, D.B.; Motta, A.; Wang, Q.J.; Chung, Y.W.; Lohr, T.L.; Marks, T.J. Highly Branched Polyethylene Oligomers via Group IV-Catalysed Polymerization in Very Nonpolar Media. Nat. Catal. 2019, 2, 236–242. [Google Scholar] [CrossRef]
  17. Langford, D.; Göttker-Schnetmann, I.; Wimmer, F.P.; Casper, L.A.; Kenyon, P.; Winter, R.F.; Mecking, S. Tetrakis[3,5-Bis(Pentafluorosulfanyl)Phenyl]Borate: A Weakly Coordinating Anion Probed in Polymerization Catalysis. Organometallics 2019, 38, 2710–2713. [Google Scholar] [CrossRef]
  18. Tanaka, R. Precise Control of Coordination Polymerization via the Modification of Methylaluminoxane (MAO). Polym. J. 2020, 52, 661–670. [Google Scholar] [CrossRef]
  19. Faler, C.; Whalley, M.; Jiang, P.; Hagadorn, J.; Atienza, C.C.; Carpenter, A.; Rodriguez, G. Non-Coordinating Anion Type Activators Containing Cation Having Large Alkyl Groups. U.S. Patent Application No. US20190330139, 31 October 2019. [Google Scholar]
  20. Marks, T.J.; Gao, Y.; Lohr, T.L.; Christianson, M.D.; Klosin, J.; Carnahan, E.M.; Young, A.J. Highly Soluble Alkyl Substituted Carbenium Borate as Co-Catalysts for Olefin Polymerizations. Patent WO2,019,191,539, 3 May 2019. [Google Scholar]
  21. Keaton, R.J.; Ozerov, O.V.; Lai, Q.; Klosin, J.; Lesniak, J.M.; Pearson, D.M.; Senecal, T.D.; Woodward, W.H.H.; Mukhopadhyay, S.; Stutzman, J.R.; et al. Alkene Functionalized Activators. Patent WO2,020,223,488, 30 April 2020. [Google Scholar]
  22. Senecal, T.D.; Mukhopadhyay, S.; Keaton, R.J.; Klosin, J.; Huacuja, R.; Pearson, D.M.; Woodward, W.H.H. Anionic Group III Complexes as Weakly Coordinating Anions for Olefin Polymerization Catalyst Activators. Patent WO2,020,198,196, 24 March 2020. [Google Scholar]
  23. Lee, H.-J.; Baek, J.-W.; Seo, Y.-H.; Lee, H.-C.; Jeong, S.-M.; Lee, J.; Lee, C.-G.; Lee, B.-Y. Preparation of High-Purity Ammonium Tetrakis(Pentafluorophenyl)Borate for the Activation of Olefin Polymerization Catalysts. Molecules 2021, 26, 2827. [Google Scholar] [CrossRef]
  24. Gao, Y.; Christianson, M.D.; Wang, Y.; Coons, M.P.; Chen, J.; Zhang, J.; Marshall, S.; Lohr, T.L.; Klosin, J.; Marks, T.J. Alkane-Soluble Bis[Tris(Alkylphenyl)Carbenium] Diborate Cocatalyst for Olefin Polymerizations. ACS Catal. 2022, 12, 7589–7597. [Google Scholar] [CrossRef]
  25. Kitphaitun, S.; Fujimoto, T.; Ochi, Y.; Nomura, K. Effect of Borate Cocatalysts toward Activity and Comonomer Incorporation in Ethylene Copolymerization by Half-Titanocene Catalysts in Methylcyclohexane. ACS Org. Inorg. Au 2022, 2, 386–391. [Google Scholar] [CrossRef]
  26. Ehm, C.; Mingione, A.; Vittoria, A.; Zaccaria, F.; Cipullo, R.; Busico, V. High-Throughput Experimentation in Olefin Polymerization Catalysis: Facing the Challenges of Miniaturization. Ind. Eng. Chem. Res. 2020, 59, 13940–13947. [Google Scholar] [CrossRef]
  27. 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]
  28. Harlan, C.J.; Barron, A.R.; Bott, S.G. Three-Coordinate Aluminum Is Not a Prerequisite for Catalytic Activity in the Zirconocene—Alumoxane Polymerization of Ethylene. J. Am. Chem. Soc. 1995, 117, 6465–6474. [Google Scholar] [CrossRef]
  29. Joshi, A.; Collins, S.; Linnolahti, M.; Zijlstra, H.S.; Liles, E.; McIndoe, J.S. Spectroscopic Studies of Synthetic Methylaluminoxane: Structure of Methylaluminoxane Activators. Chem. Eur. J. 2021, 27, 8753–8763. [Google Scholar] [CrossRef]
  30. Collins, S.; Hasan, G.; Joshi, A.; McIndoe, J.S.; Linnolahti, M. Are Methylaluminoxane Activators Sheets? ChemPhysChem 2021, 22, 1326–1335. [Google Scholar] [CrossRef] [PubMed]
  31. Zaccaria, F.; Budzelaar, P.H.M.; Cipullo, R.; Zuccaccia, C.; Macchioni, A.; Busico, V.; Ehm, C. Reactivity Trends of Lewis Acidic Sites in Methylaluminoxane and Some of Its Modifications. Inorg. Chem. 2020, 59, 5751–5759. [Google Scholar] [CrossRef] [PubMed]
  32. Zaccaria, F.; Zuccaccia, C.; Cipullo, R.; Budzelaar, P.H.M.; Macchioni, A.; Busico, V.; Ehm, C. On the Nature of the Lewis Acidic Sites in “TMA-Free” Phenol-Modified Methylaluminoxane. Eur. J. Inorg. Chem. 2020, 2020, 1088–1095. [Google Scholar] [CrossRef]
  33. Zurek, E.; Ziegler, T. Theoretical Studies of the Structure and Function of MAO (Methylaluminoxane). Prog. Polym. Sci. 2004, 29, 107–148. [Google Scholar] [CrossRef]
  34. 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]
  35. 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]
  36. Turner, H.W. Catalysts, Method of Preparing These Catalysts and Method of Using Said Catalysts. Patent EP 0277004, 27 January 1988. [Google Scholar]
  37. Hlatky, G.G.; Upton, D.J.; Turner, H.W. Supported Ionic Metallocene Catalysts for Olefin Polymerization. WO9,109,882, 11 July 1991. [Google Scholar]
  38. Sian, L.; Dall’Anese, A.; Macchioni, A.; Tensi, L.; Busico, V.; Cipullo, R.; Goryunov, G.P.; Uborsky, D.; Voskoboynikov, A.Z.; Ehm, C.; et al. Role of Solvent Coordination on the Structure and Dynamics of Ansa-Zirconocenium Ion Pairs in Aromatic Hydrocarbons. Organometallics 2022, 41, 547–560. [Google Scholar] [CrossRef]
  39. Ehm, C.; Vittoria, A.; Goryunov, G.P.; Izmer, V.V.; Kononovich, D.S.; Samsonov, O.V.; Di Girolamo, R.; Budzelaar, P.H.M.; Voskoboynikov, A.Z.; Busico, V.; et al. An Integrated High Throughput Experimentation/Predictive QSAR Modeling Approach to Ansa-Zirconocene Catalysts for Isotactic Polypropylene. Polymers 2020, 12, 1005. [Google Scholar] [CrossRef] [PubMed]
  40. Mills, R. Self-Diffusion in Normal and Heavy Water in the Range 1–45°. J. Phys. Chem. 1973, 77, 685–688. [Google Scholar] [CrossRef]
  41. Zuccaccia, D.; Macchioni, A. An Accurate Methodology to Identify the Level of Aggregation in Solution by PGSE NMR Measurements: The Case of Half-Sandwich Diamino Ruthenium(II) Salts. Organometallics 2005, 24, 3476–3486. [Google Scholar] [CrossRef]
  42. Macchioni, A.; Ciancaleoni, G.; Zuccaccia, C.; Zuccaccia, D. Determining Accurate Molecular Sizes in Solution through NMR Diffusion Spectroscopy. Chem. Soc. Rev. 2008, 37, 479–489. [Google Scholar] [CrossRef]
  43. Rocchigiani, L.; Bellachioma, G.; Ciancaleoni, G.; MacChioni, A.; Zuccaccia, D.; Zuccaccia, C. Synthesis, Characterization, Interionic Structure, and Self-Aggregation Tendency of Zirconaaziridinium Salts Bearing Long Alkyl Chains. Organometallics 2011, 30, 100–114. [Google Scholar] [CrossRef]
  44. Rocchigiani, L.; Bellachioma, G.; Ciancaleoni, G.; Crocchianti, S.; Laganà, A.; Zuccaccia, C.; Zuccaccia, D.; Macchioni, A. Anion-Dependent Tendency of Di-Long-Chain Quaternary Ammonium Salts to Form Ion Quadruples and Higher Aggregates in Benzene. ChemPhysChem 2010, 11, 3243–3254. [Google Scholar] [CrossRef] [PubMed]
  45. Bellachioma, G.; Ciancaleoni, G.; Zuccaccia, C.; Zuccaccia, D.; Macchioni, A. NMR Investigation of Non-Covalent Aggregation of Coordination Compounds Ranging from Dimers and Ion Pairs up to Nano-Aggregates. Coord. Chem. Rev. 2008, 252, 2224–2238. [Google Scholar] [CrossRef]
  46. Martin, R.B. Comparisons of Indefinite Self-Association Models. Chem. Rev. 1996, 96, 3043–3064. [Google Scholar] [CrossRef] [PubMed]
  47. Casey, C.P.; Carpenetti, D.W. Measurement of Barriers for Alkene Dissociation and for Inversion at Zirconium in a d0 Zirconium-Alkyl-Alkene Complex. Organometallics 2000, 19, 3970–3977. [Google Scholar] [CrossRef]
  48. Busico, V.; Cipullo, R.; Mingione, A.; Rongo, L. Accelerating the Research Approach to Ziegler-Natta Catalysts. Ind. Eng. Chem. Res. 2016, 55, 2686–2695. [Google Scholar] [CrossRef]
  49. Vittoria, A.; Busico, V.; Cannavacciuolo, F.D.; Cipullo, R. Molecular Kinetic Study of “Chain Shuttling” Olefin Copolymerization. ACS Catal. 2018, 8, 5051–5061. [Google Scholar] [CrossRef]
  50. Uborsky, D.V.; Mladentsev, D.Y.; Guzeev, B.A.; Borisov, I.S.; Vittoria, A.; Ehm, C.; Cipullo, R.; Hendriksen, C.; Friederichs, N.; Busico, V.; et al. C1-Symmetric Si-Bridged (2-Indenyl)(1-Indenyl) Ansa-Metallocenes as Efficient Ethene/1-Hexene Copolymerization Catalysts. Dalton Trans. 2020, 49, 3015–3025. [Google Scholar] [CrossRef]
  51. Ehm, C.; Vittoria, A.; Goryunov, G.P.; Izmer, V.V.; Kononovich, D.S.; Samsonov, O.V.; Budzelaar, P.H.M.; Voskoboynikov, A.Z.; Busico, V.; Uborsky, D.V.; et al. On the Limits of Tuning Comonomer Affinity of ‘Spaleck-Type’ Ansa-Zirconocenes in Ethene/1-Hexene Copolymerization: A High-Throughput Experimentation/QSAR Approach. Dalton Trans. 2020, 49, 10162–10172. [Google Scholar] [CrossRef]
  52. Wrobel, O.; Schaper, F.; Brintzinger, H.H. Bulky Siloxyaluminum Alkyls as Models for Al2Me6-Treated Silica Gel Surfaces. Characterization of a Dimethylaniline-Stabilized Dimethylaluminum Cation. Organometallics 2004, 23, 900–905. [Google Scholar] [CrossRef]
  53. Zijlstra, H.S.; Harder, S. Methylalumoxane—History, Production, Properties, and Applications. Eur. J. Inorg. Chem. 2015, 2015, 19–43. [Google Scholar] [CrossRef]
  54. Zaccaria, F.; Zuccaccia, C.; Cipullo, R.; Macchioni, A. Extraction of Reliable Molecular Information from Diffusion NMR Spectroscopy: Hydrodynamic Volume or Molecular Mass? Chem. Eur. J. 2019, 25, 9930–9937. [Google Scholar] [CrossRef]
  55. Cohen, Y.; Avram, L.; Frish, L. Diffusion NMR Spectroscopy in Supramolecular and Combinatorial Chemistry: An Old Parameter—New Insights. Angew. Chem. Int. Ed. 2005, 44, 520–554. [Google Scholar] [CrossRef]
  56. Sian, L.; Macchioni, A.; Zuccaccia, C. Understanding the Role of Metallocenium Ion-Pair Aggregates on the Rate of Olefin Insertion into the Metal-Carbon Bond. ACS Catal. 2020, 10, 1591–1606. [Google Scholar] [CrossRef]
  57. Macchioni, A. Ion Pairing in Transition-Metal Organometallic Chemistry. Chem. Rev. 2005, 105, 2039–2073. [Google Scholar] [CrossRef]
  58. Kaminsky, W. The Discovery of Metallocene Catalysts and Their Present State of the Art. J. Polym. Sci. A Polym. Chem. 2004, 42, 3911–3921. [Google Scholar] [CrossRef]
Figure 1. Structures of AlHAl and its hydrocarbon soluble analogue s-AlHAl.
Figure 1. Structures of AlHAl and its hydrocarbon soluble analogue s-AlHAl.
Polymers 15 01378 g001
Scheme 1. Synthesis of s-Al and s-AlHAl. DIBAL-H is shown as a monomer for simplicity.
Scheme 1. Synthesis of s-Al and s-AlHAl. DIBAL-H is shown as a monomer for simplicity.
Polymers 15 01378 sch001
Figure 2. (a) Representative 1D diffusion NMR spectra of s-AlHAl at increasing gradient strength in pentane (with acetone-d6 coaxial capillary, 298 K, 33 mM) and (b) DOSY map derived therefrom; (c) trend of the average aggregation number (N) as a function of the analytical concentration of s-AlHAl derived from diffusion NMR measurements (trendline obtained by fitting experimental data based on the IK model; see ESI). *, solvent signals. A 15% and 10% error are assumed on aggregation number and concentration, respectively [43,44].
Figure 2. (a) Representative 1D diffusion NMR spectra of s-AlHAl at increasing gradient strength in pentane (with acetone-d6 coaxial capillary, 298 K, 33 mM) and (b) DOSY map derived therefrom; (c) trend of the average aggregation number (N) as a function of the analytical concentration of s-AlHAl derived from diffusion NMR measurements (trendline obtained by fitting experimental data based on the IK model; see ESI). *, solvent signals. A 15% and 10% error are assumed on aggregation number and concentration, respectively [43,44].
Polymers 15 01378 g002
Figure 3. The ansa-zirconocene precatalyst Cat-Zr employed in polymerization screening.
Figure 3. The ansa-zirconocene precatalyst Cat-Zr employed in polymerization screening.
Polymers 15 01378 g003
Table 1. Ethene/1-hexene copolymerization results with precatalyst Cat-Zr in combination with different cocatalysts (a).
Table 1. Ethene/1-hexene copolymerization results with precatalyst Cat-Zr in combination with different cocatalysts (a).
EntryActivator[B]/[Zr]TIBAL[Altot]/[Zr] R p R p ¯   M n M n ¯   M w M w ¯ PDI[H]
µmol kg μmolZr−1 h−1kg μmolZr−1 h−1kDa mol%
1RIBS5105.0 × 1030.190.22 (3)57561411392.57.4
20.25561372.57.2
3s-AlHAl5105.0 × 1030.570.53 (5)59611411382.47.9
40.48641402.27.6
50.55611342.27.7
65001.0 × 1020.420.50 (10)59551341342.37.9
70.48511342.67.9
80.61541342.58.0
(a) in ISOPAR-G (Vtot = 6 mL), pethene = 12 bar, [H]feed = 3.7 vol%, T = 100 °C, nZr = 2 nmol. Rp, productivity; Mn, number average molecular mass; Mw, weight average molecular mass; PDI, polydispersity index; [H], 1-hexene incorporation into the polymer (determined by NMR spectroscopy).
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

Urciuoli, G.; Zaccaria, F.; Zuccaccia, C.; Cipullo, R.; Budzelaar, P.H.M.; Vittoria, A.; Ehm, C.; Macchioni, A.; Busico, V. A Hydrocarbon Soluble, Molecular and “Complete” Al-Cocatalyst for High Temperature Olefin Polymerization. Polymers 2023, 15, 1378. https://doi.org/10.3390/polym15061378

AMA Style

Urciuoli G, Zaccaria F, Zuccaccia C, Cipullo R, Budzelaar PHM, Vittoria A, Ehm C, Macchioni A, Busico V. A Hydrocarbon Soluble, Molecular and “Complete” Al-Cocatalyst for High Temperature Olefin Polymerization. Polymers. 2023; 15(6):1378. https://doi.org/10.3390/polym15061378

Chicago/Turabian Style

Urciuoli, Gaia, Francesco Zaccaria, Cristiano Zuccaccia, Roberta Cipullo, Peter H. M. Budzelaar, Antonio Vittoria, Christian Ehm, Alceo Macchioni, and Vincenzo Busico. 2023. "A Hydrocarbon Soluble, Molecular and “Complete” Al-Cocatalyst for High Temperature Olefin Polymerization" Polymers 15, no. 6: 1378. https://doi.org/10.3390/polym15061378

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