BHT-modified MAO: cage size estimation, chemical counting of

Feb 20, 2019 - BHT-modified MAO: cage size estimation, chemical counting of strongly acidic Al-sites and activation of a Ti-phosphinimide precatalyst...
5 downloads 0 Views 659KB Size
Subscriber access provided by Macquarie University

Article

BHT-modified MAO: cage size estimation, chemical counting of strongly acidic Al-sites and activation of a Ti-phosphinimide precatalyst Francesco Zaccaria, Cristiano Zuccaccia, Roberta Cipullo, Peter H.M. Budzelaar, Alceo Macchioni, Vincenzo Busico, and Christian Ehm ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.9b00076 • Publication Date (Web): 20 Feb 2019 Downloaded from http://pubs.acs.org on February 23, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

BHT-modified MAO: cage size estimation, chemical counting of strongly acidic Al-sites and activation of a Ti-phosphinimide precatalyst Francesco Zaccaria,a,b Cristiano Zuccaccia,b,* Roberta Cipullo,a Peter H. M. Budzelaar,a Alceo Macchioni,b Vincenzo Busicoa and Christian Ehma,*. a. Dipartimento di Scienze Chimiche, Università di Napoli Federico II, Via Cintia, 80126 Napoli, Italy. b. Dipartimento di Chimica, Biologia e Biotecnologie and CIRCC, Università di Perugia, Via Elce di Sotto 8, 06123 Perugia, Italy

ABSTRACT. MAO/BHT (MAO = methylaluminoxane; BHT = 2,6-di-tert-butyl-4-methylphenol) cocatalyst for olefin polymerization has been investigated by NMR spectroscopy. It has been found that it consists of oligomeric [AlOMe0.9(bht)0.1]n cages and monomeric MeAl(bht)2. Diffusion NMR indicates an average n for Al-clusters of 62-96, i.e. 2-3 times higher than that estimated for unmodified MAO under analogous conditions (n ≈ 26-41). The reactivity of MAO/BHT has been explored by monitoring the activation of the Cp*-phosphinimide titanium dichloride precatalyst Cp*(tBu3P=N)TiCl2. Comparison with independently synthesized model species and DFT modeling allowed characterization of the reaction mixtures obtained at varying aluminum to titanium ratios. Homodinuclear adducts [Cp*(tBu3P=N)TiX]2(µ-Y)+ (X, Y = Me or Cl) forming ACS Paragon Plus Environment

1

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 51

Outer Sphere Ion Pairs (OSIPs) with MAO/BHT-derived anions are dominant at low Al/Ti ratios, whereas mononuclear Inner Sphere Ion Pairs (ISIPs) [Cp*(tBu3P=N)Ti-X]+[MAO/BHT]- are formed at high Al/Ti ratios; both types of species are found to be viable precursors for the cationic active species. Activation of dibenzyl analogue Cp*(tBu3P=N)TiBn2 results in the clean formation of [Cp*(tBu3P=N)Ti-Bn]+[MAO/BHT]- OSIP, giving sharp 1H and 31P NMR signals; this reaction was exploited to quantify the amount of strongly acidic sites on Al-clusters, shedding further light on the structure and properties of MAO/BHT.

KEYWORDS. Methylaluminoxane, Modified MAO, DOSY NMR, olefin polymerization catalysis, chemical shift predictions, activation, DFT, post-metallocene.

INTRODUCTION Methylaluminoxane (MAO) is the typical cocatalyst of choice in molecular olefin polymerization catalysis, due to its remarkable combination of alkylating, activating and impurity scavenging properties.1, 2 Additionally, it offers the possibility to be heterogenized by absorption on supports such as silica or alumina, expanding the application range of molecular catalysts from solution to slurry and gas phase reactors.3 It is generally accepted that MAO consists of a distribution of cagelike oligomers (AlOMe)n decorated by and in equilibrium with trimethylaluminum (TMA),4-14 although no definitive consensus has been reached on its structure, yet. The average number of Alatoms per cage (n)7, 8, 11-21 and the amount of strongly acidic Al-sites on MAO clusters8, 12, 22-25 are some of the most debated aspects. Often, the presence of TMA represents a major Achilles’ heel of typical MAO solutions. Though playing a positive role as methylating agent and impurity scavenger,9, 10, 26 ‘free' TMA can also

ACS Paragon Plus Environment

2

Page 3 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

decrease catalyst productivity and polymer molecular weight by forming rather stable heterodinuclear adducts M-(μ-X)m-Al with group IV cationic complexes (M = transition metal, X = Me or Cl, m = 1-2), suspected to be ‘dormant’ in olefin polymerization and involved in chain transfer to Al.27-35 The highly reactive TMA can also trigger catalyst deactivation.26, 36 A growing scientific and commercial interest is therefore directed at the development of different grades of Modified Methylaluminoxanes (MMAO), aiming at addressing some of its drawbacks, while preserving desirable MAO properties.1 Replacing methyl with bulkier iso-butyl or n-octyl groups,37-39 or following non-hydrolytic processes to ensure that virtually no residual TMA remains,40, 41 represent promising commercial approaches to different grades of MMAO. Alternative strategies have been proposed to modify typical commercial MAO solutions. For instance, TMA (b.p. 125 °C) can be physically removed under vacuum, although it has been debated if this removal is quantitative.20, 22, 42 An even simpler and highly effective method to trap TMA is to add a sterically hindered phenol (HP), such as 2,6-di-tert-butyl-4-methylphenol (BHT).28 TMA is thereby transformed into the far less reactive aryloxide complex MeAl(bht)2 (bht = BHT phenolate).43, 44 This chemical modification of MAO has been successfully applied in many polymerization studies,27-29, 31, 45, 46 including industrially relevant cases,47 but it has been rarely exploited for spectroscopic mechanistic investigations.20, 27 To the best of our knowledge, the sole example of an NMR study on a ternary system, catalyst/MAO/HP,20 focused on the tendency of Cp2ZrMe+ to form Inner (ISIPs) or Outer Sphere Ion Pairs (OSIPs)48, 49 with counterions and higher aggregates thereof. In the present paper, the structure and reactivity of MAO/BHT are explored in a more comprehensive manner. Detailed insights into its composition and Al-cage size have been obtained

ACS Paragon Plus Environment

3

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 51

by diffusion ordered NMR spectroscopy (DOSY); although analogous 1D pulsed gradient spinecho (PGSE) experiments have been widely applied in this7, 8, 20, 49, 50 and other51 fields, 2D DOSY is particularly suited to analyze complex reaction mixtures52 and it has been only occasionally applied to MAO-type cocatalysts.53 The reactivity of MAO/BHT has been explored by monitoring the activation of a representative phosphinimide precatalyst Cp*(tBu3P=N)TiCl2 (1-Cl2, Figure 1) at varying Al/Ti ratios, using the P atom in the ancillary ligand as spectroscopic handle. Independently synthesized model compounds using well-defined molecular activators and DFT modelling facilitated identification of activation products in complex reaction mixtures. The catalytic activity of the observed species in 1-hexene polymerization was also evaluated, providing a rare example of a systematic study of activation of an industrially relevant catalyst class54 by a TMA-free MMAO. Finally, a simple and effective method to estimate the average number of strongly acidic sites on the Al-clusters is proposed based on the dibenzyl precatalyst 1-Bn2. Tentative structural and functional conclusions on the oligomeric fraction of MAO/BHT combining all results are presented.

Ti N tBu3P

X X

1-X2

Figure 1. Studied phosphinimide half-titanocene precatalyst. X = Me, Cl or Bn.

ACS Paragon Plus Environment

4

Page 5 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

RESULTS AND DISCUSSION Preparation and properties of MAO/BHT MAO/BHT was prepared by adding BHT to a diluted commercial solution of MAO in toluene, as described in ref. 28. The resulting mixture was subsequently dried under vacuum to avoid large peaks of toluene in 1H NMR spectra, providing solid-(MAO/BHT) [s(MAO/BHT)] as a white powder with good solubility properties in aromatic solvents. The phenol reacts preferentially with TMA,8, 27-29 estimated to account for 30-33% of the total Al in the starting MAO solution. However, in the late stages of this process, 1H NMR revealed that side reactions of small amounts of BHT with the (AlOMe)n cages occur, giving broad Al-phenolate signals (Figure S1).20 Relative integration of NMR peaks suggests an approximate formula of {[AlOMe0.9(bht)0.1]ˑ[MeAl(bht)2]0.3}n for s(MAO/BHT). Addition of a large excess of BHT does not affect the ratio between Al-Me and Al-bht groups on MAO-type clusters significantly,20 and does not kill the activation properties of the cocatalyst, suggesting that only the most sterically accessible Al-sites react with the phenol. s(MAO/BHT) was also analyzed by DOSY NMR. In typical 2D-DOSY plots, the chemical shift of the various species in solution is resolved according to their translational self-diffusion coefficients (Dt),52 as shown in Figure 2a for the case of s(MAO/BHT). Via the Stokes-Einstein equation, Dt values can be correlated to the hydrodynamic radius (rH) or volume (VH) of the analyte,55 providing information on molecular dimensions.56 Here, toluene solutions were analyzed at 298 K, using triphenylmethane (Ph3CH)57 as internal standard to account for temperature and viscosity fluctuations.56 Two components are found for s(MAO/BHT), having Dt equal to 7.03 and 2.37 ˑ 10-10 m2/s (Entries 2-3; Table 1), in line with previous reports.20 The larger

ACS Paragon Plus Environment

5

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 51

Dt value of the first component is comparable to that of isolated MeAl(bht)2, which was independently synthesized by reacting TMA with BHT (Al/BHT 1:2; Entry 6, Table 1), and it is therefore diagnostic of a free-diffusing bis(phenolate) complex. The second component, having a smaller Dt, is ascribed to [AlOMe0.9(bht)0.1]n cages; as the signal attenuation for Al-Me and Al-bht fragments on these cages exhibit a similar dependence on the gradient strength (Figure 2a), it appears that the phenolate groups are rather homogeneously distributed among the Al-clusters.

Figure 2. DOSY maps (toluene-d8, 298K) of s(MAO/BHT) (a) and MeAl(bht)2-depleted washed s(MAO/BHT) (b). [AlOMe0.9(bht)0.1] ≈ 30-35 mM. Ph3CH was used as internal viscosity standard (see main text). *Traces of residual MeAl(bht)2, estimated to account for 100 eq.) to 1X+[s(MAO/BHT)]- results in a 31P NMR spectrum (Figure 4e) analogous to the one obtained with TTB activator (Figure 3g), i.e. a major sharp signal at 54.0 ppm and other minor, broader peaks at 55.5-57.2 ppm. The dormant allyl complex 1-All+[s(MAO/BHT)]- (54.0 ppm) is likely formed via CTM, and it accumulates along with some other decomposition/inactive products (55.5-57.2 ppm, probably deriving from decomposition of in situ generated Ti-H species).68 The sharpness of the peak at 54.0 ppm indicates that also 1-All+ forms OSIP with MAO/BHT anions, as the η3-allyl coordination prevents a tight cation/anion interaction.

ACS Paragon Plus Environment

17

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 51

When the amount of 1-hexene used is lowered to only ~10 eq., 31P NMR evidence of the formation of 1-Bn+ via solvent rather than monomer activation was obtained (Figure S11). These observations are in line with those obtained with TTB activator,68 and provide direct NMR evidence for the formation of titanium allyl and benzyl type complexes under polymerization conditions also in the more complex case of s(MAO/BHT) cocatalyst. The rapid disappearance of starting 1-X+[s(MAO/BHT)]- broad peaks in the 31P NMR spectrum, indicates that at least the clear majority of ISIPs reacts with the monomer. These observations point to ISIPs dissociation being relatively facile, in contrast to what has been reported for metallocenes in combination with commercial MAO. In such cases, the presence of TMA leads to the preferential formation of OSIPs of bimetallic Ti-Al adducts and only a fraction of the transition metal cations giving a tight interaction with the MAO-counterions.7,

24, 34, 66

Due to the high

dissociation barriers predicted for the latter, ISIPs have been proposed as plausible dormant sites.23, 34, 78, 80

Our somewhat different observations indicate that a close binding with MAO-type anions

does not necessarily prevent monomer coordination to the cationic active species, at least for the present catalyst class. Chemical estimation of strongly Lewis acidic Al-sites The Al/Ti ratio of ~280 necessary for full activation of 1-Cl2 indicates that only a rather small fraction of Al atoms on s(MAO/BHT) actually reacts with the precatalyst, as it is generally accepted for MAO-activators.6,

12, 23

In principle, this type of reaction monitoring provides an

estimate of the stoichiometry of the activation process in terms of total Al vs Ti atoms, from which the percentage of reactive Al-center can be derived. However, this approach faces some practical hurdles, such as: A) possible side reactions like dimer formation can compete with catalyst

ACS Paragon Plus Environment

18

Page 19 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

activation; B) the difficulties in accurately quantifying the species in solution; C) concentrationdependent Al-cage reorganizations that might be induced upon sequential additions of cocatalyst to the reaction mixture.20 The double alkylating/abstracting role carried out by MAO in activating dichloride complexes adds further complexity. Based on the above described results concerning 1-X2 complexes, it was envisaged that quantitative monitoring of the activation of 1-Bn2 might allow to circumvent all these complications, and provide a reliable chemical estimate of the sole abstracting sites on s(MAO/BHT) cages.81 1-Bn2 is an already ‘alkylated’ species that, upon activation, does not undergo dimer formation but only yields monomeric 1-Bn+ (Figure 5). Furthermore, the Ti-Bn+ cation forms an OSIP with s(MAO/BHT) counterions, giving a sharp

31P

NMR signal and

diagnostic 1H NMR peaks (e.g. the ortho-H of the Ti-benzyl group at δH = 7.18 ppm) appearing in relatively clean regions of the spectra, making accurate integration (i.e. quantification) of this species particularly easy. Finally, experiments were designed to keep the concentration of s(MAO/BHT) relatively low ([AlOMe0.9(bht)0.1] ≈ 24 mM) and constant throughout the reaction, to prevent gelification/precipitation and avoid concentration dependent self-aggregation of Alcages. The latter should be rather negligible for phenol-modified MAO, especially at such low concentrations (see Experimental Section for further details), as previously reported.20

ACS Paragon Plus Environment

19

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Ti N

Bn

Page 20 of 51

A-

s(MAO/BHT) Ti

Bn

N tBu3 P

tBu3 P 1-Bn2

1-Bn+[s(MAO/BHT)]A-

= [s(MAO/BHT)]-

Figure 5. Exemplifying 31P NMR (toluene-d8, 298 K) spectra of the activation of 1-Bn2 after sequential additions of s(MAO/BHT), highlighting the clean formation of an OSIP of 1-Bn+ and MAO/BHT counterion as the only reaction product. The build-up of 1-Bn+ was monitored by NMR up to a constant concentration value; most of the product is formed already in the first 5-10 min (Figure 6). The initial concentration of Al-Me groups on s(MAO/BHT) cages and the final concentration of cationic Ti-benzyl complex were

ACS Paragon Plus Environment

20

Page 21 of 51

measured with respect to an external standard. In this way, it was found that 1 out of ~42 Al-sites on [AlOMe0.9(bht)0.1]n clusters is capable to abstract a benzyl group from the Ti-center.82 Analogous experiments were carried out on MeAl(bht)2-depleted s(MAO/BHT), prepared by extracting the bis(aryloxide) complex with pentane. Very similar results were obtained (1 of ~45 Al-sites), further indicating that removal of MeAl(bht)2 has only marginal effects on the oligomeric component of s(MAO/BHT). With respect to established methods using organic Lewis bases as molecular probes (e.g. TEMPO),19 the proposed procedure offers the advantage of exploiting a real abstraction reaction for the quantification of strongly acidic sites, although it cannot not contribute to the elucidation of the abstraction mechanism (e.g. involving sterically strained Al atoms on MAO cages4, 16, 83 or transient [Me2Al]+ cations17, 63, 84).

0.7 0.6 0.5

[1-Bn+] (mM)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

0.4 0.3 0.2 0.1 0.0 0

20

40

60

t (min)

80

100

Figure 6 Concentration of 1-Bn+ as a function of 1-Bn2 + s(MAO/BHT) reaction time. Determined with respect to an external standard by using 1H NMR (toluene-d8, 298 K). 15% error assumed on concentrations.

ACS Paragon Plus Environment

21

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Assuming one ‘working’ acidic site per cage,6,

19, 23

Page 22 of 51

these results suggest that the average

[AlOMe0.9(bht)0.1]n cluster should contain n ≈ 42-45 Al atoms. This value is somewhat smaller than that estimated by diffusion NMR (n ≈ 62-96), but - considering all the approximations involved - the agreement is quite satisfactory. Alternatively, one might speculate that the average number of acidic sites per cage is actually >1; examples of dianionic MAO cages have been occasionally reported, for instance by mass spectrometry.11 Although this second explanation appears plausible,39 these results do not allow to draw definitive conclusions in this respect.

CONCLUSIONS NMR spectroscopy has been used to investigate the structure and reactivity of TMA-free solid(MAO/BHT) cocatalyst. First, some structural features have been identified by diffusion NMR, such as A) an approximate formula of {[AlOMe0.9(bht)0.1]ˑ[MeAl(bht)2]0.3}n, B) the free-diffusing nature of the MeAl(bht)2 component and C) an average cage size of 62-96 metal atoms. Then, the activation properties of s(MAO/BHT) have been tested by reacting it with the phosphinimide half-titanocene 1-Cl2. With the help of ad-hoc synthesized model species and DFT calculations, it has been found that at low Al/Ti ratios, rather stable Cl-bridged and partially methylated Ti-dimers are formed, while no indications for Ti/Al heterodinuclear adducts are observed. The dimers easily dissociate in the presence of 1-hexene, releasing a cationic alkyl complex that initiates chain growth. Addition of monomer seems to have a beneficial effect on the activation of the remaining chlorinated species, likely because it induces dissociation of the dimers, making the neutral precursors more prone to react with s(MAO/BHT). At Al/Ti ratios around 280, the abstraction process is complete, and the dimers are fully converted to ISIPs with s(MAO/BHT)

ACS Paragon Plus Environment

22

Page 23 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

anions. Also in this case, addition of 1-hexene rapidly leads to dissociation of the ion pair, showing that close binding of MAO-type counterions to the transition metal cation is not necessarily a significant hurdle for monomer capture. The activated species 1-X+ containing MAO-type anions exhibit a similar reactivity to that recently reported for analogous ion pairs with [B(C6F5)4]counterions; homolytic solvent and monomer activation are observed when reacting 1-X+ with suitable amounts of 1-hexene. The activation process of 1-Cl2 has been compared to that of analogous 1-Me2 and 1-Bn2. In particular, accurate monitoring of the activation of the dibenzyl complex was exploited to derive another important structural property of s(MAO/BHT), namely the fraction of strongly acidic Alcenters (1 of ~45). Under the assumption of only one strongly acidic site per cage, the method also provides an estimate for the average cage size that is not too different from the one based on diffusion NMR (n ≈ 62-96). However, it cannot be excluded that the discrepancy between these two estimations is actually due to the presence of slightly more than one reactive center on MAO/BHT clusters, on average. The higher amount of s(MAO/BHT) necessary to activate 1-Cl2 rather than 1-Bn2 is likely due to the double alkylating/abstracting role of the cocatalyst, as well as to competing dimer formation. In any case, the Al/Ti ratios used here are significantly lower than those typically used in bulk polymerization experiments to achieve optimal productivities (in the order of 103-104),1,

85

suggesting that a large part of this Al excess serves different purposes like impurity scavenging. This is in line with analogous observations on MAO17, 18, 86 and dMAO.22

EXPERIMENTAL PART

ACS Paragon Plus Environment

23

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 51

Synthesis. 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 (36h. 1-MeBn. 1H NMR (400 MHz, toluene-d8): 1.91 (s, 15H, Cp- CH3), 2.80 (d, 1H, 2J = 10.6 Hz, TiCH2-Ph), 1.84 (d, 1H, Ti-CH2-Ph), 1.25 (d, 27H, JPH = 12.6 Hz, C(CH3)3), 0.53 (s, 3H, Ti-Me) ppm.

13C

NMR (100 MHz, toluene-d8): 118.8 (C5Me5), 68.3 (Ti-CH2-Ph), 48.5 (Ti-Me), 41.5

(C(CH3)3), 29.8 (C(CH3)3), 12.0 (Cp-CH3) ppm. 31P NMR (161 MHz, toluene-d8): 33.0 ppm. Diffusion NMR experiments. DOSY NMR measurements were performed at 298K in toluene-d8 on a Bruker Avance III HD 400 spectrometer equipped with a smart probe with a z gradient coil, by using the standard double-stimulated echo pulse sequence without spinning. The shape of the gradients was rectangular, their duration (δ) was 4 ms, and their strength (G) was varied during the experiments. All the spectra were acquired using 32K points, 256 increments, 128 scans, a spectral width of 6400 Hz, an acquisition time of 0.5 s and a relaxation delay of 1 s per transient. The DOSY spectra were processed by means of the Bruker Dynamics Center software package (version 2.5.5), using a line broadening of 1.0 Hz in the direct dimension. The DOSY maps were obtained by using the Inverse Laplace Transform routine and choosing 256 points in the vertical dimension. Since the aliphatic resonance of toluene-d8 overlaps with that MAO/BHT, selfdiffusion coefficients (Dt) were calculated with respect to that of Ph3CH, which was used as internal reference standard.58 The Dt of Ph3CH in toluene-d8 was separately calibrated using the diffusion coefficient of residual solvent resonance in a dilute solution (0.1 mM). The latter was calibrated using an external sample of HDO in D2O88 under the same exact conditions. Dt data were treated as described in the literature in order to derive the hydrodynamic dimensions.56

ACS Paragon Plus Environment

27

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 51

Low concentrations of oligomeric Al-Me groups have been used (30-35 mM), which represent a good compromise between those typically used in catalysis1, 85 and the often higher ones used in NMR studies to guarantee a good signal to noise ratio.7,

8, 20

Furthermore, they should be

sufficiently low to prevent any gelification/precipitation and, more generally, minimize cage selfaggregation phenomena.20, 89 DFT Calculations. Following the protocol described in ref. 90, all geometries were fully optimized by using the Gaussian 09 software package91 in combination with the OPTIMIZE routine of Baker92-93 and the BOpt software package.94 All relevant minima and transition states were fully optimized at the TPSSTPSS level95 of theory employing correlation-consistent polarized valence double-ζ Dunning (DZ) basis sets (cc-pVDZ quality)96, 97 from the EMSL basis set exchange library.98 The density fitting approximation was used at the optimization stage (Resolution of Identity, RI).99 All calculations were performed at the standard Gaussian 09 quality settings [Scf=Tight and Int(Grid=Fine)]. All structures represent either true minima (as indicated by the absence of imaginary frequencies) or transition states (with exactly one imaginary frequency corresponding to the reaction coordinate). Final single-point energies were calculated at the M062X level of theory99 employing triple-ζ Dunning (TZ) basis sets (cc-pVTZ quality).96 Although M06-2X is usually recommended for main group thermochemistry, it has been found that it accurately reproduces different experimental problems concerning d0 and d1 early transition metal systems in olefin polymerization32, 45, 68, 90, 101, 102 and hydrodefluorination.103 Aluminum chemistry is similarly well described,13,

35, 104

including MAO related problems.13 Solvent effects were

included with the polarizable continuum model approach (PCM; solvent=toluene, ε=2.3741) at this stage.105 Enthalpies and Gibbs free energies were then obtained from TZ single-point energies and thermal corrections from the TPSSTPSS/cc-pVDZ-(PP) vibrational analyses; entropy

ACS Paragon Plus Environment

28

Page 29 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

corrections were scaled by a factor of 0.67 to account for decreased entropy in the condensed phase.100, 106 The GIAO method107 at the TPSSTPSS level in combination with basis sets optimized for this method (aug-cc-pVTZ-J basis set108 for Ti and IGLO-III109-110 for all other atoms) and PCM was used for 31P NMR chemical shift predictions. AUTHOR INFORMATION Corresponding Author *E-mail for C.Z.: [email protected]. *E-mail for C.E.: [email protected]. Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. ASSOCIATED CONTENT Supporting Information. The following files are available free of charge. Additional NMR and DFT details (PDF) Coordinates for optimized geometries (XYZ) ACKNOWLEDGMENTS Part of this work has been financially supported by PRIN 2015 (20154X9ATP_004), University of Perugia and MIUR (AMIS, “Dipartimenti di Eccellenza - 2018-2022” program). F.Z. thanks INSTM and CIRCC for a post-doc grant. ABBREVIATIONS

ACS Paragon Plus Environment

29

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 51

MAO, methylaluminoxane; MMAO, modified methylaluminoxane; TMA, trimethylaluminum; BHT, 2,6-di-tert-butyl-4-methylphenol; HP, hindered phenol; ISIP, inner sphere ion pair; OSIP, other sphere ion pair; DOSY, diffusion ordered NMR spectroscopy; PGSE, pulsed gradient spin echo; DFT, density functional theory; TTB, trityl borate; CTM, chain transfer to monomer; CTS, chain transfer to solvent; TEMPO, 2,2,6,6-Tetramethylpiperidine 1-oxyl. REFERENCES 1.

Zijlstra, H. S.; Harder, S., Methylalumoxane – History, Production, Properties, and

Applications. Eur. J. Inorg. Chem. 2015, 2015, 19-43. 2.

Kaminsky, W., Discovery of Methylaluminoxane as Cocatalyst for Olefin Polymerization.

Macromolecules 2012, 45, 3289-3297. 3.

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. 4.

Zurek, E.; Ziegler, T., Theoretical Studies of The Structure and Function of MAO

(Methylaluminoxane). Prog. Polym. Sci. 2004, 29, 107-148. 5.

Tritto, I.; Méalares, C.; Sacchi, M. C.; Locatelli, P., Methylaluminoxane: NMR Analysis,

Cryoscopic Measurements and Cocatalytic Ability in Ethylene Polymerization. Macromol. Chem. Phys. 1997, 198, 3963-3977. 6.

Falls, Z.; Tymińska, N.; Zurek, E., The Dynamic Equilibrium between (AlOMe)n Cages

and (AlOMe)n·(AlMe3)m Nanotubes in Methylaluminoxane (MAO): A First-Principles Investigation. Macromolecules 2014, 47, 8556-8569.

ACS Paragon Plus Environment

30

Page 31 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

7.

Babushkin, D. E.; Brintzinger, H.-H., Activation of Dimethyl Zirconocene by

Methylaluminoxane (MAO) Size Estimate for Me-MAO- Anions by Pulsed Field-Gradient NMR. J. Am. Chem. Soc. 2002, 124, 12869-12873. 8.

Ghiotto, F.; Pateraki, C.; Tanskanen, J.; Severn, J. R.; Luehmann, N.; Kusmin, A.;

Stellbrink, J.; Linnolahti, M.; Bochmann, M., Probing the Structure of Methylalumoxane (MAO) by a Combined Chemical, Spectroscopic, Neutron Scattering, and Computational Approach. Organometallics 2013, 32, 3354-3362. 9.

Tritto, I.; Sacchi, M. C.; Li, S., NMR Study of the Reactions in Cp2TiMeCl/AlMe3 and

Cp2TiMeCl/Methylalumoxane Systems, Catalysts for Olefin Polymerization. Macromol. Rapid. Commun. 1994, 15, 217-223. 10. Tritto, I.; Sacchi, M. C.; Locatelli, P.; Li, S. X., Metallocene Ion Pairs: A Direct Insight Into the Reaction Equilibria and Polymerization by

13C

NMR Spectroscopy. Macromol. Symp.

1995, 89, 289-298. 11. Trefz, T. K.; Henderson, M. A.; Wang, M. Y.; Collins, S.; McIndoe, J. S., Mass Spectrometric Characterization of Methylaluminoxane. Organometallics 2013, 32, 3149-3152. 12. Linnolahti, M.; Severn, J. R.; Pakkanen, T. A., Formation of Nanotubular Methylaluminoxanes and the Nature of the Active Species in Single-Site α-Olefin Polymerization Catalysis. Angew. Chem. 2008, 120, 9419-9423. 13. Linnolahti, M.; Collins, S., Formation, Structure, and Composition of Methylaluminoxane. ChemPhysChem 2017, 18, 3369-3374.

ACS Paragon Plus Environment

31

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 51

14. Zurek, E.; Woo, T. K.; Firman, T. K.; Ziegler, T., Modeling the Dynamic Equilibrium between Oligomers of (AlOCH3)n in Methylaluminoxane (MAO). A Theoretical Study Based on a Combined Quantum Mechanical and Statistical Mechanical Approach. Inorg. Chem. 2001, 40, 361-370. 15. a) Linnolahti, M.; Severn, J. R.; Pakkanen, T. A., Are Aluminoxanes Nanotubular? Structural Evidence from a Quantum Chemical Study. Angew. Chem. Int. Ed. 2006, 45, 33313334; b) Boudene, Z.; De Bruin, T.; Toulhoat, H.; Raybaud, P., A QSPR Investigation of Thermal Stability of [Al(CH3)O]n Oligomers in Methylaluminoxane Solution: the Identification of a Geometry-Based Descriptor. Organometallics 2012, 31, 8312-8322. 16. Zurek, E.; Ziegler, T., A Combined Quantum Mechanical and Statistical Mechanical Study of the Equilibrium of Trimethylaluminum (TMA) and Oligomers of (AlOCH3)n Found in Methylaluminoxane (MAO) Solution. Inorg. Chem. 2001, 40, 3279-3292. 17. Trefz, T. K.; Henderson, M. A.; Linnolahti, M.; Collins, S.; McIndoe, J. S., Mass Spectrometric Characterization of Methylaluminoxane-Activated Metallocene Complexes. Chem. Eur. J. 2015, 21, 2980-2991. 18. Collins, S.; Linnolahti, M.; Zamora, M. G.; Zijlstra, H. S.; Rodríguez Hernández, M. T.; Perez-Camacho, O., Activation of Cp2ZrX2 (X = Me, Cl) by Methylaluminoxane As Studied by Electrospray Ionization Mass Spectrometry: Relationship to Polymerization Catalysis. Macromolecules 2017, 50, 8871-8884. 19. Talsi, E. P.; Semikolenova, N. V.; Panchenko, V. N.; Sobolev, A. P.; Babushkin, D. E.; Shubin, A. A.; Zakharov, V. A., The Metallocene/Methylaluminoxane Catalysts Formation: EPR

ACS Paragon Plus Environment

32

Page 33 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Spin Probe Study of Lewis Acidic Sites of Methylaluminoxane. J. Mol. Cat. A: Chem. 1999, 139, 131-137. 20. Rocchigiani, L.; Busico, V.; Pastore, A.; Macchioni, A., Probing the Interactions between All Components of The Catalytic Pool for Homogeneous Olefin Polymerisation by Diffusion NMR Spectroscopy. Dalton Trans. 2013, 42, 9104-9111. 21. Hansen, E. W.; Blom, R.; Kvernberg, P. O., Diffusion of Methylaluminoxane (MAO) in Toluene Probed by 1H NMR Spin-Lattice Relaxation Time. Macromol. Chem. Phys. 2001, 202, 2880-2889. 22. Pédeutour, J.-N.; Radhakrishnan, K.; Cramail, H.; Deffieux, A., Use of “TMA-depleted” MAO for The Activation of Zirconocenes in Olefin Polymerization. J. Mol. Cat. A: Chem. 2002, 185, 119-125. 23. Zurek, E.; Ziegler, T., Toward the Identification of Dormant and Active Species in MAO (Methylaluminoxane)-Activated,

Dimethylzirconocene-Catalyzed

Olefin

Polymerization.

Organometallics 2002, 21, 83-92. 24. Wieser, U.; Schaper, F.; Brintzinger, H.-H., Methylalumoxane (MAO)-Derived MeMAO− Anions in Zirconocene-Based Polymerization Catalyst Systems – A UV-Vis Spectroscopic Study. Macromol. Symp. 2006, 236, 63-68. 25. Endres, E.; Zijlstra, H. S.; Collins, S.; McIndoe, J. S.; Linnolahti, M., Oxidation of Methylalumoxane Oligomers: A Theoretical Study Guided by Mass Spectrometry. Organometallics 2018, 37, 3936-3942.

ACS Paragon Plus Environment

33

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 51

26. Theurkauff, G.; Bondon, A.; Dorcet, V.; Carpentier, J.-F.; Kirillov, E., Heterobi- and trimetallic Ion Pairs of Zirconocene-Based Isoselective Olefin Polymerization Catalysts with AlMe3. Angew. Chem. Int. Ed. 2015, 54, 6343-6346. 27. Ghiotto, F.; Pateraki, C.; Severn, J. R.; Friederichs, N.; Bochmann, M., Rapid Evaluation of Catalysts and MAO Activators by Kinetics: What Controls Polymer Molecular Weight and Activity in Metallocene/MAO Catalysts? Dalton Trans. 2013, 42, 9040-9048. 28. Busico, V.; Cipullo, R.; Cutillo, F.; Friederichs, N.; Ronca, S.; Wang, B., Improving the Performance of Methylalumoxane:  A Facile and Efficient Method to Trap “Free” Trimethylaluminum. J. Am. Chem. Soc. 2003, 125, 12402-12403. 29. Busico, V.; Cipullo, R.; Pellecchia, R.; Talarico, G.; Razavi, A., Hafnocenes and MAO: Beware of Trimethylaluminum! Macromolecules 2009, 42, 1789-1791. 30. a) Song, F.; Cannon, R. D.; Bochmann, M., Zirconocene-Catalyzed Propene Polymerization:  A Quenched-Flow Kinetic Study. J. Am. Chem. Soc. 2003, 125, 7641-7653; b) Machat, M. R.; Jandl, C.; Rieger, B., Titanocenes in Olefin Polymerization: Sustainable Catalyst System or an Extinct Species? Organometallics 2017, 36, 1408-1418. 31. Cipullo, R.; Melone, P.; Yu, Y.; Iannone, D.; Busico, V., Olefin Polymerisation Catalysts: When Perfection is Not Enough. Dalton Trans. 2015, 44, 12304-12311. 32. Ehm, C.; Cipullo, R.; Budzelaar, P. H. M.; Busico, V., Role(s) of TMA in Polymerization. Dalton Trans. 2016, 45, 6847-6855.

ACS Paragon Plus Environment

34

Page 35 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

33. Bochmann, M.; Lancaster, S. J., Monomer–Dimer Equilibria in Homo- and Heterodinuclear Cationic Alkylzirconium Complexes and Their Role in Polymerization Catalysis. Angew. Chem. Int. Ed. Eng. 1994, 33, 1634-1637. 34. 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. 35. Ehm, C.; Antinucci, G.; Budzelaar, P. H. M.; Busico, V., Catalyst Activation and The Dimerization Energy of Alkylaluminium Compounds. J. Organomet. Chem. 2014, 772–773, 161171. 36. a) Kickham, J. E.; Guérin, F.; Stewart, J. C.; Stephan, D. W., Five-Coordinate Carbides in Ti-Al-C Complexes. Angew. Chem. Int. Ed. 2000, 39, 3263-3266; b) Kickham, J. E.; Guérin, F.; Stewart, J. C.; Urbanska, E.; Stephan, D. W., Multiple C−H Bond Activation:  Reactions of Titanium−Phosphinimide Complexes with Trimethylaluminum. Organometallics 2001, 20, 11751182. 37. Dall’Occo, T.; Galimberti, M.; Resconi, L.; Albizzati, E.; Pennini, Catalysts and Processes for the Polymerization Of Olefins. PCT Int. Appl. WO 96/02580, 1996. 38. Galimberti, M.; Destro, M.; Fusco, O.; Piemontesi, F.; Camurati, I., Ethene/Propene Copolymerization from Metallocene-Based Catalytic Systems:  Role of the Alumoxane. Macromolecules 1999, 32, 258-263. 39. Zijlstra, H. S.; Joshi, A.; Linnolahti, M.; Collins, S.; McIndoe, J. S., Modifying Methylalumoxane via Alkyl Exchange. Dalton Trans. 2018, 47, 17291-17298.

ACS Paragon Plus Environment

35

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 51

40. Smith, G. M.; Palmaka, S. W.; Roger, J. S.; Malpass, D. B.; Monfiston, D. J., Polyalkylaluminoxane Compositions Formed by Non-Hydrolytic Means. PCT Int. Appl. WO9723288, 1997. 41. Stellbrink, J.; Niu, A.; Allgaier, J.; Richter, D.; Koenig, B. W.; Hartmann, R.; Coates, G. W.; Fetters, L. J., Analysis of Polymeric Methylaluminoxane (MAO) via Small Angle Neutron Scattering. Macromolecules 2007, 40, 4972-4981. 42. Tritto, I.; Sacchi, M. C.; Locatelli, P.; Li, S. X., Metallocenes/methylalumoxanes: A 13C NMR Study of the Reaction Equilibria and Polymerization. Macromol. Symp. 1995, 97, 101-108. 43. a) Stapleton, R. A.; Al-Humydi, A.; Chai, J.; Galan, B. R.; Collins, S., Sterically Hindered Aluminum Alkyls:  Weakly Interacting Scavenging Agents of Use in Olefin Polymerization. Organometallics 2006, 25, 5083-5092; b) Stapleton, R. A.; Galan, B. R.; Collins, S.; Simons, R. S.; Garrison, J. C.; Youngs, W. J., Bulky Aluminum Alkyl Scavengers in Olefin Polymerization with Group 4 Catalysts. J. Am. Chem. Soc. 2003, 125, 9246-9247. 44. Williams, V. C.; Dai, C.; Li, Z.; Collins, S.; Piers, W. E.; Clegg, W.; Elsegood, M. R. J.; Marder, T. B., Activation of [Cp2ZrMe2] with New Perfluoroaryl Diboranes: Solution Chemistry and Ethylene Polymerization Behavior in the Presence of MeAl(BHT)2. Angew. Chem. Int. Ed. 1999, 38, 3695-3698. 45. a) Ehm, C.; Cipullo, R.; Passaro, M.; Zaccaria, F.; Budzelaar, P. H. M.; Busico, V., Chain Transfer to Solvent in Propene Polymerization with Ti Cp-phosphinimide Catalysts: Evidence for Chain Termination via Ti–C Bond Homolysis. ACS Catal. 2016, 6, 7989-7993; b) Zaccaria, F.; Ehm, C.; Budzelaar, P. H. M.; Busico, V.; Cipullo, R., Catalyst Mileage in Olefin Polymerization: the Peculiar Role of Toluene. Organometallics 2018, 37, 2872-2879.

ACS Paragon Plus Environment

36

Page 37 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

46. a) Bashir, M. A.; Monteil, V.; Boisson, C.; McKenna, T. F. L., The Effect of Aluminum Alkyls and BHT-H on Reaction Kinetics of Silica Supported Metallocenes and Polymer Properties in Slurry Phase Ethylene Polymerization. J. Appl. Polym. Sci. 2018, 135, 45670; b) Collins, R. A.; Russell, A. F.; Scott, R. T. W.; Bernardo, R.; van Doremaele, G. H. J.; Berthoud, A.; Mountford, P., Monometallic and Bimetallic Titanium κ1-Amidinate Complexes as Olefin Polymerization Catalysts. Organometallics 2017, 36, 2167-2181; c) Sun, Y.; Xu, B.; Shiono, T.; Cai, Z., Highly Active ansa-(Fluorenyl)(amido)titanium-Based Catalysts with Low Load of Methylaluminoxane for Syndiotactic-Specific Living Polymerization of Propylene. Organometallics 2017, 36, 30093012; d) Cipullo, R.; Busico, V.; Fraldi, N.; Pellecchia, R.; Talarico, G., Improving the Behavior of Bis(phenoxyamine) Group 4 Metal Catalysts for Controlled Alkene Polymerization. Macromolecules 2009, 42, 3869-3872; e) Tynys, A.; Eilertsen, J. L.; Seppälä, J. V.; Rytter, E., Propylene Polymerizations with a Binary Metallocene System—Chain Shuttling Caused by Trimethylaluminium between Active Catalyst Centers. J. Polym. Sci. A: Polym. Chem. 2007, 45, 1364-1376; f) Turunen, J. P. J.; Pakkanen, T. T., Suppression of Metallocene Catalyst Leaching by the Removal of Free Trimethylaluminum from Methylaluminoxane. J. Appl. Polym. Sci. 2006, 100, 4632-4635; g) Romano, D.; Ronca, S.; Rastogi, S., Activation of a Bis-(Phenoxyimine) Titanium (IV) Catalyst Using Different Aluminoxane Co-Catalysts. Macromol. Symp. 2015, 356, 61-69; h) Tynys, A.; Eilertsen, J. L.; Rytter, E., Zirconocene Propylene Polymerisation: Controlling Termination Reactions. Macromol. Chem. Phys. 2006, 207, 295-303; i) Descour, C.; Sciarone, T. J. J.; Cavallo, D.; Macko, T.; Kelchtermans, M.; Korobkov, I.; Duchateau, R., Exploration of the Effect of 2,6-(t-Bu)2-4-Me-C6H2OH (BHT) in Chain Shuttling Polymerization. Polym. Chem. 2013, 4, 4718-4729; l) Quintanilla, E.; di Lena, F.; Chen, P., Chain Transfer to Aluminium in MAO-activated Metallocene-Catalyzed Polymerization Reactions. Chem. Commun.

ACS Paragon Plus Environment

37

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 51

2006, 4309-4311; m) Eric P. Wasserman, E. P.; Westwood, A. D.; Yu, Z.; Oskamb, J. H.; Duenas, S. L., Synchrotron XAS Studies on a Half-Sandwich Titanium-based Polyethylene Catalyst. J. Mol. Cat. A; 2001, 172, 67-80. 47. van Doremaele, G.; van Duin, M.; Valla, M.; Berthoud, A., On the Development of Titanium κ1-Amidinate Complexes, Commercialized as Keltan ACE™ Technology, Enabling the Production of an Unprecedented Large Variety of EPDM Polymer Structures. J. Polym. Sci. A: Polym. Chem. 2017, 55, 2877-2891. 48. Macchioni, A., Ion Pairing in Transition-Metal Organometallic Chemistry. Chem. Rev. 2005, 105, 2039-2074. 49. Zuccaccia, C.; Stahl, N. G.; Macchioni, A.; Chen, M.-C.; Roberts, J. A.; Marks, T. J., NOE and PGSE NMR Spectroscopic Studies of Solution Structure and Aggregation in Metallocenium Ion-Pairs. J. Am. Chem. Soc. 2004, 126, 1448-1464. 50. a) Alonso-Moreno, C.; Lancaster, S. J.; Zuccaccia, C.; Macchioni, A.; Bochmann, M., Evidence for Mixed-Ion Clusters in Metallocene Catalysts:  Influence on Ligand Exchange Dynamics and Catalyst Activity. J. Am. Chem. Soc. 2007, 129, 9282-9283; b) Rocchigiani, L.; Zuccaccia, C.; Zuccaccia, D.; Macchioni, A., Self-Aggregation Tendency of Zirconocenium Ion Pairs Which Model Polymer-Chain-Carrying Species in Aromatic and Aliphatic Solvents with Low Polarity. Chem. Eur. J. 2008, 14, 6589-6592; c) Alonso-Moreno, C.; Lancaster, S. J.; Wright, J. A.; Hughes, D. L.; Zuccaccia, C.; Correa, A.; Macchioni, A.; Cavallo, L.; Bochmann, M., Ligand Mobility

and

Solution

Structures

of

the

Metallocenium

Ion

Pairs

[Me2C(Cp)(fluorenyl)MCH2SiMe3+···X−] (M = Zr, Hf; X = MeB(C6F5)3, B(C6F5)4). Organometallics 2008, 27, 5474-5487; d) Rocchigiani, L.; Bellachioma, G.; Ciancaleoni, G.;

ACS Paragon Plus Environment

38

Page 39 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

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. 51. a) 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; b) Pregosin, P. S., Applications of NMR Diffusion Methods with Emphasis on ion Pairing in Inorganic Chemistry: a Mini-review. Magn. Res. Chem. 2017, 55, 405-413; c) Rocchigiani, L.; Macchioni, A., Disclosing the Multi-faceted World of Weakly Interacting Inorganic Systems by Means of NMR Spectroscopy. Dalton Trans. 2016, 45, 2785-2790; d) Macchioni, A.; Romani, A.; Zuccaccia, C.; Guglielmetti, G.; Querci, C., Experimental Evidence for the Aggregation of [(Phen)2Pd2(μ-H)(μ-CO)]+ in Solution. Organometallics 2003, 22, 15261533; e) Zuccaccia, D.; Pirondini, L.; Pinalli, R.; Dalcanale, E.; Macchioni, A., Dynamic and Structural NMR Studies of Cavitand-Based Coordination Cages. J. Am. Chem. Soc. 2005, 127, 7025-7032; f) Pettirossi, S.; Bellachioma, G.; Ciancaleoni, G.; Zuccaccia, C.; Zuccaccia, D.; Macchioni, A., Diffusion and NOE NMR Studies on Multicationic DAB-Organoruthenium Dendrimers: Size-Dependent Noncovalent Self-Assembly to Megamers and Ion Pairing. Chem. Eur. J. 2009, 15, 5337-5347; g) Rocchigiani, L.; Bellachioma, G.; Ciancaleoni, G.; Crocchianti, S.; Laganà, A.; Zuccaccia, C.; Zuccaccia, D.; Macchioni, A., Anion-Dependent Tendency of DiLong-Chain Quaternary Ammonium Salts to Form Ion Quadruples and Higher Aggregates in Benzene. ChemPhysChem 2010, 11, 3243-3254. 52. a) Li, D.; Keresztes, I.; Hopson, R.; Williard, P. G., Characterization of Reactive Intermediates by Multinuclear Diffusion-Ordered NMR Spectroscopy (DOSY). Acc. Chem. Res. 2009, 42, 270-280; b) Macchioni, A.; Ciancaleoni, G.; Zuccaccia, C.; Zuccaccia, D., Diffusion

ACS Paragon Plus Environment

39

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 51

Ordered NMR Spectroscopy (DOSY). In Supramolecular Chemistry, Gale, P. A.; Steed, J. W., Eds. John Wiley & Sons, Ltd.: Hoboken, New Jersey, USA, 2012; pp 319-330. 53. Eilertsen, J. L.; Hall, R. W.; Simeral, L. S.; Butler, L. G., Tools and Strategies for Processing Diffusion-Ordered 2D NMR Spectroscopy (DOSY) of a Broad, Featureless Resonance: an Application to Methylaluminoxane (MAO). Anal. Bioanal. Chem. 2004, 378, 15741578. 54. Stephan, D. W., The Road to Early-Transition-Metal Phosphinimide Olefin Polymerization Catalysts. Organometallics 2005, 24, 2548-2560. 55. Edward, J. T., Molecular Volumes and the Stokes-Einstein Equation. J. Chem. Educ. 1970, 47, 261. 56. 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. 57. The stability of s(MAO/BHT) in the presence of Ph3CH has been tested over >16h, observing no significant changes in the 1H NMR spectrum. 58. 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. 59. This strategy has been occasionally adopted to simplify interpretation of NMR spectra (see Supporting Information).

ACS Paragon Plus Environment

40

Page 41 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

60. 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. 61. Benn, R.; Janssen, E.; Lehmkuhl, H.; Rufińska, A.; Angermund, K.; Betz, P.; Goddard, R.; Krüger, C., Drei- oder Vierfach-Koordination des Aluminiums in Alkylaluminiumphenoxiden und deren Unterscheidung durch 27Al-NMR-Spektroskopie. J. Organomet. Chem. 1991, 411, 37-55. 62. Upper limit for VH0: 0.9*80 + 0.1*460 = 118 Å3; lower limit (assuming 430 Å3 for AlO(bht)): 0.9*67 + 0.1*430 = 103.3 Å3. The 1.8 correction factor for VH proposed by Stellbrink, Linnolahti and Bochmann (ref. 8) is considered for estimating n. 63. Kuklin, M. S.; Hirvi, J. T.; Bochmann, M.; Linnolahti, M., Toward Controlling the Metallocene/Methylaluminoxane-Catalyzed Olefin Polymerization Process by a Computational Approach. Organometallics 2015, 34, 3586-3597. 64. Zijlstra, H.; Collins, S.; McIndoe, J. S., Oxidation of Methylalumoxane Oligomers. Chem. Eur. J. 2018, 24, 5506-5512. 65. a) von Lacroix, K.; Heitmann, B.; Sinn, H., Behaviour of Differently Produced Methylalumoxanes in The Phase Separation with Diethyl Ether and Molecular Weight Estimations. Macromol. Symp. 1995, 97, 137-142; b) Eilertsen, J. L.; Rytter, E.; Ystenes, M., In Situ FTIR Spectroscopy During Addition of Trimethylaluminium (TMA) to Methylaluminoxane (MAO) Shows No Formation of MAO–TMA Compounds. Vib. Spectrosc. 2000, 24, 257-264.

ACS Paragon Plus Environment

41

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 42 of 51

66. Bryliakov, K. P.; Talsi, E. P.; Bochmann, M., 1H and 13C NMR Spectroscopic Study of Titanium(IV) Species Formed by Activation of Cp2TiCl2 and [(Me4C5)SiMe2NtBu]TiCl2 with Methylaluminoxane (MAO). Organometallics 2004, 23, 149-152. 67. Ma, K.; Piers, W. E.; Parvez, M., Competitive ArC−H and ArC−X (X = Cl, Br) Activation in Halobenzenes at Cationic Titanium Centers. J. Am. Chem. Soc. 2006, 128, 3303-3312. 68. Zaccaria, F.; Zuccaccia, C.; Cipullo, R.; Budzelaar, P. H. M.; Macchioni, A.; Busico, V.; Ehm, C., Toluene and α-Olefins as Radical Scavengers: Direct NMR Evidence for Homolytic Chain Transfer Mechanism Leading to Benzyl and “Dormant” Titanium Allyl Complexes. Organometallics 2018, 37, 4189-4194. 69. Bochmann, M.; Lancaster, S. J., Base-free Cationic Zirconium Benzyl Complexes as Highly Active Polymerization Catalysts. Organometallics 1993, 12, 633-640. 70. The high field shifted ipso carbon of the benzylic phenyl group in 1-Bn+[B(C6F5)4]- (δC = 146.4 ppm) with respect to 1-Bn2 (δC = 152.8 ppm) is in line with the hypothesis of an η2-benzyl coordination (see ref. 69), although no significant broadening of the Ti-CH2Ph signal is observed even after cooling a dichloromethane-d2 solution to 213 K. 71. a) Cabrera, L.; Hollink, E.; Stewart, J. C.; Wei, P.; Stephan, D. W., Cationic Methyl- and Chlorotitanium Phosphinimide Complexes. Organometallics 2005, 24 1091-1098; b) Smith, J. C.; Ma, K.; Piers, W. E.; Parvez, M.; McDonald, R., A New Weakly Coordinating Aluminate Anion Incorporating a Chelating Perfluoro-bis-anilido Ligand. Dalton Trans. 2010, 39, 10256-10263. 72. Based on 2D NMR characterization, the two most stable dimer configurations seem to be independent of the nature of the counterion (B(C6F5)4- vs [s(MAO/BHT)]- anions, Figure S4).

ACS Paragon Plus Environment

42

Page 43 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

73. Bochmann, M., Kinetic and Mechanistic Aspects of Metallocene Polymerisation Catalysts. J. Organomet. Chem. 2004, 689, 3982-3998. 74. Beck, S.; Prosenc, M.-H.; Brintzinger, H.-H.; Goretzki, R.; Herfert, N.; Fink, G., Binuclear Zirconocene Cations with μ-CH3-Bridges in Homogeneous Ziegler-Natta Catalyst Systems. J. Mol. Cat. A: Chem. 1996, 111, 67-79. 75. Siedle, A. R.; Newmark, R. A.; Lamanna, W. M.; Schroepfer, J. N., Exchange Reactions between Dialkylzirconocene and Alkylaluminium Compounds. Polyhedron 1990, 9, 301-308. 76. a) Coevoet, D.; Cramail, H.; Deffieux, A., U.V./visible Spectroscopic Study of the RacEt(Ind)2ZrCl2/MAO Olefin Polymerization Catalytic System, 1. Investigation in Toluene. Macromol. Chem. Phys. 1998, 199, 1451-1457; b) Beck, S.; Brintzinger, H. H., Alkyl Exchange between Aluminum Trialkyls and Zirconocene Dichloride Complexes — a Measure of Electron Densities at the Zr Center. Inorg. Chim. Acta 1998, 270, 376-381. 77. Bryliakov, K. P.; Babushkin, D. E.; Talsi, E. P.; Voskoboynikov, A. Z.; Gritzo, H.; Schröder, L.; Damrau, H.-R. H.; Wieser, U.; Schaper, F.; Brintzinger, H. H., ansa-Titanocene Catalysts

for

α-Olefin

Polymerization.

Syntheses,

Structures,

and

Reactions

with

Methylaluminoxane and Boron-Based Activators. Organometallics 2005, 24, 894-904. 78. Tymińska, N.; Zurek, E., DFT-D Investigation of Active and Dormant Methylaluminoxane (MAO) Species Grafted onto a Magnesium Dichloride Cluster: A Model Study of Supported MAO. ACS Catal. 2015, 5, 6989-6998.

ACS Paragon Plus Environment

43

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 44 of 51

79. Xu, Z.; Vanka, K.; Firman, T.; Michalak, A.; Zurek, E.; Zhu, C.; Ziegler, T., Theoretical Study of the Interactions between Cations and Anions in Group IV Transition-Metal Catalysts for Single-Site Homogeneous Olefin Polymerization. Organometallics 2002, 21, 2444-2453. 80. a) Zurek, E.; Ziegler, T., A theoretical Study of the Insertion Barrier of MAO (Methylaluminoxane)-Activated, Cp2ZrMe2-Catalyzed Ethylene Polymerization: further Evidence for the Structural Assignment of Active and Dormant Species. Faraday Discuss. 2003, 124, 93109; b) Oliva, L.; Oliva, P.; Galdi, N.; Pellecchia, C.; Sian, L.; Macchioni, A.; Zuccaccia, C., Solution Structure and Reactivity with Metallocenes of AlMe2F: Mimicking Cation–Anion Interactions in Metallocenium–Methylalumoxane Inner-Sphere Ion Pairs. Angew. Chem. Int. Ed. 2017, 56, 14227-14231. 81. Unfortunately, this approach cannot be easily extended to unmodified MAO since 1) phosphinimide precatalysts tend to decompose in the presence of TMA (ref. 36); 2) the presence of 'free' TMA would likely lead to more rapid Me/Bn scrambling competing with Bn abstraction; 3) Ti-Me+ ISIP give broad NMR signals complicating accurate quantification. 82. Strictly speaking, the results are not transferable to 1-Me2 activation, for which is more difficult to obtain a reliable estimation due to competing dimer formation and, possibly, solvent activation (Figure 3a). Nevertheless, it appears that the amount of MAO/BHT needed for full activation of 1-Me2 and 1-Bn2 are comparable, i.e. acidic sites abstract Bn and Me groups similarly easily. 83. a) Mason, M. R.; Smith, J. M.; Bott, S. G.; Barron, A. R., Hydrolysis of Tri-tertbutylaluminum: the First Structural Characterization of Alkylalumoxanes [(R2Al)2O]n and (RAlO)n. J. Am. Chem. Soc. 1993, 115, 4971-4984; b) Falls, Z.; Zurek, E.; Autschbach, J.,

ACS Paragon Plus Environment

44

Page 45 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

Computational Prediction and Analysis of the

27Al

Solid-state NMR Spectrum of

Methylaluminoxane (MAO) at Variable Temperatures and Field Strengths. Phys. Chem. Chem. Phys. 2016, 18, 24106-24118. 84. a) Hirvi, J. T.; Bochmann, M.; Severn, J. R.; Linnolahti, M., Formation of Octameric Methylaluminoxanes by Hydrolysis of Trimethylaluminum and the Mechanisms of Catalyst Activation in Single-Site α-Olefin Polymerization Catalysis. ChemPhysChem 2014, 15, 27322742; b) Velthoen, M. E. Z.; Muñoz-Murillo, A.; Bouhmadi, A.; Cecius, M.; Diefenbach, S.; Weckhuysen, B. M., The Multifaceted Role of Methylaluminoxane in Metallocene-Based Olefin Polymerization Catalysis. Macromolecules 2018, 51, 343-355. 85. a) Kleinschmidt, R.; van der Leek, Y.; Reffke, M.; Fink, G., Kinetics and Mechanistic Insight into Propylene Polymerization with Different Metallocenes and Various Aluminium Alkyls as Cocatalysts. J. Mol. Cat. A: Chem. 1999, 148, 29-41; b) Jüngling, S.; Mülhaupt, R., The Influence of Methylalumoxane Concentration on Propene Polymerization with Homogeneous Metallocene-based Ziegler-Natta Catalysts. J. Organomet. Chem. 1995, 497, 27-32. 86. Sishta, C.; Hathorn, R. M.; Marks, T. J., Group 4 Metallocene-Alumoxane Olefin Polymerization Catalysts. CPMAS-NMR Spectroscopic Observation of Cation-like Zirconocene Alkyls. J. Am. Chem. Soc. 1992, 114, 1112-1114. 87. Stephan, D. W.; Stewart, J. C.; Guérin, F.; Courtenay, S.; Kickham, J.; Hollink, E.; Beddie, C.; Hoskin, A.; Graham, T.; Wei, P.; Spence, R. E. v. H.; Xu, W.; Koch, L.; Gao, X.; Harrison, D. G., An Approach to Catalyst Design:  Cyclopentadienyl-Titanium Phosphinimide Complexes in Ethylene Polymerization. Organometallics 2003, 22, 1937-1947.

ACS Paragon Plus Environment

45

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 46 of 51

88. Mills, R., Self-diffusion in Normal and Heavy Water in the Range 1-45.deg. J. Phys. Chem. 1973, 77, 685-688. 89. Zijlstra, H. S.; Stuart, M. C. A.; Harder, S., Structural Investigation of Methylalumoxane Using Transmission Electron Microscopy. Macromolecules 2015, 48, 5116-5119. 90. Ehm, C.; Budzelaar, P. H. M.; Busico, V., Calculating accurate barriers for olefin insertion and related reactions. J. Organomet. Chem. 2015, 775, 39-49. 91. Gaussian 09 Revision B.1; for the full citation see the Supporting Information. 92. J. Baker, PQS, version 2.4, Parallel Quantum Solutions: Fayetteville, AR, 2001. 93. Baker, J., An Algorithm for the Location of Transition States. J. Comput. Chem. 1986, 7, 385-395. 94. Budzelaar, P. H. M., Geometry Optimization Using Generalized, Chemically Meaningful Constraints. J. Comput. Chem. 2007, 28, 2226-2236. 95. Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E., Climbing the Density Functional Ladder: Nonempirical Meta-Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401. 96. Balabanov, N. B.; Peterson, K. A., Systematically Convergent Basis Sets for Transition Metals. I. All-electron Correlation Consistent Basis Sets for the 3d Elements Sc–Zn. J. Chem. Phys. 2005, 123, 064107.

ACS Paragon Plus Environment

46

Page 47 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

97. Balabanov, N. B.; Peterson, K. A., Basis Set Limit Electronic Excitation Energies, Ionization Potentials, and Electron Affinities for the 3d Transition Metal Atoms: Coupled Cluster and Multireference Methods. J. Chem. Phys. 2006, 125, 074110. 98. Schuchardt, K. L.; Didier, B. T.; Elsethagen, T.; Sun, L. S.; Gurumoorthi, V.; Chase, J.; Li, J.; Windus, T. L., Basis Set Exchange: A Community Database for Computational Sciences. J. Chem. Inf. Model 2007, 47, 1045-1052. 99. a) Whitten, J. L., Coulombic Potential Energy Integrals and Approximations. J. Chem. Phys. 1973, 58, 4496; b) Baerends, E. J.; Ellis, D. E.; Ros, P., Self-consistent Molecular Hartree— Fock—Slater Calculations I. The Computational Procedure. Chem. Phys. 1973, 2, 41-51; c) Feyereisen, M.; Fitzgerald, G.; Komornicki, A., Use of Approximate Integrals in Ab Initio Theory. An Application in MP2 Energy Calculations. Chem. Phys. Lett. 1993, 208, 359-363; d) Vahtras, O.; Almlöf, J.; Feyereisen, M. W., Integral Approximations for LCAO-SCF Calculations. Chem. Phys. Lett. 1993, 213, 514-518. 100. Zhao, Y.; Truhlar, D. G., The M06 Suite Of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing af Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215-241. 101. a) Zaccaria, F.; Ehm, C.; Budzelaar, P. H. M.; Busico, V., Accurate Prediction of Copolymerization Statistics in Molecular Olefin Polymerization Catalysis: The Role of Entropic, Electronic, and Steric Effects in Catalyst Comonomer Affinity. ACS Catal. 2017, 7, 1512-1519; b) Zaccaria, F.; Cipullo, R.; Budzelaar, P. H. M.; Busico, V.; Ehm, C., Backbone Rearrangement During Olefin Capture as the Rate Limiting Step in Molecular Olefin Polymerization Catalysis

ACS Paragon Plus Environment

47

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 48 of 51

and its Effect on Comonomer Affinity. J. Polym. Sci. A: Polym. Chem. 2017, 55, 2807-2814; c) Ehm, C.; Vittoria, A.; Goryunov, G. P.; Kulyabin, P. S.; Budzelaar, P. H. M.; Voskoboynikov, A. Z.; Busico, V.; Uborsky, D. V.; Cipullo, R., Connection of Stereoselectivity, Regioselectivity, and Molecular

Weight

Capability

in

rac-R’2Si(2-Me-4-R-indenyl)2ZrCl2

Type

Catalysts.

Macromolecules 2018, 51, 8073-8083. 102. a) Ehm, C.; Budzelaar, P. H. M.; Busico, V., Metal–Carbon Bond Strengths Under Polymerization Conditions: 2,1-Insertion as a Catalyst Stress Test. J. Catal. 2017, 351, 146-152; b) Ehm, C.; Budzelaar, P. H. M.; Busico, V., Tuning the Relative Energies of Propagation and Chain Termination Barriers in Polyolefin Catalysis through Electronic and Steric Effects. Eur. J. Inorg. Chem. 2017, 3343-3349; c) Zorve, P.; Linnolahti, M., Adsorption of Titanium Tetrachloride on Magnesium Dichloride Clusters. ACS Omega 2018, 3, 9921-9928. 103. a) Ehm, C.; Krüger, J.; Lentz, D., How a Thermally Unstable Metal Hydrido Complex Can Yield High Catalytic Activity Even at Elevated Temperatures. Chem. Eur. J. 2016, 22, 9305-9310; b) Krüger, J.; Leppkes, J.; Ehm, C.; Lentz, D., Competition of Nucleophilic Aromatic Substitution, σ-Bond Metathesis, and syn Hydrometalation in Titanium(III)-Catalyzed Hydrodefluorination of Arenes. Chem. Asian J. 2016, 11, 3062-3071; c) Kruger, J.; Ehm, C.; Lentz, D., Improving Selectivity in Catalytic Hydrodefluorination by Limiting SNV Reactivity. Dalton Trans. 2016, 45, 16789-16798. 104. a) Zaccaria, F.; Vittoria, A.; Correa, A.; Ehm, C.; Budzelaar, P. H. M.; Busico, V.; Cipullo, R., Internal Donors in Ziegler-Natta Systems: is Reduction by AlR3 a Requirement for Donor Clean-Up? ChemCatChem 2018, 10, 984-988; b) Jaeger, A. D.; Ehm, C.; Lentz, D., Organocatalytic C-F Bond Activation with Alanes. Chem. Eur. J. 2018, 24, 6769-6777.

ACS Paragon Plus Environment

48

Page 49 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

105. Tomasi, J., Thirty Years of Continuum Solvation Chemistry: a Review, and Prospects for the Near Future. Theor. Chem. Acc. 2004, 112, 184-203. 106. a) Tobisch, S.; Ziegler, T., Catalytic Oligomerization of Ethylene to Higher Linear αOlefins Promoted by the Cationic Group 4 [(η5-Cp-(CMe2-bridge)-Ph)MII(ethylene)2]+ (M = Ti, Zr, Hf) Active Catalysts:  A Density Functional Investigation of the Influence of the Metal on the Catalytic Activity and Selectivity. J. Am. Chem. Soc. 2004, 126, 9059-9071; b) Dunlop-Brière, A. F.; Budzelaar, P. H. M.; Baird, M. C., α- and β-Agostic Alkyl–Titanocene Complexes. Organometallics 2012, 31, 1591-1594. 107. a) McWeeny, R., Perturbation Theory for the Fock-Dirac Density Matrix. Phys. Rev. 1962, 126, 1028-1034; b) Ditchfield, R., Self-consistent Perturbation Theory of Diamagnetism. Mol. Phys. 1974, 27, 789-807; c) Wolinski, K.; Hinton, J. F.; Pulay, P., Efficient Implementation of the Gauge-Independent Atomic Orbital Method for NMR Chemical Shift Calculations. J. Am. Chem. Soc. 1990, 112, 8251-8260; d) Cheeseman, J. R.; Trucks, G. W.; Keith, T. A.; Frisch, M. J., A Comparison of Models for Calculating Nuclear Magnetic Resonance Shielding Tensors. J. Chem. Phys. 1996, 104, 5497-5509. 108. a) Enevoldsen, T.; Oddershede, J.; Sauer, S. P. A., Correlated Calculations of Indirect Nuclear

spin-spin

Coupling

Constants

using

Second-Order

Polarization

Propagator

Approximations: SOPPA and SOPPA(CCSD). Theor. Chem. Acc. 1998, 100, 275-284; b) Sauer, S. P. A.; Raynes, W. T., Unexpected Differential Sensitivity of Nuclear Spin–Spin-Coupling Constants to Bond Stretching in BH4−, NH4+, and SiH4. J. Chem. Phys. 2000, 113, 3121-3129; c) Sauer, S. P. A.; Raynes, W. T.; Nicholls, R. A., Nuclear Spin–Spin Coupling in Silane and its Isotopomers: Ab Initio Calculation and Experimental Investigation. J. Chem. Phys. 2001, 115,

ACS Paragon Plus Environment

49

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 50 of 51

5994-6006; d) Provasi, P. F.; Aucar, G. A.; Sauer, S. P. A., The Effect of Lone Pairs and Electronegativity on the Indirect Nuclear Spin–Spin Coupling Constants in CH2X (X=CH2, NH, O, S): Ab Initio Calculations Using Optimized Contracted Basis Sets. J. Chem. Phys. 2001, 115, 1324-1334; e) Barone, V.; Provasi, P. F.; Peralta, J. E.; Snyder, J. P.; Sauer, S. P. A.; Contreras, R. H., Substituent Effects on Scalar 2J(19F,19F) and 3J(19F,19F) NMR Couplings:  A Comparison of SOPPA and DFT Methods. J. Phys. Chem. A 2003, 107, 4748-4754. 109. Schindler, M.; Kutzelnigg, W., Theory of Magnetic Susceptibilities and NMR Chemical Shifts in Terms of Localized Quantities. II. Application to Some Simple Molecules. J. Chem. Phys. 1982, 76, 1919-1933. 110. Kutzelnigg, W.; Fleischer, U.; Schindler, M., The IGLO-Method: Ab Initio Calculation Interpretation of NMR Chemical Shifts Magnetic Susceptibilities. In Deuterium and Shift Calculation, Fleischer, U.; Kutzelnigg, W.; Limbach, H. H.; Martin, G. J.; Martin, M. L.; Schindler, M., Eds. Springer: Berlin, Heidelberg, 1991; pp 165-262.

ACS Paragon Plus Environment

50

Page 51 of 51 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Catalysis

TOC Graphic

ACS Paragon Plus Environment