Organometallic Zirconium Compounds in an Oxygen-Rich

Jan 9, 2018 - Synopsis. Tris(oxoimidazolyl)hydroborato ligands, which serve as L2X [O3] donors, have been employed to obtain organometallic zirconium ...
0 downloads 9 Views 3MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Organometallic Zirconium Compounds in an Oxygen-Rich Coordination Environment: Synthesis and Structural Characterization of Tris(oxoimidazolyl)hydroboratozirconium Compounds Ahmed Al-Harbi, Matthew J. Hammond, and Gerard Parkin* Department of Chemistry, Columbia University, New York, New York 10027, United States S Supporting Information *

ABSTRACT: A series of tris(oxoimidazolyl)hydroborato ligands, which serve as L2X [O3] donors, have been employed to obtain organometallic zirconium compounds in an uncommon oxygen-rich coordination environment. For example, Cp[ToMeBenz]ZrCl2 has been synthesized via the reaction of [ToMeBenz]Na with CpZrCl3 and bears a structural resemblance to the bent metallocene dichloride derivative Cp2ZrCl2. t In addition, the half-sandwich counterparts [ToMeBenz]ZrCl3 and [ToBu ]ZrCl3 have t been obtained by metathesis of ZrCl4 with [ToMeBenz]Na and [ToBu ]Na, respectively. The structurally related zirconium benzyl compounds [ToRBenz]Zr(CH2Ph)3 (R = Me, But, 1-Ad) have also been synthesized via the reactions of [ToRBenz]Tl with Zr(CH2Ph)4, and X-ray diffraction studies demonstrate that the benzyl ligands in these compounds are conformationally flexible and exhibit a large range of Zr−CH2−Ph bond angles (94.7−131.7°). Protolytic cleavage of one of the benzyl ligands of [ToRBenz]Zr(CH2Ph)3 (R = But, 1-Ad) may be achieved by treatment with [PhNHMe2][B(C6F5)4] to generate {[ToRBenz]Zr(CH2Ph)2}[B(C6F5)4], which are catalysts for the t polymerization of ethylene. The molecular structure of the ether adduct, {[ToBu Benz]Zr(CH2Ph)2(OEt2)}[B(C6F5)4], has been determined by X-ray diffraction. In addition to the use of tris(oxoimidazolyl)hydroborato ligands, bis(oxoimidazolyl)hydroborato ligands have also been used to obtain zirconium benzyl compounds in oxygen-rich environments, namely, [BoMeBenz]2Zr(CH2Ph)2 and [BoAdBenz]2Zr(CH2Ph)2.





INTRODUCTION

RESULTS AND DISCUSSION One approach for incorporating [ToR] ligands into zirconium compounds involves thet metathesis of a Zr−Cl bond with [ToR]Na.15 Thus, [ToBu ]ZrCl3 and [ToMeBenz]ZrCl3 may be t obtained via the respective reactions of ZrCl4 with [ToBu ]Na and MeBenz ]Na (Scheme 1), and their molecular structures have [To been determined by X-ray diffraction, as illustrated in Figures 1 and 2. In addition to metathesis reactions with ZrCl4, CpZrCl3 may be employed to afford the hybrid compound Cp[ToMeBenz]ZrCl2, which contains both Cp and [ToMeBenz] ligands (Scheme 1). The successful synthesis of the zirconocene analogue Cp[ToMeBenz]ZrCl2, which has been structurally characterized by X-ray diffraction (Figure 3), is of significance because the corresponding reaction of CpZrCl3 with [CpCo{P(O)(OEt)2}3]Na does not yield Cp[CpCo{P(O)(OEt)2}3]ZrCl2. Specifically, rather than undergoing metathesis of the Zr−Cl bond, CpZrCl3 reacts with [CpCo{P(O)(OEt)2}3]Na to afford [CpCo{P(O)(OEt)2}3]ZrCl3 due to the preferential substitution of the cyclopentadienyl ligand.16,17 Another illustration of the difficulty of obtaining zirconocene analogues using the [CpCo{P(O)(OEt)2}3] ligand is provided by the fact that the reaction of ZrCl4(THF)2 with [CpCo{P(O)(OEt)2}3]Na forms a dinuclear compound, resulting from the dealkylation of two phosphonate alkyl groups per

Cyclopentadienyl ligands have played an important role in the development of organometallic chemistry, as exemplified by the metallocene systems and their use as olefin polymerization catalysts.1,2 Electronically related to cyclopentadienyl are C3-symmetric L2X3 tripodal ligands that feature a variety of donor arrays,4,5 of which the tris(pyrazolyl)hydroborato [N3] donor ligand system, [TpR,R′], has been studied most extensively.4 By comparison, C3-symmetric [O3] donors are less common6−9 but offer potential for mimicking molecular species that are grafted to oxide surfaces.10−13 With respect to L2X [O3] donors, the most commonly encountered class of ligands is that of the [CpCo{P(O)(OR)2}3] ligand system.6 Despite the formal analogy between [CpCo{P(O)(OR)2}3] and [TpR,R′] ligands, however, an important distinction is that the different placement of the R substituents results in the [O3] ligands being much less sterically demanding than [TpR,R′]. Therefore, since sterically demanding ligands are often advantageous for supporting metal centers with uncommon coordination environments, we recently introduced a new class of L2X [O3] ligands that are based on oxoimidazolyl donors, namely, tris(oxoimidazolyl)hydroborato, [ToR].14 In addition, we have also reported related bis(oxoimidazolyl)hydroborato ligands, [BoR], which provide complementary LX [O2] donor arrays. Here, we describe the application of these bis- and tris(oxoimidazolyl)hydroborato ligands to organozirconium chemistry. © XXXX American Chemical Society

Received: November 6, 2017

A

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Scheme 1

Figure 3. Molecular structure of Cp[ToMeBenz]ZrCl2.

zirconium, rather than [CpCo{P(O)(OEt)2}3]2ZrCl2.16 Moreover, ZrCl4(THF)2 reacts with the ethyl counterpart [CpCo{P(O)Et2}3]Na to afford ionic {[CpCo{P(O)Et)2}3]2Zr}Cl2, in which the chlorides do not coordinate to the zirconium, and the complex possesses a six-coordinate octahedral geometry.16 The use of the [ToMeBenz] ligand, therefore, provides access to a long sought-after zirconocene analogue in which a cyclopentadienyl ligand is replaced with an [O3] donor array. In accord with the bent metallocene analogy, the molecular structure of Cp[ToMeBenz]ZrCl2 (Figure 3) bears a resemblance to that of Cp2ZrCl2,18 as summarized in Table 1. For example, Table 1. Comparison of the Metrical Data for Cp[ToMeBenz]ZrCl2, Cp[TmMe]ZrCl2, and Cp2ZrCl2

Zr−Crange/Å Zr−Cav/Å Zr−Ccent/Å Zr−Cl/Å Cl−Zr−Cl/deg Cpcent−Zr−Y/deg

t

Cp[TmMe]ZrCl2a

Cp2ZrCl2b

Cp[ToMeBenz]ZrCl2

2.51−2.57 2.54 2.26 2.52 97.9 133.5 (Y = B)

2.47−2.52 2.50 2.20 2.45 97.0 129.2 (Y = Cpcent)

2.51−2.57 2.54 2.24 2.48 93.9 130.5 (Y = B)

Figure 1. Molecular structure of [ToBu ]ZrCl3.

a

Figure 2. Molecular structure of [ToMeBenz]ZrCl3.

the Cpcent−Zr−B angle of Cp[ToMeBenz]ZrCl2 (130.5°) is similar to the Cpcent−Zr−Cpcent angle of Cp2ZrCl2 (129.2°),18a while the Cl−Zr−Cl bond angle of Cp[ToMeBenz]ZrCl2 (93.9°) is similar to that of Cp2ZrCl2 (97.0°). The molecular structure of Cp[ToMeBenz]ZrCl2 is also similar to that of the mercapto counterpart Cp[TmMe]ZrCl2,19 but an interesting difference pertains to the 1H NMR spectroscopic properties. Specifically, whereas the methyl groups of the [ToMeBenz] ligand of Cp[ToMeBenz]ZrCl2 adopt a 2:1 pattern,20 those of the [TmMe] ligand of Cp[TmMe]ZrCl2 adopt a 1:1:1 pattern.19 The origin of this difference is associated with the chirality of Cp[ToMeBenz]ZrCl2 and Cp[TmMe]ZrCl2 due to the propeller-like twist21 of the [ToMeBenz] and [TmMe] ligands. Thus, in the absence of enantiomer interconversion, a 1:1:1 pattern of methyl groups is predicted, while a 2:1 pattern is predicted if interconversion is facile and the ligand arm that is trans to the cyclopentadienyl ligand remains in place. The observation that Cp[ToMeBenz]ZrCl2 exhibits a 2:1 pattern, while Cp[TmMe]ZrCl2 retains a 1:1:1 pattern,19 indicates that the barrier to B

Data taken from ref 19. bData taken from ref 18a.

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry enantiomer interconversion of Cp[ToMeBenz]ZrCl2 is significantly lower than that of Cp[TmMe]ZrCl2.22 Enantiomer interconversion requires the “inversion” of each arm, and the mechanism could involve either dissociation of the chalcogen23 or direct inversion at the coordinated chalcogen. Assuming that the enantiomer interconversions for Cp[ToMeBenz]ZrCl2 and Cp[TmMe]ZrCl2 occur via a common mechanism, the observation that the process is more facile for Cp[ToMeBenz]ZrCl2 suggests that the mechanism does not involve dissociation of an arm because the Zr−O bonds are expected to be stronger than the corresponding Zr−S bonds.24 Additional support for a nondissociative mechanism is provided by the observation that barriers to inversion are typically lower for second-row rather than third-row elements.25 Calculations on [TmH]ZnCl and [TmH]Mn(CO)3 provide further evidence that enantiomer interconversion of such ligands may occur via a nondissociative mechanism that features a transition state in which both the metal and boron reside in the mercaptoimidazolyl plane.23b The approach to the transition state is accompanied by an increase in the M−S−C bond angle, and since the Zr−O−C angles for Cp[ToMeBenz]ZrCl2 (132.3− 137.1°) are significantly greater than the corresponding Zr−S−C angles for Cp[TmMe]ZrCl2 (107.9−113.8°),19,26,27 it is evident that it would be easier for the oxoimidazolyl system Cp[ToMeBenz]ZrCl2 to access the transition state than would the mercapto system Cp[TmMe]ZrCl2. Thus, the tris(oxoimidazolyl)hydroborato ligands may be considered to be more flexible than their sulfur counterparts in these systems. In addition to the synthesis of chloride derivatives, we also sought benzyl derivatives because such compounds have been used extensively as polymerization catalysts,28 including heterogenized single-site catalysts.29 Although compounds of the class [ToR]Zr(CH2Ph)3 could, in principle, be synthesized via the alkylation of [ToR]ZrCl3, previous studies have noted synthetic difficulties associated with this approach for group 4 metal alkyl compounds of the type [TpR,R′]MRxX3−x.30 An alternative approach for the synthesis of [TpR,R′]M alkyl derivatives, however, involves the direct metathesis of M−R bonds. For example, we previously reported that thallium compounds of the type [TpR,R′]Tl are effective reagents for the synthesis of [TpR,R′]MRx−1 derivatives (e.g., M = Mg, Zn, Al) via reactions with MRx,31,32 because decomposition of the incipient TlR byproduct provides an effective driving force for the reaction.33 Accordingly, the benzyl compounds [ToR]Zr(CH2Ph)3 may be conveniently obtained by the metathesis of Zr(CH2Ph)4 with the corresponding thallium reagent [ToR]Tl (Scheme 2), an approach that has

Scheme 2

Figure 4. Molecular structure of [ToMeBenz]Zr(CH2Ph)3 (modification A).

t

also been employed for [TmBu ]M(CH2Ph)3 (M = Zr, Hf)34 and [TpMs*]Hf(CH2Ph)3.30,35 The molecular structures of [ToR]Zr(CH2Ph)3 have been determined by X-ray diffraction, as illustrated in Figures 4−9. Previous studies have demonstrated that benzyl ligands can adopt a variety of coordination modes, which include η1, η2, η3, η4, and η7 hapticities, and we have recently provided a means to classify these coordination modes.36,37 In this regard, it is interesting to note that the methylbenzimidazolyl derivative, [ToMeBenz]Zr(CH2Ph)3, has been structurally characterized in several modifications that exhibit different conformations of the benzyl ligands, with a range of bond angles. The conformation of the benzyl ligands in [ToR]Zr(CH2Ph)3 can be quantified by the magnitude of the B···Zr−C−Ph torsion angle (Figure 10), with values of |τ| > 90° being classified as “up” (with the phenyl group directed away from the [ToMeBenz]Zr moiety) and values of |τ| < 90° being classified as “down”

Figure 5. Molecular structure of [ToMeBenz]Zr(CH2Ph)3 (modification B).

(with the phenyl group directed towards the [ToMeBenz]Zr moiety). With this perspective, one of the modifications, which has approximately C3 molecular symmetry, possesses a “3 down” conformation, whereas the other modifications possess “1 down:2 up” conformations. Adopting the criteria that Zr−CH2−Ph bond angles of >97° and ≤97° are respectively used to classify η1 and η2 coordination,36 the three benzyl ligands of the “3 down” conformer, with Zr−CH2−Ph bond angles in the range 103.7−111.4°, are appropriately described as η1, which is the most commonly encountered coordination mode for benzyl ligands.38 In contrast, the “1 down:2 up” conformers each possess one benzyl group that approaches an η2 description, with Zr−CH2−Ph bond angles of 94.6° and 99.8° (Table 2). Such variability of the benzyl ligand coordination mode is in accord with our previous study, which demonstrates that the C

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 6. Molecular structure of [ToMeBenz]Zr(CH2Ph)3 (modification C). Figure 9. Molecular structure of [ToAdBenz]Zr(CH2Ph)3.

t

Figure 7. Molecular structure of [ToBu Benz]Zr(CH2Ph)3 (modification A).

Figure 8. Molecular structure of [ToBu Benz]Zr(CH2Ph)3 (modification B).

Figure 10. Classification of of the benzyl conformation according to the B···Zr−CH2−Ph torsion angle (τ).

Zr−CH2−Ph bond angle is very flexible and that relatively little energy is required to perturb it.36 For example, the Zr−CH2−Ph bond angles in structurally characterized compounds listed in the Cambridge Structural Database (CSD)27 range from 62.1°39 to

144.4°.40 As an illustration of the flexibility within a single compound, we have reported that the Zr−CH2−Ph bond angles within Zr(CH2Ph)4 range from 81.6(1)° to 106.7(2)°,36 while Arnold has noted that {CyNC[N(SiMe3)2]NCy}Zr(CH2Ph)3

t

D

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

and adamantyl groups that would destabilize “down” conformations. Another interesting feature of [ToRBenz]Zr(CH2Ph)3 is that the two methylene hydrogen atoms of each benzyl ligand appear as a singlet in the 1H NMR spectrum, even though each molecule possesses a molecular C3 axis, such that the hydrogen atoms are expected to be diastereotopic and, hence, chemically inequivalent.43 The fact that the methylene groups appear as a singlet is, nevertheless, rationalized by facile enantiomer interconversion due to propeller inversion of the [ToRBenz] ligand, as noted above for Cp[ToMeBenz]ZrCl2. In this regard, it is pertinent to note that the methylene hydrogen atoms of the mercaptoimidazolyl t counterpart [TmBu ]Zr(CH2Ph)3 appear as two doublets in the 1 H NMR spectrum.34 As such, the result is in accord with the spectral differences noted above for Cp[ToMeBenz]ZrCl2 and Cp[TmMe]ZrCl2, which also indicate that the oxoimidazolyl system is more flexible than the sulfur counterpart. In addition to incorporating [ToR] ligands via the reactions of Zr(CH2Ph)4 with the thallium reagent, the bis(oxobenzimidazolyl) compounds [BoMeBenz]2Zr(CH2Ph)2 and [BoAdBenz]2Zr(CH2Ph)2 may be synthesized analogously (Scheme 3).

Table 2. Benzyl Ligand Conformations of [ToR]Zr(CH2Ph)3 As Classified by B···Zr−C−Ph Torsion (τ) and Zr−CH2−Ph Bond Angles τ/deg [ToMeBenz]Zr(CH2Ph)3 (modification A)

[ToMeBenz]Zr(CH2Ph)3 (modification B)

[ToMeBenz]Zr(CH2Ph)3 (modification C)

t

[ToBu Benz]Zr(CH2Ph)3 (modification A)

t

[ToBu Benz]Zr(CH2Ph)3 (modification B)

[ToAdBenz]Zr(CH2Ph)3

a

Zr−CH2−Ph/deg classificationa

39.43

121.19

d, η1

114.31 129.81 40.47

115.67 94.62 118.07

u, η1 u, η2 d, η1

117.96 129.58 54.95

116.88 99.84 111.44

u, η1 u, η2 d, η1

57.59 61.81 141.83

103.70 111.30 120.86

d, η1 d, η1 u, η1

141.83 141.83 55.63

120.86 120.86 131.73

u, η1 u, η1 d, η1

81.08 178.37 142.49 142.49 142.50

128.77 122.25 122.75 122.75 122.75

d, η1 u, η1 u, η1 u, η1 u, η1

Scheme 3

d = down; u = up.

exists as two polymorphs with Zr−CH2−Ph bond angles that span the ranges 88.7(4)−123.2(4)° and 104.6(2)−115.9(2)°.41,42 Although density functional theory (DFT) geometry optimization (B3LYP) on several conformations of [ToMeBenz]Zr(CH2Ph)3 indicates that the “3 down” conformer is the more stable (Figure 11), the difference is small, such that crystal packing effects are sufficient to allow the observation of different conformers. t Interestingly, X-ray diffraction studies on [ToBu Benz]Zr(CH2Ph)3 AdBenz (Figures 7 and 8) and [To ]Zr(CH2Ph)3 (Figure 9) reveal conformations that are different to those observed experimentally t for [ToMeBenz]Zr(CH2Ph)3. Of particular note, both [ToBu Benz]Zr(CH2Ph)3 (Figure 7) and [ToAdBenz]Zr(CH2Ph)3 (Figure 9) exhibit “3 up” conformations. By comparison to [ToMeBenz]Zr(CH2Ph)3, the occurrence of the “3 up” conformations for t [ToBu Benz]Zr(CH2Ph)3 and [ToAdBenz]Zr(CH2Ph)3 is undoubtedly facilitated by the greater steric demands of the tert-butyl

The molecular structure of [BoMeBenz]2Zr(CH2Ph)2 (Figure 12) has been determined by X-ray diffraction, thereby demonstrating that coordination of the two [BoRBenz] ligands is supplemented by three-center−two-electron44 Zr···H−B secondary interactions.45 This coordination mode is reproduced well by DFT geometry optimization on [BoMeBenz]2Zr(CH2Ph)2 (Figure 13), as illustrated by a comparison of the experimental (3.33 and 3.45 Å) and calculated (3.36 and 3.54 Å) Zr···B distances.

Figure 11. Relative energies of DFT geometry-optimized conformers of [ToMeBenz]Zr(CH2Ph)3. E

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

t

Figure 14. Molecular structure of {[ToBu Benz]Zr(CH2Ph)2(OEt2)}[B(C6F5)4] (only the cation is shown for clarity).

Figure 12. Molecular structure of [BoMeBenz]2Zr(CH2Ph)2.

t

{[ToBu Benz]Zr(CH2Ph)2(OEt2)}+ is noteworthy because treatment of [TpMe2]Zr(CH2Ph)3 with [Ph3C][B(C6F5)4] at −60 °C generates {[TpMe2]Zr(CH2Ph)2}+, but decomposition occurs upon warming to 0 °C, resulting in the formation of {[PhCH2BpMe2]Zr(CH2Ph)(pzMe2)}+.40 t Despite its cationic nature, however, {[ToBu Benz]Zr(CH2Ph)2(OEt2)}[B(C6F5)4] is not an effective active catalyst for ethylene t polymerization. Postulating that the inefficiency of {[ToBu Benz]Zr(CH2Ph)2(OEt2)}[B(C6F5)4] as a catalyst is a consequence of the coordinated ether inhibiting ethylene binding, the possibility of obtaining a catalytic system by using a less strongly coordinating solvent was investigated. It is, therefore, significant that treatment of t [ToBu Benz]Zr(CH2Ph)3 with [PhNHMe2][B(C6F5)4] in benzene t

affords {[ToBu Benz]Zr(CH2Ph)2}[B(C6F5)4], which does serve as an effective catalyst, with an activity of 48 kg of polyethylene (PE) [mol of Zr]−1 [h]−1 [atm of C2H4]−1. [ToAdBenz]Zr(CH2Ph)3 activated by [PhNHMe2][B(C6F5)4] behaves similarly, with an activity of 160 kg of PE [mol of Zr]−1 [h]−1 [atm of C2H4]−1.49 In accord with the suggestion that the low activity of t {[ToBu Benz]Zr(CH2Ph)2(OEt2)}+ is a consequence of coordiof ethylene, DFT nated Et2O inhibiting the association t geometry optimization of {[ToBu Benz]Zr(CH2Ph)2(OEt2)}+ and t {[ToBu Benz]Zr(CH2Ph)2}+ (Figure 15) indicates that coordination of Et2O to zirconium is accompanied by an energy change of ΔESCF = −4.43 kcal mol−1.50

Figure 13. DFT geometry-optimized structure of [BoMeBenz]2Zr(CH2Ph)2 (hydrogen atoms on carbon atoms are omitted for clarity).

Generation of Cationic {[ToRBenz]Zr(CH2Ph)2(OEt2)}+ and {[ToRBenz]Zr(CH2Ph)2}+ and Their Use as Polymerization Catalysts. The use of cyclopentadienyl metal complexes, including metallocenes, for the generation of single-site olefin polymerization catalysts is now well established and is of considerable industrial interest.1 A more recent extension of this area has focused on the discovery of noncyclopentadienyl catalysts.46,47 For example, attention has been directed towards the use of the N3 donor [TpR,R′] ligands, which belong to the same L2X covalent bond classification3 as cyclopentadienyl.48 As an illustration, both [TpR,R′]ZrCl348a and [TpR,R′]Zr(CH2Ph)340,48b have been employed as precatalysts for olefin polymerization. Therefore, we have investigated the ability of the [O3] donor [ToRBenz] ligand to provide a catalytic system. Common methods for activating neutral zirconium alkyl compounds for the polymerization of olefins involve (i) alkyl abstraction via reaction with either [Ph3C][B(C6F5)4] or B(C6F5)3 and (ii) protolytic cleavage with [PhNHMe2][B(C6F5)4].1 In t this regard, treatment of [ToBu Benz]Zr(CH2Ph)3 with [PhNHMe2]-



SUMMARY In summary, bis- and tris(oxoimidazolyl)hydroborato ligands have been employed to obtain zirconium compounds in oxygenrich coordination environments that are uncommon for organometallic derivatives of this element. For example, the synthesis of the zirconocene counterpart Cp[ToMeBenz]ZrCl2 is noteworthy because the corresponding reaction of CpZrCl3 with [CpCo{P(O)(OEt)2}3]Na does not yield the analogous zirconocene species, Cp[CpCo{P(O)(OEt)2}3]ZrCl2, but rather forms [CpCo{P(O)(OEt)2}3]ZrCl3 due to the preferential displacement of the cyclopentadienyl ligand from zirconium. The reactivity of [ToRBenz]Tl (R = Me, But, 1-Ad) towards Zr(CH2Ph)4 provides a means to obtain benzyl compounds [ToRBenz]Zr(CH2Ph)3, which are counterparts to half-sandwich derivatives CpRZr(CH2Ph)3. In accord with this analogy,

t

[B(C6F5)4] in Et2O affords {[ToBu Benz]Zr(CH2Ph)2(OEt2)}[B(C6F5)4] (Scheme 2), which has been structurally characterized by X-ray diffraction (Figure 14). The ability to isolate F

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

suspension. The mixture was treated with n-hexane (5 mL) to precipitate more material, which was isolated by filtration. The precipitate was washed with n-hexane, dried in vacuo, and then extracted with dichloromethane (5 mL). The volatile components were removed in vacuo to give [ToMeBenz]CpZrCl2 as a white powder (30 mg, 67%). Anal. Calcd for Cp[ToMeBenz]ZrCl2: C, 51.2; H, 4.0; N, 12.3. Found: C, 50.8; H, 3.9; N, 11.3. t

t

Figure 15. DFT geometry-optimized structures of {[ToBu Benz]Zrt

(CH2Ph)2(OEt2)}+ and {[ToBu Benz]Zr(CH2Ph)2}+ (hydrogen atoms on carbon atoms are omitted for clarity).

protolytic cleavage of [ToRBenz]Zr(CH2Ph)3 with [PhNHMe2][B(C6F5)4] affords {[ToRBenz]Zr(CH2Ph)2}+, which can serve as a catalyst for the polymerization of ethylene.



t

Synthesis of [ToBu ]ZrCl3. A mixture of [ToBu ]Na (16 mg, 0.04 mmol) and ZrCl4 (12.3 mg, 0.05 mmol) was treated with benzene (1 mL) and heated at 60 °C for 4 h. The mixture was filtered, and the solid was extracted with chloroform (3 × 3 mL). The volatile components were removed in vacuo, and the residue obtained was washed t with hexanes (3 mL), yielding [ToBu ]ZrCl3 as an off-white powder (6 mg, 27%). Synthesis of [ToMeBenz]Zr(CH2Ph)3. A solution of Zr(CH2Ph)4 (27.7 mg, 0.06 mmol) in benzene (6 mL) was added to [ToMeBenz]Tl (40.0 mg, 0.06 mmol). The mixture was stirred at room temperature for 10 min, resulting in the immediate deposition of thallium. The mixture was filtered and the filtrate was treated with pentane (20 mL), thereby precipitating [ToMeBenz]Zr(CH2Ph)3. The supernatant was decanted, and the product was washed with pentane (5 mL) and dried in vacuo to afford [ToMeBenz]Zr(CH2Ph)3 (20.0 mg, 40%) as a yellow powder. Crystals of [ToMeBenz]Zr(CH2Ph)3 suitable for X-ray diffraction were obtained from a solution in CH2Cl2. Anal. Calcd for [ToMeBenz]Zr(CH2Ph)3: C, 64.5; H, 5.5; N, 10.8. Found: C, 64.3; H, 5.4; N, 10.6.

EXPERIMENTAL SECTION

General Considerations. All manipulations were performed using a combination of glovebox, high vacuum, and Schlenk techniques under a nitrogen or argon atmosphere.51 Solvents were purified and degassed by standard procedures. 1H NMR spectra were measured on Bruker 300 DRX, Bruker 400 DRX, Bruker 400 Cyber-enabled Avance III, and Bruker Avance 500 DMX spectrometers. 1H NMR chemical shifts (see the Supporting Information, SI) are reported in ppm relative to SiMe4 (δ = 0) and were referenced internally with respect to the protio solvent impurity (δ 7.16 for C6D5H and 5.32 for CDHCl2).52 13C NMR chemical shifts (see the SI) are reported in ppm relative to SiMe4 (δ = 0) and were referenced internally with respect to the solvent (δ 128.06 for C6D6 and 53.84 for CD2Cl2).52 Coupling constants are given in hertz. IR spectra were recorded on a PerkinElmer Spectrum Two spectrometer and are reported in reciprocal centimeters. Mass spectra were obtained on a Jeol JMS-HX110H tandem double-focusing mass spectrometer with a 10 kV accelerating voltage equipped with a fast-atom-bombardment ion source. CpZrCl3 and Cp2ZrCl2 were obtained from Sigma-Aldrich, while Zr(CH2Ph)436 and [ToR]M (M = Na, Tl)14,53 were prepared by literature methods. Caution! Thallium compounds are toxic and must be handled and disposed of safely. X-ray Structure Determinations. X-ray diffraction data were collected on a Bruker Apex II diffractometer. The structures were solved using direct methods and standard difference map techniques and were refined by full-matrix least-squares procedures on F2 with SHELXTL (version 2014/7).54 Crystallographic data have been deposited with the Cambridge Crystallographic Data Centre (CCDC 1584045−1584052).55 Computational Details. Calculations were carried out using DFT, as implemented in the Jaguar 8.9 (release 15) suite of ab initio quantum chemistry programs.56 Geometry optimizations were performed with the B3LYP density functional using the LACVP** basis set. Cartesian coordinates, and the energies of the geometry-optimized structures are provided in the SI. Synthesis of [ToMeBenz]ZrCl3. A mixture of [ToMeBenz]Na·diglyme (40 mg, 0.07 mmol) and ZrCl4 (18 mg, 0.08 mmol) was placed in an ampoule, treated with dichloromethane (6 mL), and heated overnight at 50 °C. After this period, the mixture was filtered and the volatile components were removed from the filtrate in vacuo. The solid residue was washed with acetonitrile (3 mL) and hexane (3 mL) to yield [ToMeBenz]ZrCl3 as a white powder (14 mg, 33%). Anal. Calcd for [ToMeBenz]ZrCl3·CH2Cl2: C, 40.8; H, 3.3; N, 11.4. Found: C, 41.0; H, 3.4; N, 11.0. Synthesis of Cp[ToMeBenz]ZrCl2. A mixture of CpZrCl3 (18 mg, 0.07 mmol) and [ToMeBenz]Na·diglyme (40 mg, 0.07 mmol) was placed in an ampoule and treated with benzene (5 mL). The mixture was stirred at room temperature for a period of 2 h, during which it became a

t

Synthesis of [ToBu Benz]Zr(CH2Ph)3. A solution of Zr(CH2Ph)4 t

(23 mg, 0.05 mmol) in toluene (3 mL) was added to [ToBu Benz]Tl (40 mg, 0.05 mmol). The mixture was stirred at room temperature for 10 min, resulting in the deposition of thallium. After this period, the t

mixture was filtered. Yellow crystals of [ToBu Benz]Zr(CH2Ph)3 were obtained by the diffusion of pentane into the filtrate at −15 °C, isolated, washed with pentane, and dried in vacuo (27.0 mg, 57%). Crystals of t [ToBu Benz]Zr(CH2Ph)3 suitable for X-ray diffraction were obtained by t the slow diffusion of pentane into a toluene solution of [ToBu Benz]ZrButBenz (CH2Ph)3 at −15 °C. Anal. Calcd for [To ]Zr(CH2Ph)3: C, 68.7; H, 6.5; N, 8.9. Found: C, 67.5; H, 6.5; N, 8.5. Synthesis of [ToAdBenz]Zr(CH2Ph)3. A solution of Zr(CH2Ph)4 (17.2 mg, 0.038 mmol) in benzene (4 mL) was added to [ToAdBenz]Tl· THF (41.0 mg, 0.038 mmol). The mixture was stirred at room temperature for 10 min, thereby resulting in the immediate deposition of thallium. The mixture was filtered and the filtrate was treated with pentane (15 mL), thereby precipitating [ToAdBenz]Zr(CH2Ph)3. The supernatant was decanted, and the product was washed with pentane (5 mL) and dried in vacuo to afford [ToAdBenz]Zr(CH2Ph)3 (27 mg, 61%) as a yellow powder. Yellow crystals of [ToAdBenz]Zr(CH2Ph)3 suitable for X-ray diffraction were obtained from a solution of [ToAdBenz]Zr(CH2Ph)3 in CH2Cl2. Anal. Calcd for [ToAdBenz]Zr(CH2Ph)3·0.4CH2Cl2: C, 71.7; H, 6.6; N, 6.9. Found: C, 71.7; H, 7.1; N, 6.9. Synthesis of [BoMeBenz]2Zr(CH2Ph)2. A mixture of Zr(CH2Ph)4 (17.8 mg, 0.04 mmol) and [BoMeBenz]Tl (40.0 mg, 0.08 mmol) was treated with benzene (2 mL). The mixture was stirred at room temperature for 5 min, thereby resulting in the immediate deposition of thallium. The mixture was filtered and the filtrate was treated with pentane (20 mL), thereby precipitating [BoMeBenz]2Zr(CH2Ph)2. The supernatant was decanted, and the product was washed with pentane (10 mL) and dried in vacuo to afford [BoMeBenz]2Zr(CH2Ph)2 (18.0 mg, 52%) as a yellow powder. Yellow crystals of [BoMeBenz]2Zr(CH2Ph)2 suitable for X-ray diffraction were obtained from the slow diffusion of pentane into a toluene solution of [BoMeBenz]2Zr(CH2Ph)2 at −15 °C. Synthesis of [BoAdBenz]2Zr(CH2Ph)2. A mixture of Zr(CH2Ph)4 (17.3 mg, 0.04 mmol) and [BoAdBenz]Tl·0.5THF (60 mg, 0.08 mmol) was treated with toluene (4 mL). The mixture was stirred at room temperature for 10 min, thereby resulting in the immediate deposition of thallium. The mixture was filtered and the filtrate was treated with pentane (20 mL), thereby precipitating [BoAdBenz]2Zr(CH2Ph)2. The supernatant was decanted, and the product was washed with G

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Inorganic Chemistry



pentane (10 mL) and dried in vacuo to afford [BoAdBenz]2Zr(CH2Ph)2 (25.0 mg, 48%) as a yellow powder.

ACKNOWLEDGMENTS We thank the U.S. Department of Energy, Office of Basic Energy Sciences (Grant DE-FG02-93ER14339), for support of this research.57 A.A.-H. thanks the Government of Saudi Arabia for a scholarship.

t

Synthesis of {[ToBu Benz]Zr(CH2Ph)2(OEt2)}[B(C6F5)4]. A mixture t

of [ToBu Benz]Zr(CH2Ph)3 (29.5 mg, 0.03 mmol) and [PhNHMe2][B(C6F5)4] (25.0 mg, 0.03 mmol) was treated with Et2O (4 mL), stirred at room temperature for several minutes, and filtered. The mixture t was filtered, and yellow crystals of {[ToBu Benz]Zr(CH2Ph)2(OEt2)[B(C6F5)4] were obtained by the diffusion of pentane into the filtrate at t −15 °C. The crystals of {[ToBu Benz]Zr(CH2Ph)2(OEt2)[B(C6F5)4] were isolated, washed with pentane (3 mL), and dried (15.0 mg, 30%). t Generation and Ethylene Polymerization Activity of {[ToBu Benz]t Zr(CH2Ph)2}[B(C6F5)4]. A solution of [ToBu Benz]Zr(CH2Ph)3 (10 mg, 0.01 mmol) in C6D6 (2 mL) was added to [PhNHMe2][B(C6F5)4] (8.5 mg, 0.01 mmol), and the combination was mixed for several minutes, filtered, and analyzed by 1H NMR spectroscopy, thereby t demonstrating the formation of a mixture of {[ToBu Benz]Zr(CH2Ph)2}+, Me2NPh, and PhCH3. Benzene (4 mL) was added to the above solution, which was degassed and treated with ethylene (1 atm). The reaction mixture was stirred at room temperature for 10 min while maintaining an ethylene pressure of 1 atm. After this period, the mixture was quenched with methanol (5 mL), followed by dilute HClaq (1 M, 20 mL). The polymer was collected by filtration, washed with methanol (10 mL), and then dried in vacuo to a constant weight (80 mg, corresponding to an activity of 48 kg of PE [mol of Zr]−1 [h]−1 [atm of C2H4]−1). Generation and Ethylene Polymerization Activity of {[ToAdBenz]Zr(CH2Ph)2}[B(C6F5)4]. A solution of [ToAdBenz]Zr(CH2Ph)3 (7 mg, 0.006 mmol) in C6D6 (2 mL) was added to [PhNHMe2][B(C6F5)4] (4.7 mg, 0.006 mmol), and the combination was mixed for several minutes, filtered, and analyzed by 1H NMR spectroscopy, thereby demonstrating the formation of a mixture of {[ToAdBenz]Zr(CH2Ph)2}+, Me2NPh, and PhCH3. Benzene (4 mL) was added to the above solution, which was degassed and treated with ethylene (1 atm). The reaction mixture was stirred at room temperature for 10 min while maintaining an ethylene pressure of 1 atm. After this period, the mixture was quenched with methanol (5 mL), followed by dilute HClaq (1 M, 20 mL). The polymer was collected by filtration, washed with methanol (10 mL), and dried in vacuo to a constant weight (160 mg, corresponding to an activity of 160 kg of PE [mol of Zr]−1 [h]−1 [atm of C2H4]−1).





REFERENCES

(1) (a) Chen, E. Y.; Marks, T. J. Cocatalysts for metal-catalyzed olefin polymerization: Activators, activation processes, and structure-activity relationships. Chem. Rev. 2000, 100, 1391−1434. (b) Delferro, M.; Marks, T. J. Multinuclear olefin polymerization catalysts. Chem. Rev. 2011, 111, 2450−2485. (c) Collins, R. A.; Russell, A. F.; Mountford, P. Group 4 metal complexes for homogeneous olefin polymerisation: a short tutorial review. Appl. Petrochem. Res. 2015, 5, 153−171. (d) Bochmann, M. The chemistry of catalyst activation: The case of group 4 polymerization catalysts. Organometallics 2010, 29, 4711−4740. (2) In addition to the use of simple cyclopentadienyl ligands, variants that contain pendent donors have also received much attention. For example, see: (a) Shapiro, P. J.; Bunel, E.; Schaefer, W. P.; Bercaw, J. E. [((η5-C5Me4)Me2Si(η1-NCMe3))(PMe3)ScH]2: A unique example of a single-component α-olefin polymerization Catalyst. Organometallics 1990, 9, 867−869. (b) Doufou, P.; Abboud, K. A.; Boncella, J. M. Synthesis and characterization of organozirconium complexes bearing a new trianionic multidentate cyclopentadienyl ligand by use of amine elimination. J. Organomet. Chem. 2000, 603, 213−219. (c) Wang, C.; Erker, G.; Kehr, G.; Wedeking, K.; Frohlich, R. Synthesis, structural features, and formation of organometallic derivates of C1-bridged Cp/ amido titanium and zirconium “CpCN-constrained geometry” systems. Organometallics 2005, 24, 4760−4773. (d) Okuda, J. Bifunctional cyclopentadienyl ligands in organotransition metal chemistry. Comments Inorg. Chem. 1994, 16, 185−205. (e) Braunschweig, H.; Breitling, F. M. Constrained geometry complexes − Synthesis and applications. Coord. Chem. Rev. 2006, 250, 2691−2720. (3) For the classification of ligands as L, X or Z, see: (a) Green, M. L. H. A new approach to the formal classification of covalent compounds. J. Organomet. Chem. 1995, 500, 127−148. (b) Parkin, G. Classification of organotransition metal compounds. Comprehensive Organometallic Chemistry III; Crabtree, R. H., Mingos, D. M. P., Eds.; Elsevier: Oxford, U.K., 2006; Vol. 1, Chapter 1. (c) Green, M. L. H.; Parkin, G. Application of the Covalent Bond Classification method for the teaching of inorganic chemistry. J. Chem. Educ. 2014, 91, 807−816. (4) (a) ScorpionatesThe Coordination Chemistry of Polypyrazolylborate Ligands; Trofimenko, S., Ed.; Imperial College Press: London, 1999. (b) Scorpionates II: Chelating Borate Ligands; Pettinari, C., Ed.; Imperial College Press: London, 2008. (c) Parkin, G. Synthetic analogues relevant to the structure and function of zinc enzymes. Chem. Rev. 2004, 104, 699−767. (d) Parkin, G. The bioinorganic chemistry of zinc: Synthetic analogues of zinc enzymes that feature tripodal ligands. Chem. Commun. 2000, 1971−1985. (e) Parkin, G. Synthetic analogues of zinc enzymes. In Metal Ions in Biological Systems; Sigel, A., Sigel, H., Eds.; Marcel Dekker: New York, 2001; Vol. 38, Chapter 14, pp 411−460. (f) Parkin, G. Alkyl, hydride, and hydroxide derivatives of the s- and pblock elements supported by poly(pyrazolyl)borato ligation: Models for carbonic anhydrase, receptors for anions, and the study of controlled crystallographic disorder. Adv. Inorg. Chem. 1995, 42, 291−393. (g) Parkin, G. Effect of tris(pyrazolyl)hydroborato ligation. In Handbook of Grignard Reagents; Silverman, G. S., Rakita, P. E., Eds.; Marcel Dekker: New York, 1996. (h) Santini, C.; Pellei, M.; Lobbia, G. G.; Papini, G. Synthesis and properties of poly(pyrazolyl)borate and related boron-centered scorpionate ligands. Part A: Pyrazole-based systems. Mini-Rev. Org. Chem. 2010, 7, 84−124. (5) For other examples of L2X ligands with different donor arrays, see: (a) Rabinovich, D. Synthetic bioinorganic chemistry: Scorpionates turn 50. Struct. Bonding 2016, 172, 139−157. (b) Spicer, M. D.; Reglinski, J. Soft scorpionate ligands based on imidazole-2-thione donors. Eur. J. Inorg. Chem. 2009, 1553−1574. (c) Smith, J. M. Strongly donating scorpionate ligands. Comments Inorg. Chem. 2008, 29, 189−233. (d) Soares, L. F.; Silva, R. M. Hydrotris(methimazolyl)borate. Inorg.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b02832. Spectroscopic data and Cartesian coordinates for geometryoptimized structures (PDF) Accession Codes

CCDC 1584045−1584052 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Gerard Parkin: 0000-0003-1925-0547 Notes

The authors declare no competing financial interest. H

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Syn th. 2002 , 33, 199−202. (e) R abinovic h, D. Poly(mercaptoimidazolyl)borate complexes of cadmium and mercury. Struct. Bonding 2006, 120, 143−162. (f) Parkin, G. Applications of tripodal [S3] and [Se3] L2X donor ligands to zinc, cadmium and mercury chemistry: organometallic and bioinorganic perspectives. New J. Chem. 2007, 31, 1996−2014. (6) For examples of X3 [O3] donors, see: (a) Dinger, M. B.; Scott, M. J. Extended structures built on a triphenoxymethane platform − C3symmetric, conformational mimics of calix[n]arenes. Eur. J. Org. Chem. 2000, 2467−2478. (b) Dinger, M. B.; Scott, M. J. Alkali salts of C3symmetric, linked aryloxides: Selective binding of substrates with metal aggregates. Inorg. Chem. 2000, 39, 1238−1254. (c) Dinger, M. B.; Scott, M. J. Sterically crowded, rigid, C3 symmetric phosphites: Synthesis, structure, and preparation of metal complexes. Inorg. Chem. 2001, 40, 856−864. (d) Akagi, F.; Matsuo, T.; Kawaguchi, H. Titanium and zirconium complexes of preorganized tripodal triaryloxide ligands. J. Am. Chem. Soc. 2005, 127, 11936−11937. (e) Akagi, F.; Matsuo, T.; Kawaguchi, H. Dinitrogen cleavage by a diniobium tetrahydride complex: Formation of a nitride and its conversion into imide species. Angew. Chem., Int. Ed. 2007, 46, 8778−8781. (f) Chaplin, A. B.; Harrison, J. A.; Nielson, A. J.; Shen, C. H.; Waters, J. M. A case for linear agostic interactions: Identification by DFT calculation in a complex of Ta containing a uniquely caged triphenylmethyl C-H hydrogen. Dalton Trans. 2004, 2643−2648. (g) Feher, F. J.; Budzichowski, T. A. Silasesquioxanes as ligands in inorganic and organometallic chemistry. Polyhedron 1995, 14, 3239−3253. (h) Levitskii, M. M.; Smirnov, V. V.; Zavin, B. G.; Bilyachenko, A. N.; Rabkina, A. Y. Metalasiloxanes: New structure formation methods and catalytic properties. Kinet. Catal. 2009, 50, 490−507. (7) For examples of L2X [O3] donors, see: (a) Kläui, W.; Asbahr, H. O.; Schramm, G.; Englert, U. Tris-chelating oxygen ligands: New synthetic routes to sterically demanding ligands. Chem. Ber. 1997, 130, 1223− 1229. (b) Leung, W.-H.; Zhang, Q.-F.; Yi, X.-Y. Recent developments in the coordination and organometallic chemistry of Kläui oxygen tripodal ligands. Coord. Chem. Rev. 2007, 251, 2266−2279. (c) Kläui, W. The coordination chemistry and organometallic chemistry of tridentate oxygen ligands with π-donor properties. Angew. Chem., Int. Ed. Engl. 1990, 29, 627−637. (8) For non-tripodal [O3] donors, see: Matsuo, T.; Kawaguchi, H. Anisole-diphenoxide ligands and their zirconium dichloride and dialkyl complexes. Inorg. Chem. 2007, 46, 8426−8434. (9) For tetradentate LX3 (a−d), L3X (e), and X4 (d and f−h) [O3] donor ligands, see: (a) Verkade, J. G. Main group atranes: chemical and structural features. Coord. Chem. Rev. 1994, 137, 233−295. (b) Verkade, J. G. Atranes: New examples with unexpected properties. Acc. Chem. Res. 1993, 26, 483−489. (c) Kuzu, I.; Krummenacher, I.; Meyer, J.; Armbruster, F.; Breher, F. Multidentate ligand systems featuring dual functionality. Dalton Trans. 2008, 5836−5865. (d) Akagi, F.; Matsuo, T.; Kawaguchi, H. Titanium and zirconium complexes of preorganized tripodal triaryloxide ligands. J. Am. Chem. Soc. 2005, 127, 11936−11937. (e) Al-Harbi, A.; Rong, Y.; Parkin, G. Synthesis and structural characterization of tris(2-pyridonyl)methyl complexes of zinc and thallium: a new class of metallacarbatranes and a monovalent thallium alkyl compound. Dalton Trans. 2013, 42, 14053−14057. (f) Kobayashi, J.; Goto, K.; Kawashima, T.; Schmidt, M. W.; Nagase, S. Synthesis, structure, and bonding properties of 5-carbaphosphatranes: A new class of main group atrane. J. Am. Chem. Soc. 2002, 124, 3703−3712. (g) Nakanishi, Y.; Ishida, Y.; Kawaguchi, H. Synthesis and reactions of a zirconium naphthalene complex bearing a tetraanionic C-capped triaryloxide ligand. Dalton Trans. 2016, 45, 15879−15885. (h) Nakanishi, Y.; Ishida, Y.; Kawaguchi, H. Zirconium hydride complex supported by a tetradentate carbon-centered tripodal tris(aryloxide) ligand: Synthesis, structure, and reactivity. Inorg. Chem. 2016, 55, 3967− 3973. (10) Coperet, C.; Comas-Vives, A.; Conley, M. P.; Estes, D. P.; Fedorov, A.; Mougel, V.; Nagae, H.; Núñez-Zarur, F.; Zhizhko, P. A. Surface organometallic and coordination chemistry toward single-site heterogeneous catalysts: Strategies, methods, structures, and activities. Chem. Rev. 2016, 116, 323−421.

(11) (a) Limberg, C. Calixarene-based oxovanadium complexes as molecular models for catalytically active surface species and homogeneous catalysts. Eur. J. Inorg. Chem. 2007, 3303−3314. (b) Quadrelli, E. A.; Basset, J. M. On silsesquioxanes’ accuracy as molecular models for silica-grafted complexes in heterogeneous catalysis. Coord. Chem. Rev. 2010, 254, 707−728. (c) Floriani, C. Metal reactivity on oxo surfaces modeled by calix[4]arenes: Metaldriven reactions in an oxo-quasiplanar environment. Chem. - Eur. J. 1999, 5, 19−23. (d) Floriani, C.; Floriani-Moro, R. The M-C bond functionalities bonded to an oxo-surface modeled by calix[4]arenes. Adv. Organomet. Chem. 2001, 47, 167−233. (e) Giannini, L.; Solari, E.; Zanotti-Gerosa, A.; Floriani, C.; Chiesi-Villa, A.; Rizzoli, C. The organometallic chemistry of zirconium on an oxo surface provided by ptert-butylcalix[4]arene. Angew. Chem., Int. Ed. Engl. 1996, 35, 85−87. (f) Redshaw, C. Metallocalixarene catalysts: α-olefin polymerization and ROP of cyclic esters. Dalton Trans. 2016, 45, 9018−9030. (g) Espinas, J.; Darbost, U.; Pelletier, J.; Jeanneau, E.; Duchamp, C.; Bayard, F.; Boyron, O.; Broyer, J. P.; Thivolle-Cazat, J.; Basset, J. M.; Taoufik, M.; Bonnamour, I. Titanacalixarenes in homogeneous catalysis: Synthesis, conformation and catalytic activity in ethylene polymerization. Eur. J. Inorg. Chem. 2010, 1349−1359. (12) (a) Salinier, V.; Corker, J. M.; Lefebvre, F.; Bayard, F.; Dufaud, V.; Basset, J. M. Silica-supported zirconium complexes and their polyoligosilsesquioxane analogues in the transesterification of acrylates: Part 1. Synthesis and characterization. Adv. Synth. Catal. 2009, 351, 2155−2167. (b) Salinier, V.; Niccolai, G. P.; Dufaud, V.; Basset, J. M. Silica-supported zirconium complexes and their polyoligosilsesquioxane analogues in the transesterification of acrylates: Part 2. Activity, recycling and regeneration. Adv. Synth. Catal. 2009, 351, 2168−2177. (c) Eedugurala, N.; Wang, Z. R.; Chaudhary, U.; Nelson, N.; Kandel, K.; Kobayashi, T.; Slowing, I. I.; Pruski, M.; Sadow, A. D. Mesoporous silica-supported amidozirconium-catalyzed carbonyl hydroboration. ACS Catal. 2015, 5, 7399−7414. (d) Franceschini, F. C.; Tavares, T. T. R.; Bianchini, D.; Alves, M. C. M.; Ferreira, M. L.; dos Santos, J. H. Z. Characterization and evaluation of supported rac-dimethylsilylenebis(indenyl)zirconium dichloride on ethylene polymerization. J. Appl. Polym. Sci. 2009, 112, 563−571. (13) For examples of cyclopentadienylzirconium compounds supported on oxygen surfaces, see: (a) Jezequel, M.; Dufaud, V.; RuizGarcia, M. J.; Carrillo-Hermosilla, F.; Neugebauer, U.; Niccolai, G. P.; Lefebvre, F.; Bayard, F.; Corker, J.; Fiddy, S.; Evans, J.; Broyer, J. P.; Malinge, J.; Basset, J. M. Supported metallocene catalysts by surface organometallic chemistry. Synthesis, characterization, and reactivity in ethylene polymerization of oxide-supported mono- and biscyclopentadienyl zirconium alkyl complexes: Establishment of structure/reactivity relationships. J. Am. Chem. Soc. 2001, 123, 3520−3540. (b) Bashir, M. A.; Vancompernolle, T.; Gauvin, R. M.; Delevoye, L.; Merle, N.; Monteil, V.; Taoufik, M.; McKenna, T. F. L.; Boisson, C. Silica/MAO/ (n-BuCp)2ZrCl2 catalyst: Effect of support dehydroxylation temperature on the grafting of MAO and ethylene polymerization. Catal. Sci. Technol. 2016, 6, 2962−2974. (c) Popoff, N.; Gauvin, R. M.; De Mallmann, A.; Taoufik, M. On the fate of silica-supported halfmetallocene cations: Elucidating a catalyst’s deactivation pathways. Organometallics 2012, 31, 4763−4768. (d) Fisch, A. G.; Cardozo, N. S. M.; Secchi, A. R.; dos Santos, J. H. Z. Supported metallocene catalysts for polyolefin production: Review of the immobilization strategies. Quim. Nova 2011, 34, 646−657. (e) O’Brien, S.; Tudor, J.; Maschmeyer, T.; O’Hare, D. EXAFS analysis of a chiral alkene polymerisation catalyst incorporated in the mesoporous silicate MCM-41. Chem. Commun. 1997, 1905−1906. (f) Franceschini, F. C.; Tavares, T. T. R.; Bianchini, D.; Alves, M. C. M.; Ferreira, M. L.; dos Santos, J. H. Z. Characterization and evaluation of supported rac-dimethylsilylenebis(indenyl)zirconium dichloride on ethylene polymerization. J. Appl. Polym. Sci. 2009, 112, 563−571. (14) (a) Al-Harbi, A.; Sattler, W.; Sattler, A.; Parkin, G. Synthesis and structural characterization of tris(2-oxo-1-t-butylimidazolyl) and tris(2oxo-1-methylbenzimidazolyl)hydroborato complexes: A new class of tripodal oxygen donor ligand. Chem. Commun. 2011, 47, 3123−3125. (b) Al-Harbi, A.; Rong, Y.; Parkin, G. Synthesis and structural I

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

(27) Searches of the CSD were performed with version 5.38. See: Groom, C. R.; Bruno, I. J.; Lightfoot, M. P.; Ward, S. C. The Cambridge Structural Database. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 2016, 72, 171−179. (28) (a) Oliva, L.; Longo, P.; Pellecchia, C. Isotactic polypropylene by polymerization of propene in the presence of some achiral soluble transition-metal compounds and methylaluminoxane. Makromol. Chem., Rapid Commun. 1988, 9, 51−55. (b) Zambelli, A.; Oliva, L.; Pellecchia, C. Soluble catalysts for syndiotactic polymerization of styrene. Macromolecules 1989, 22, 2129−2130. (c) Longo, P.; Grassi, A.; Oliva, L. Copolymerization of styrene and ethylene in the presence of different syndiospecific catalysts. Makromol. Chem. 1990, 191, 2387−2396. (d) Bochmann, M.; Lancaster, S. J. Base-free cationic zirconium benzyl complexes as highly-active polymerization catalysts. Organometallics 1993, 12, 633−640. (e) Long, R. J.; Gibson, V. C.; White, A. J. P. Group 4 metal olefin polymerization catalysts stabilized by bidentate O,P ligands. Organometallics 2008, 27, 235−245. (f) Thorn, M. G.; Etheridge, Z. C.; Fanwick, P. E.; Rothwell, I. P. Cationic group 4 metal alkyl compounds containing aryloxide ligation: Synthesis, structure, reactivity and polymerization studies. J. Organomet. Chem. 1999, 591, 148−162. (g) Thorn, M. G.; Etheridge, Z. C.; Fanwick, P. E.; Rothwell, I. P. Cationic group 4 metal alkyl compounds containing oarylphenoxide ligation. Organometallics 1998, 17, 3636−3638. (29) Popoff, N.; Macqueron, B.; Sayhoun, W.; Espinas, J.; Pelletier, J.; Boyron, O.; Boisson, C.; Merle, N.; Szeto, K. C.; Gauvin, R. M.; De Mallmann, A.; Taoufik, M. Well-defined silica-supported zirconiumbenzyl cationic species: Improved heterogenization of single-site polymerization catalysts. Eur. J. Inorg. Chem. 2014, 888−895. (30) Lee, H.; Nienkemper, K.; Jordan, R. F. Synthesis and reactivity of a sterically crowded tris(pyrazolyl)borate hafnium benzyl complex. Organometallics 2008, 27, 5075−5081. (31) See, for example, refs 4f and 4g. (a) Han, R.; Bachrach, M.; Parkin, G. Competitive alkyl and halide metathesis in the reactions of Grignard reagents with Tl{η3-HB(3-Butpz)3}. Polyhedron 1990, 9, 1775−1778. (b) Han, R.; Parkin, G. (Tris(pyrazolyl)hydroborato)magnesium alkyl derivatives: Synthetic and structural studies. Organometallics 1991, 10, 1010−1020. (c) Han, R.; Parkin, G. Synthesis, structure, and reactivity of {η3-HB(3-Butpz)3}BeCH3, a terminal alkyl complex supported by tris(3-tert-butylpyrazolyl)hydroborato ligation. Inorg. Chem. 1993, 32, 4968−4970. (d) Looney, A.; Saleh, A.; Zhang, Y. H.; Parkin, G. Tris(pyrazolyl)hydroborato complexes of cadmium: A bidentate nitrate derivative and its relevance to carbonic anhydrase activity. Inorg. Chem. 1994, 33, 1158−1164. (e) Looney, A.; Han, R.; Gorrell, I. B.; Cornebise, M.; Yoon, K.; Parkin, G.; Rheingold, A. L. Monomeric alkyl and hydride derivatives of zinc supported by poly(pyrazolyl)hydroborato ligation: Synthetic, structural, and reactivity studies. Organometallics 1995, 14, 274−288. (32) For related examples, see: (a) Kimblin, C.; Bridgewater, B. M.; Hascall, T.; Parkin, G. The synthesis and structural characterization of bis(mercaptoimidazolyl)(pyrazolyl)hydroborato and tris(mercaptoimidazolyl)hydroborato complexes of thallium(I) and thallium(III). J. Chem. Soc., Dalton Trans. 2000, 1267−1274. (b) Melnick, J. G.; Parkin, G. Methyl and arylchalcogenolate complexes of cadmium in a sulfur rich coordination environment: Syntheses and structural characterization of the tris(2-mercapto-1-tert-butylimidazolyl)hydroborato cadmi-

characterization of bis(2-oxoimidazolyl)hydroborato complexes: A new class of bidentate oxygen donor ligand. Inorg. Chem. 2013, 52, 10226− 10228. (c) Al-Harbi, A.; Kriegel, B.; Gulati, S.; Hammond, M. J.; Parkin, G. Bis- and Tris(2-oxobenzimidazolyl)hydroborato complexes of sodium and thallium: New classes of bidentate and tridentate oxygen donor ligands. Inorg. Chem. 2017, 56, 15271−15284. (15) The synthesis of the chloride derivatives reported herein has been previously communicated. See ref 14a. (16) Ward, T. R.; Duclos, S.; Therrien, B.; Schenk, K. Coordination properties of Klaui’s tripodal oxygen donor toward zirconium(IV). Organometallics 1998, 17, 2490−2494. (17) For other zirconium compounds that feature [CpCo{P(O)(OR)2}3] ligands, see: (a) Yi, X.-Y.; Zhang, Q.-F.; Lam, T. C. H.; Chan, E. Y. Y.; Williams, I. D.; Leung, W.-H. Phosphato, chromato, and perrhenato complexes of titanium(IV) and zirconium(IV) containing Kläui’s tripodal ligand. Inorg. Chem. 2006, 45, 328−335. (b) Yi, X.-Y.; Zhang, Q.-F.; Williams, I. D.; Leung, W.-H. Synthesis and structures of zirconium(IV) hydrogensulfato and carboxylato complexes with Klaui’s oxygen tripodal ligand. J. Organomet. Chem. 2009, 694, 4256−4260. (c) Yi, X.-Y.; Lam, T. C. H.; Williams, I. D.; Leung, W.-H. Hydrolysis of bis(p-nitrophenyl)phosphate by tetravalent metal complexes with Kläui’s oxygen tripodal ligand. Inorg. Chem. 2010, 49, 2232−2238. (d) Yi, X.-Y.; Wang, G.-C.; Ip, H.-F.; Wong, W.-Y.; Chen, L.; Sung, H. H.-Y.; Williams, I. D.; Leung, W.-H. Oxidation of alkylbenzenes with cerium complexes containing a tripodal oxygen ligand. Eur. J. Inorg. Chem. 2014, 6097−6103. (18) (a) Repo, T.; Klinga, M.; Mutikainen, I.; Su, Y. C.; Leskelä, M.; Polamo, M. Low-temperature crystal structures of dichlorobis(η5indenyl)zirconium(IV) at 153 K and dichlorobis(η5-cyclopentadienyl)zirconium(IV) at 193 K. Acta Chem. Scand. 1996, 50, 1116−1120. (b) Corey, J. Y.; Zhu, X.-H.; Brammer, L.; Rath, N. P. Zirconocene dichloride. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1995, 51, 565−567. (19) Buccella, D.; Shultz, A.; Melnick, J. G.; Konopka, F.; Parkin, G. Cp[TmMe]ZrCl2, a tris(2-mercapto-1-methylimidazolyl)hydroborato complex of zirconium and a new type of precatalyst for olefin polymerization. Organometallics 2006, 25, 5496−5499. (20) A 2:1 pattern is also maintained at −73 °C. (21) Mislow, K.; Gust, D.; Finocchiaro, P.; Boettcher, R. J. Stereochemical correspondence among molecular propellers. Top. Curr. Chem. 1974, 47, 1−28. (22) For another example of more facile enantiomer interconversion for an oxygen-donor ligand than a sulfur-donor ligand, see: Ghosh, P.; Parkin, G. Modeling the Active Sites of Bacteriophage T7 Lysozyme, Bovine 5-Aminolevulinate Dehydratase, and Peptide Deformylase: Synthesis and Structural Characterization of a Bis(pyrazolyl)(thioalkoxy)hydroborato Zinc Complex, [(Ph2CHS)BpBut,Pri]ZnI. Chem. Commun. 1998, 413−414. (23) (a) Foreman, M. R. St.-J.; Hill, A. F.; White, A. J. P.; Williams, D. J. Hydrotris(methimazolyl)borato alkylidyne complexes of tungsten. Organometallics 2003, 22, 3831−3840. (b) Bailey, P. J.; Dawson, A.; McCormack, C.; Moggach, S. A.; Oswald, I. D. H.; Parsons, S.; Rankin, D. W. H.; Turner, A. Barriers to racemization in C3-symmetric complexes containing the hydrotris(2-mercapto-1-ethylimidazolyl)borate (TmEt) Ligand. Inorg. Chem. 2005, 44, 8884−8898. (c) Puchta, R.; van Eikema Hommes, N. V.; Meier, R.; van Eldik, R. A new enantiomerization mechanism for tripodal penta-coordinate Zn-II(nta) complexes. Theoretical clarification of the observed 1H NMR spectrum. Dalton Trans. 2006, 3392−3395. (24) Luo, Y.-R. Comprehensive Handbook of Chemical Bond Energies; CRC Press, 2007, Chapter 14. (25) (a) Kutzelnigg, W. Chemical bonding in higher main group elements. Angew. Chem., Int. Ed. Engl. 1984, 23, 272−295. (b) Cherry, W.; Epiotis, N. The inversion barrier in AH3 molecules. J. Am. Chem. Soc. 1976, 98, 1135−1140. (26) Two-coordinate oxygen compounds typically exhibit larger bond angles than do two-coordinate sulfur compounds. For example, the mean X−O−X angle for two-coordinate compounds listed in the CSD27 is 117.0°, compared to a value of 97.7° for X−S−X compounds.

t

t

um complexes [ToBu ]CdMe, and [ToBu ]CdEAr (E = O, S, Se, Te) and analysis of the bonding in chalcogenolate compounds. Dalton Trans. 2006, 4207−4210. (c) Melnick, J. G.; Docrat, A.; Parkin, G. Methyl, hydrochalcogenido, and phenylchalcogenolate complexes of zinc in a sulfur rich coordination environment: syntheses and structural characterization of the tris(2-mercapto-1-tert-butylimidazolyl)hydroboratozinc t

t

t

complexes [ToBu ]ZnMe, [ToBu ]ZnEH (E = S, Se) and [ToBu ]ZnEPh (E = O, S, Se, Te). Chem. Commun. 2004, 2870−2871. (d) Kumar, M.; Papish, E. T.; Zeller, M.; Hunter, A. D. New alkylzinc complexes with bulky tris(triazolyl)borate ligands: Surprising water stability and reactivity. Dalton Trans. 2010, 39, 59−61. (e) Reference 30. (33) (a) Janiak, C. (Organo)thallium (I) and (II) chemistry: Syntheses, structures, properties and applications of subvalent thallium J

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry complexes with alkyl, cyclopentadienyl, arene or hydrotris(pyrazolyl)borate ligands. Coord. Chem. Rev. 1997, 163, 107−216. (b) McKillop, A.; Elsom, L. F.; Taylor, E. C. Thallium in organic synthesis.3. Coupling of aryl and alkyl Grignard reagents. J. Am. Chem. Soc. 1968, 90, 2423−2424. (c) Gilman, H.; Jones, R. G. Trimethylthallium. J. Am. Chem. Soc. 1946, 68, 517−520. (d) Lee, A. G. Organothallium chemistry. Q. Rev., Chem. Soc. 1970, 24, 310−329. (e) Schwerdtfeger, P.; Boyd, P. D. W.; Bowmaker, G. A.; Mack, H. G.; Oberhammer, H. Theoretical studies on the stability of Tl-C σ-bonds in aliphatic organothallium compounds. J. Am. Chem. Soc. 1989, 111, 15−23. (34) Rong, Y.; Sambade, D.; Parkin, G. Molecular structures of tris(1tert-butyl-2-mercaptoimidazolyl)hydroborate complexes of titanium, zirconium and hafnium. Acta Crystallogr., Sect. C: Struct. Chem. 2016, 72, 806−812. (35) For examples of zirconium benzyl compounds with oxygen donors in the coordination environment, see ref 9d. (a) Kiesewetter, E. T.; Randoll, S.; Radlauer, M.; Waymouth, R. M. Stereospecific octahedral group 4 bis(phenolate) ether complexes for olefin polymerization. J. Am. Chem. Soc. 2010, 132, 5566−5567. (b) Matsuo, T.; Kawaguchi, H. A synthetic cycle for H2/CO activation and allene synthesis using recyclable zirconium complexes. J. Am. Chem. Soc. 2005, 127, 17198−17199. (c) Giannini, L.; Solari, E.; Zanotti-Gerosa, A.; Floriani, C.; Chiesi-Villa, A.; Rizzoli, C. The organometallic chemistry of zirconium on an oxo surface provided by p-tert-butylcalix[4]arene. Angew. Chem., Int. Ed. Engl. 1996, 35, 85−87. (d) Groysman, S.; Goldberg, I.; Kol, M.; Genizi, E.; Goldschmidt, Z. From THF to furan: Activity tuning and mechanistic insight via sidearm donor replacement in group IV amine bis(phenolate) polymerization catalysts. Organometallics 2003, 22, 3013−3015. (e) Watanabe, T.; Ishida, Y.; Matsuo, T.; Kawaguchi, H. Syntheses and structures of zirconium(IV) complexes supported by 2,6-di-adamantylaryloxide ligands and formation of arenebridged dizirconium complexes with an inverse sandwich structure. Dalton Trans. 2010, 39, 484−491. (f) Giannini, L.; Caselli, A.; Solari, E.; Floriani, C.; Chiesi-Villa, A.; Rizzoli, C.; Re, N.; Sgamellotti, A. Organometallic reactivity on a calix[4]arene oxo surface. The stepwise migratory insertion of carbon monoxide and isocyanides into zirconiumcarbon bonds anchored to a calix[4]arene moiety. J. Am. Chem. Soc. 1997, 119, 9709−9719. (g) Giannini, L.; Caselli, A.; Solari, E.; Floriani, C.; Chiesi-Villa, A.; Rizzoli, C.; Re, N.; Sgamellotti, A. Organometallic reactivity on a calix[4]arene oxo surface. Synthesis and rearrangement of Zr-C functionalities anchored to a calix[4]arene moiety. J. Am. Chem. Soc. 1997, 119, 9198−9210. (h) Radlauer, M. R.; Agapie, T. Bimetallic zirconium amine bis(phenolate) polymerization catalysts: Enhanced activity and tacticity control for polyolefin synthesis. Organometallics 2014, 33, 3247−3250. (i) Shao, P. C.; Gendron, R. A. L.; Berg, D. J.; Bushnell, G. W. Dibenzylzirconium complexes of chelating aminodiolates. Synthesis, structural studies, thermal stability, and insertion chemistry. Organometallics 2000, 19, 509−520. (j) Groysman, S.; Goldberg, I.; Kol, M.; Genizi, E.; Goldschmidt, Z. Group IV complexes of an amine bis(phenolate) ligand featuring a THF sidearm donor: from highly active to living polymerization catalysts of 1-hexene. Inorg. Chim. Acta 2003, 345, 137−144. (k) Nakata, N.; Toda, T.; Saito, Y.; Watanabe, T.; Ishii, A. Highly active and isospecific styrene polymerization catalyzed by zirconium complexes bearing aryl-substituted [OSSO]type bis(phenolate) ligands. Polymers 2016, 8, 31. (36) Rong, Y.; Al-Harbi, A.; Parkin, G. Highly variable Zr−CH2−Ph bond angles in tetrabenzylzirconium: Analysis of benzyl ligand coordination modes. Organometallics 2012, 31, 8208−8217. (37) Edwards, P. G.; Andersen, R. A.; Zalkin, A. Preparation of tetraalkyl phosphine complexes of the f-block metals. Crystal structure of Th(CH 2 Ph) 4 (Me 2 PCH 2 CH 2 PMe 2 ) and U(CH 2 Ph) 3 Me(Me2PCH2CH2PMe2). Organometallics 1984, 3, 293−298. (38) For example, a recent analysis indicated that approximately 93% of benzyl ligands coordinate in an η1 manner. See ref 36. (39) Pellecchia, C.; Immirzi, A.; Pappalardo, D.; Peluso, A. A novel η7 coordination mode of a benzyl ligand in a cationic zirconium complex. Organometallics 1994, 13, 3773−3775. (40) Lee, H.; Jordan, R. F. Unusual reactivity of tris(pyrazolyl)borate zirconium. J. Am. Chem. Soc. 2005, 127, 9384−9385.

(41) Giesbrecht, G. R.; Whitener, G. D.; Arnold, J. Crystal packing forces dictate η1- versus η2-coordination of benzyl groups in [guanidinate]Zr(CH2Ph)3. Organometallics 2000, 19, 2809−2812. (42) For a recent review of crystal packing effects on bond lengths, see: Minasian, S. G.; Krinsky, J. L.; Arnold, J. Evaluating f-element bonding from structure and thermodynamics. Chem. - Eur. J. 2011, 17, 12234− 12245. (43) It should be noted that the methylene groups of the tris(pyrazolyl)hydroborato complex, [TpMe2]Zr(CH2Ph)3 also appear as a singlet,40 but this is to be expected in view of the fact that the [TpMe2] Zr moiety possesses local C3 symmetry. (44) Green, J. C.; Green, M. L. H.; Parkin, G. The occurrence and representation of three-centre two-electron bonds in covalent inorganic compounds. Chem. Commun. 2012, 48, 11481−11503. (45) For related compounds with Zr···H−B interactions, see refs 34 and 40. (a) Fryzuk, M. D.; Mylvaganam, M.; Zaworotko, M. J.; Macgillivray, L. R. Synthesis of mononuclear paramagnetic zirconium(III) derivatives. J. Am. Chem. Soc. 1993, 115, 10360−10361. (b) Kasani, A.; Gambarotta, S.; Bensimon, C. Zirconium alkyl and borohydride complexes stabilized by a sterically demanding anionic organic amide. The crystal structures of ZrMeL3 (2) and Zr(BH4)L3 (3) (L = (3,5Me2Ph)N(Ad); Ad = adamantyl). Can. J. Chem. 1997, 75, 1494−1499. (c) Liu, F.-C.; Chen, S.-C.; Lee, G.-H.; Peng, S.-M. Effect of solvent on reactions of Cp2Zr{(μ-H)2BHR}2 and Cp2ZrH{(μ-H)2BHR}2 (R = CH3, Ph) with B(C6F5)3. J. Organomet. Chem. 2007, 692, 2375−2384. (d) Ding, E. R.; Liu, F. C.; Liu, S. M.; Meyers, E. A.; Shore, S. G. Synthesis, structural characterization, and reactions of cyclic organohydroborate half-zirconocene compounds. Inorg. Chem. 2002, 41, 5329−5335. (e) Marks, T. J.; Kolb, J. R. Covalent transition-metal, lanthanide, and actinide tetrahydroborate complexes. Chem. Rev. 1977, 77, 263−293. (f) Xu, Z.; Lin, Z. Transition metal tetrahydroborato complexes: An orbital interaction analysis of their structure and bonding. Coord. Chem. Rev. 1996, 156, 139−162. (g) Besora, M.; Lledós, A. Coordination modes and hydride exchange dynamics in transition metal tetrahydroborate complexes. Struct. Bonding 2008, 130, 149−202. (46) (a) Gibson, V. C.; Spitzmesser, S. K. Advances in nonmetallocene olefin polymerization catalysis. Chem. Rev. 2003, 103, 283−315. (b) Britovsek, G. J. P.; Gibson, V. C.; Wass, D. F. The search for new-generation olefin polymerization catalysts: Life beyond metallocenes. Angew. Chem., Int. Ed. 1999, 38, 428−447. (47) (a) Peng, D. Q.; Yan, X. W.; Yu, C.; Zhang, S. W.; Li, X. F. Transition metal complexes bearing tridentate ligands for precise olefin polymerization. Polym. Chem. 2016, 7, 2601−2634. (b) Budagumpi, S.; Keri, R. S.; Biffis, A.; Patil, S. A. Olefin poly/oligomerizations by metal precatalysts bearing non-heterocyclic N-donor ligands. Appl. Catal., A 2017, 535, 32−60. (48) (a) Nakazawa, H.; Ikai, S.; Imaoka, K.; Kai, Y.; Yano, T. Polymerization of olefins with titanium and zirconium complexes containing hydrotris(pyrazolyl)borate or hydrotris(3,5dimethylpyrazolyl)borate. J. Mol. Catal. A: Chem. 1998, 132, 33−41. (b) Nienkemper, K.; Lee, H.; Jordan, R. F.; Ariafard, A.; Dang, L.; Lin, Z. Y. Synthesis of double-end-capped polyethylene by a cationic tris(pyrazolyl)borate zirconium benzyl complex. Organometallics 2008, 27, 5867−5875. (c) Karam, A.; Jimeno, M.; Lezama, J.; Catarí, E.; Figueroa, A.; de Gascue, B. R. Ethylene polymerization by TpRTi(OCH3)3−nCln complexes. J. Mol. Catal. A: Chem. 2001, 176, 65−72. (d) Karam, A.; Casas, E.; Catarí, A.; Pekerar, S.; Albornoz, A.; Méndez, B. Effect of the alkoxyl ligands on ethylene polymerization by TpTiCl2(OR) complexes. J. Mol. Catal. A: Chem. 2005, 238, 233− 240. (e) Karam, A.; Pastran, J.; Casas, E.; Mendéz, B. Styrene polymerization by TpTiCl2(OR) precatalysts. Polym. Bull. 2005, 55, 11−17. (f) Murtuza, S.; Casagrande, O. L., Jr.; Jordan, R. F. Ethylene polymerization behavior of tris(pyrazolyl)borate titanium(IV) complexes. Organometallics 2002, 21, 1882−1890. (g) Lopez-Linares, F.; Barrios, A. D.; Ortega, H.; Karam, A.; Agrifoglio, G.; Gonzalez, E. Modification of polyethylene polydispersity by blending a Ziegler-Natta catalyst with a group of IV-half metallocene or scorpionate complexes. J. Mol. Catal. A: Chem. 2002, 179, 87−92. (h) Gil, M. P.; dos Santos, J. H. Z.; Casagrande, O. L., Jr. Copolymerization of ethylene with l-hexene K

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry using sterically hindered tris(pyrazolyl)borate titanium (IV) compounds. Macromol. Chem. Phys. 2001, 202, 319−324. (i) Gil, M. P.; dos Santos, J. H. Z.; Casagrande, O. L.; Simplício, L. M. T.; da Rocha, Z. N. Spectroscopic and voltametric studies in titanium tris(pyrazolyl)borate catalysts. J. Mol. Catal. A: Chem. 2005, 238, 96−101. (j) Junges, F.; de Souza, R. F.; dos Santos, J. H. Z.; Casagrande, O. L. Ethylene polymerization using combined Ni and Ti catalysts supported in situ on MAO-modified silica. Macromol. Mater. Eng. 2005, 290, 72−77. (k) Kunrath, F. A.; Mauler, R. S.; de Souza, R. F.; Casagrande, O. L. Synthesis and properties of branched polyethylene/high-density polyethylene blends using a homogeneous binary catalyst system composed of early and late transition metal complexes. Macromol. Chem. Phys. 2002, 203, 2058−2068. (l) Asandei, A. D.; Moran, I. W. The ligand effect in Ti-mediated living radical styrene polymerizations initiated by epoxide radical ring opening. 2. Scorpionate and half-sandwich LTiCl3 complexes. J. Polym. Sci., Part A: Polym. Chem. 2005, 43, 6039−6047. (m) Kunrath, F. A.; de Souza, R. F.; Casagrande, O. L., Jr. Combination of nickel and titanium complexes containing nitrogen ligands as catalyst for polyethylene reactor blending. Macromol. Rapid Commun. 2000, 21, 277−280. (n) Janiak, C.; Lange, K. C. H.; Scharmann, T. G. Zirconiumchelate and mono-cyclopentadienyl zirconium-chelate/methylalumoxane systems as soluble Ziegler-Natta olefin polymerization catalysts. Appl. Organomet. Chem. 2000, 14, 316−324. (49) These activities are classified as moderate and good, respectively. See ref 46b. (50) As would be anticipated for a formally 10-electron species, one of

authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof.

t

the benzyl ligands of {[ToBu Benz]Zr(CH2Ph)2}+ adopts an η2coordination mode, with a Zr−CH2−Ph bond angle of 93.2°. (51) (a) McNally, J. P.; Leong, V. S.; Cooper, N. J. Cannula techniques for the manipulation of air-sensitive materials. In Experimental Organometallic Chemistry; Wayda, A. L., Darensbourg, M. Y., Eds.; American Chemical Society: Washington, DC, 1987; Chapter 2, pp 6− 23. (b) Burger, B. J.; Bercaw, J. E. Vacuum line techniques for handling air-sensitive organometallic compounds. In Experimental Organometallic Chemistry; Wayda, A. L., Darensbourg, M. Y., Eds.; American Chemical Society: Washington, DC, 1987; Chapter 4, pp 79−98. (c) Shriver, D. F.; Drezdzon, M. A. The Manipulation of Air-Sensitive Compounds, 2nd ed.; Wiley-Interscience: New York, 1986. (52) Fulmer, G. R.; Miller, A. J. M.; Sherden, N. H.; Gottlieb, H. E.; Nudelman, A.; Stoltz, B. M.; Bercaw, J. E.; Goldberg, K. I. NMR chemical shifts of trace impurities: Common laboratory solvents, organics, and gases in deuterated solvents relevant to the organometallic chemist. Organometallics 2010, 29, 2176−2179. (53) Also see the Supporting Information of ref 14a. (54) (a) Sheldrick, G. M. SHELXTL, An Integrated System for Solving, Refining, and Displaying Crystal Structures from Diffraction Data; University of Göttingen: Göttingen, Germany, 1981. (b) Sheldrick, G. M. A short history of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (c) Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (55) Also see: ANATIT, ANATOZ, and ANAWIW. (56) (a) Jaguar, version 8.9; Schrodinger, Inc.: New York, 2015. (b) Bochevarov, A. D.; Harder, E.; Hughes, T. F.; Greenwood, J. R.; Braden, D. A.; Philipp, D. M.; Rinaldo, D.; Halls, M. D.; Zhang, J.; Friesner, R. A. Jaguar: A high-performance quantum chemistry software program with strengths in life and materials sciences. Int. J. Quantum Chem. 2013, 113, 2110−2142. (57) This report was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of L

DOI: 10.1021/acs.inorgchem.7b02832 Inorg. Chem. XXXX, XXX, XXX−XXX