Synthesis of Semicrystalline Polyolefin Materials: Precision Methyl

Apr 23, 2018 - (11) We believe this π-stacking interaction persists in solution at low reaction temperature, causing the catalyst to maintain a very ...
1 downloads 3 Views 603KB Size
Subscriber access provided by Warwick University Library

Synthesis of Semicrystalline Polyolefin Materials: Precision Methyl Branching via Stereoretentive Chain Walking David N. Vaccarello, Kyle S. O'Connor, Pasquale Iacono, Jeffrey M Rose, Anna E Cherian, and Geoffrey W. Coates J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b02963 • Publication Date (Web): 23 Apr 2018 Downloaded from http://pubs.acs.org on April 23, 2018

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 5 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

Journal of the American Chemical Society

Synthesis of Semicrystalline Polyolefin Materials: Precision Methyl Branching via Stereoretentive Chain Walking David N. Vaccarello,‡ Kyle S. O’Connor,‡ Pasquale Iacono, Jeffrey M. Rose, Anna E. Cherian, and Geoffrey W. Coates* Department of Chemistry and Chemical Biology, Baker Laboratory, Cornell University, Ithaca, New York 14853-1301, United States

Supporting Information Placeholder ABSTRACT: We report the discovery of C2-symmetric nickel catalysts capable of the regio- and isoselective polymerization of 1-butene to produce isotactic 4,2-poly(1-butene), a new semi-crystalline polyolefin. The catalyst exhibits enantioface selectivities as high as 84% and the resulting polymers display melting temperatures up to 86 °C. This system marks a rare example of preserving stereochemistry through a chain walking polymerization process.

Given the massive global production of polyethylene (PE) and isotactic polypropylene (iPP) and the importance of these materials to modern society, new semicrystalline polyolefins produced from readily available, inexpensive feedstocks are of significant interest. The olefin 1-butene meets these criteria and can easily be accessed from both petroleum and biorenewable sources.1 Polymerization of 1-butene using Ziegler-Natta catalysis results in isotactic 1,2-poly(1-butene) which exhibits superior mechanical properties relative to iPP and PE.2 However, its complex crystallization hinders its use in many practical applications.3 Crystallization from the melt gives a kinetically favored polymorph, which slowly transforms over several days at room temperature to yield a more thermodynamically stable structure resulting in dimensional changes of the material. Recent advances have attempted to address this issue, potentially allowing iso-1,2-poly(1-butene) to be more widely utilized.3 Since their discovery for olefin polymerization, α-diimine nickel(II) and palladium(II) catalysts have received considerable attention.4 A unique mechanistic feature of these catalysts is their ability to undergo an isomerization event involving successive β-hydrogen eliminations followed by metal hydride reinsertions with opposite regiochemistry. Commonly known as “chain walking,” this isomerization event places the active species at various positions along the polymer backbone, allowing for the preparation of numerous polymer topologies from simple olefin feedstocks. Exhibiting control over polymer tacticity using late metal catalysts, due to chain walking, is a challenge. By nature of the chain walking mechanism, previously installed stereocenters can easily be planarized and potentially racemized during polymerization. Notable advancements have been achieved, with Brookhart and DuPont reporting the polymerization of cyclopentene using α-diimine Ni(II) and Pd(II) complexes to form partially isotactic cis-1,3-polycyclopentene.5 Takeuchi has accessed a variety of stereoregular structures through the polymerization of non-conjugated dienes and cycloolefins.6 Additionally, Sigman and coworkers identified a rare example of enantioselective chain walking in a redox-relay oxidative Heck-arylation using chiral

palladium catalysts. After arylation, the catalyst walks through a branch point and maintains a high enantiomeric ratio of the original stereocenter.7 Despite this progress, there are very few reports for the chain walking polymerization of simple linear α-olefins into precision branched, stereoregular structures. Wagener and coworkers have pursued acyclic diene metathesis as an elegant strategy for producing precision branched polyolefins of the type desired from precise chain-walking polymerization, but this method has not yet provided stereoregularity.8 We reported the polymerization of trans-2-butene into iso-1,3-poly(2butene);9 however, the catalyst system was only modestly isoselective ([mm] = 0.41–0.64) with the resulting polymers exhibiting low melting temperatures (Tm < 67 °C). We believed a highly isoselective catalyst capable of preserving the stereochemistry from the initial insertion event could produce isotactic materials with an improved melting temperature from more readily available 1-butene (Scheme 1). Herein, we report on the development of α-diimine nickel complexes for the polymerization of 1-butene (Figure 1). To the best of our knowledge, there are no reports of catalysts that can perform both isoselective 1alkene insertion and stereoretentive chain walking polymerization in tandem.

Scheme 1. 4,2-Polymerization of 1-Butene via Stereoretentive Chain Walking. Ar R'

1-butene

4,2-poly(1-butene) 4

R

N Ni N Br 2 R' Ar + [Al(Me)O] x cocat.

+

LNi P

R

+

P

β-H

β-H

P +

NiL H

+

LNi

ins.

ins.

LNi

1

LNi

P + NiL

H

P LNi

+

H

2

n

+

stereoretentive chain walking

P

3

P

ins. β-H

P +

L Ni H

Application of complex 1 activated with methylaluminoxane (MAO) for the polymerization of 1-butene resulted in amorphous polymer (Table 1, entry 1).10 Analysis of the material by 13C NMR spectroscopy revealed a microstructure composed primarily of 4,2poly(1-butene) (Scheme 1). However, a lack of stereocontrol as evidenced by a low [mm] triad value resulted in a lack of crystallinity (Table 1, entry 1). We hypothesized that using a cumyl-derived

ACS Paragon Plus Environment

Journal of the American Chemical Society 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 5

Table 1. Catalyst Screen for the Polymerization of 1-Butene. a

entry complex

yield (mg)

TOFb (h−1)

Mn(theo) (kDa)

Mnc (kDa)

Đc (M w /M n )

44

59.8

70.0

enchainment typed (mole fraction)

[mm]d

αe

Tm f (°C)

4, 2

4, 1

1, 2

1.26

0.88

0.06

0.02h

0.43

0.74

–g

1

1

598

2

2

122

9

12.2

12.8

1.60

0.94

0.04

0.02

0.75

0.91

77

3

3

233

17

23.3

23.5

1.48

0.93

0.03

0.04

0.79

0.92

80

4

4

220

16

22.0

21.0

1.51

0.91

0.04

0.05

0.73

0.90

73

0.80

0.93

85

h

5

5

234

17

23.4

14.7

1.57

0.93

0.03

0.03

6

6

200

15

20.0

18.5

1.60

0.91

0.03

0.06

0.80

0.93

86

7

7

298

22

29.8

20.5

1.69

0.93

0.03

0.04

0.84

0.94

85

a

Polymerization conditions: 3.0 ± 0.2 g 1-butene (54 mmol), 10 µmol Ni complex in 2 mL CH2Cl2, 1.0 mmol MAO; trxn = 24 h, Trxn = −40 °C. bTurnover frequency, TOF = (mol 1-butene consumed)(mol complex • time)−1. cDetermined by GPC versus polyethylene standards. dDetermined using 13 C NMR spectroscopy; see Supporting Information. eProbability factor of three consecutive 4,2-units having the same relative stereochemistry, determined using the equation [mm] = α3+(1−α)3. fDetermined by DSC. gTm not detectable. hAdditional signals are present that correspond to long chain branching. 1: R 2: R 3: R 4: R 5: R 6: R 7: R

Ar R' R

N

N Ni Br 2 R' Ar

R

= Me, R’ = H, Ar = Mesityl = Me, R' = Me, Ar = Phenyl = F, R' = Me, Ar = Phenyl = OMe, R' = Me, Ar = Phenyl = tBu, R' = Me, Ar = Phenyl = Me, R' = Me, Ar = 1-Naphthyl = tBu, R' = Me, Ar = 1-Naphthyl

Figure 1. α-Diimine nickel complexes used in this study. complex (Figure 1, complex 2) may be more effective for the regioand stereoselective polymerization of 1-butene, with the additional steric bulk potentially preserving the stereochemical information imparted by the initial insertion event Activation of complex 2 with MAO in the presence of 1-butene resulted in the production of a white, powdery material. GPC and DSC analysis revealed a polymer with modest molecular weight (12.8 kDa), controlled unimodal dispersity (1.60), and a melting transition temperature (Tm) of 77 °C (Table 1, entry 2). Analysis by 13C NMR spectroscopy revealed a microstructure composed primarily of 4,2-poly(1butene) units arising from 4,2-enchainment (Scheme 2) and high amounts of isotacticity ([mm] = 0.75). The observed tacticity is considerably higher than that reported previously for the polymerization of trans-2-butene.9

Scheme 2. Enchainment Pathways for the Polymerization of 1-Butene Using α-Diimine Nickel Catalysts. P

1,2-ins. +

LNi P

P

+

LNi

+

3

4

LNi

n

1,2-poly(1-butene)

full chain walk

1-butene

2

1

LNi

+

P

4

2

3

+

LNi P 2,1-ins. +

NiL

P

1 n

full chain walk

4,2-poly(1-butene)

LNi

+

P

4

3

2

1

n

4,1-poly(1-butene)

The addition of ortho-cumyl groups had a dramatic effect on the isoselectivity of the resulting 4,2-poly(1-butene). We believe that the steric environment created from this substitution pattern facilitates the necessary enantioface coordination event leading to stereoretentive chain isomerization. A π-stacking interaction with the ortho-aryl substituents and the acenaphthene backbone is consistently observed in

Figure 2. X-ray crystal structure of complex 5. Hydrogen atoms are omitted for clarity. Atoms drawn at the 50% probability level. the solid state for cumyl-derived complexes (Figure 2).11 We believe this π-stacking interaction persists in solution at low reaction temperature, causing the catalyst to maintain a very specific steric environment which contributes to the observed selectivity. We also cannot rule out the possibility that the catalyst structure may prevent racemization of the stereocenter after planarization during chain walking. To probe our assumptions regarding the source of selectivity for 1butene polymerization, perturbations to the ligand framework were made. Electronic variants of the ortho-cumyl nickel complexes were first studied. It was found that MAO-activated 3 with electron withdrawing para-fluoro substituents produced materials with higher molecular weight (23.3 kDa), slightly elevated Tm (80 °C), and improved isotacticity ([mm]= 0.79) compared to 2/MAO. MAO-activated 4, with electron donating para-methoxy substituents, however, produced materials with a decreased Tm (73 °C) and lower isotacticity ([mm]= 0.73). Selectivity for 4,2-enchainment also decreased slightly for 4/MAO (0.91). We also prepared and studied complex 5 with paratert-butyl substituents that are slightly more electron donating than the methyl substituents of complex 2. To our surprise, 5/MAO produced iso-4,2-poly(1-butene) with high levels of isotacticity ([mm] = 0.80) and an improved Tm of 85 °C (Figure 3). One possibility is that the tert-butyl substituents in 5 affect the ion pairing interaction of the cationic nickel complex and the MAO activator such that an improvement in selectivity is observed.12 We continued probing the system by modifying the aryl substituent of the cumyl group from phenyl (2) to 1-naphthyl (6). Complex 6/MAO produced polymer with similar tacticity to that of complex 5/MAO ([mm] = 0.80) and a similar Tm of 86 °C. We suspect the πstacking interaction between the 1-naphthyl group and the acenaphthene backbone is enhanced due to the additional π overlap, resulting in a more rigidified structure that contributes to the desired selectivity.

ACS Paragon Plus Environment

Page 3 of 5 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

Journal of the American Chemical Society After observing the enhanced selectivities for materials produced by MAO-activated 5 and 6, we attempted to increase the steric bulk of the system further by preparing complex 7, bearing both tert-butyl and 1naphthyl substituents (Figure 1). The polymer produced by 7/MAO exhibited the highest isotacticity out of all samples previously generated ([mm] = 0.84). Interestingly, despite the increase in tacticity, there was no improvement in the observed melting temperature of the resulting material (Table 1, entry 7). C C

C

mm

B

A

D

E

F

L M

m

r

m

r

r

A

F 4,1-unit

K

m

4,2-unit

B mr

J

rr

I

1,2-unit H

K

JD

L F

M

E, I

H

Figure 3. 13C NMR spectrum of iso-4,2-poly(1-butene) produced by complex 5/MAO (Table 1, entry 5). The thermal properties of iso-4,2-poly(1-butene) are influenced not only by tacticity, but also by branch composition. A small percent of ethyl branches were detected in these systems which can be installed through 1,2-insertion of 1-butene without chain walking. Ethyl branches can considerably reduce Tm and crystallinity compared to methyl branches.13 Another defect observed in this system arises from 4,1enchainment, where 2,1-insertion followed by complete chain walking produces a linear “polyethylene” segment. Due to the small amount of these units (< 5%), it is difficult to quantify its effect on the properties of the resulting iso-4,2-poly(1-butene).14 Recent studies have shown that iPP with < 5% 3,1-insertions, made with related cumyl-substituted catalysts retained most of its crystallinity, suggesting that the effect in this system may also be small.15

Table 2. Temperature and Concentration Effects on the Polymerization of 1-Butene Using 5/MAO. a entry

Trxn (°C)

conc (M)

yield (mg)

Mnb Đb Tm c [mm]d (kDa) (Mw/Mn) (°C) e

n.d.

f

1

−10

7.8

262

10.8

1.95



2

−30

7.8

420

23.7

1.80

66

0.53

3

−40

7.8

234

14.7

1.57

85

0.80

4

−50

7.8

173

10.6

1.81

85

0.80

5

−40

6.0

g

163

12.4

1.46

80

0.70

6

−40

9.1h

464

16.3

2.17

73

0.59

a

Polymerization conditions: 3.0 ± 0.2 g 1-butene (54 mmol), 10 µmol of complex 5 in 2 mL CH2Cl2, 100 equiv. MAO; trxn = 24 h. bDetermined by GPC versus PE standards. cDetermined by DSC. dDetermined using 13C NMR spectroscopy. eNo detectable Tm. fNot determined. g1.5 ± 0.2 g 1-butene (27 mmol). h6.0 ± 0.2 g 1-butene (107 mmol). We further studied the effects of reaction conditions on the thermal properties of the resulting polymers produced by 5/MAO (Table 2). The reaction temperature had a dramatic effect on polymer thermal properties. By increasing the temperature from −40 °C to −30 °C, the melting temperature of the resulting material dropped considerably from 85 °C to 66 °C corresponding to a significant loss in tacticity (Table 2, entry 2). Increasing the reaction temperature further to −10 °C resulted in a complete loss of crystallinity. Cooling the reaction

further to −50 °C decreased the polymerization rate and polymer yield, and resulted in materials with similar thermal properties to those produced at −40 °C (Table 2, entry 3). The effect of 1-butene concentration was also studied. Increasing the concentration from the standard 7.8 M to 9.1 M (Table 2, entry 6). resulted in higher polymer yield but a marked decrease in tacticity ([mm] = 0.59) and melting temperature (Tm = 73 °C). A decrease in 4,2-enchainment in favor of 1,2-enchainment was also observed at a higher concentration of 1-butene.16 Decreasing the concentration of 1butene to 6.0 M resulted in lower polymer yield, but also a slight decrease in tacticity ([mm] = 0.70) and melting temperature (Tm = 80 °C: Table 2, entry 5). Lower 1-butene concentrations may increase the rate of chain isomerization relative to insertion, allowing introduction of stereoerrors. Returning to our assertion of stereoselective insertion and stereoretentive chain walking, it is not immediately clear why complex 1/MAO produces iPP from propylene but atactic polymer from 1-butene. Propylene polymerization can proceed through two different mechanisms that lead to the same iPP product: (1) consecutive isoselective insertions without chain walking, or (2) consecutive isoselective insertions with full stereoretentive chain walking. To confirm the operative mechanism for propylene, we performed a study using [3-d3]propylene at −78 °C and 1/MAO. If chain walking occurs, the deuterium atoms should scramble throughout the polymer backbone. Under these conditions, however, the deuterium atoms remained on the methyl branches of the polymer, demonstrating that chain walking during propylene polymerization does not occur (Scheme 3).16 This suggests that for the polymerization of 1-butene, complex 1/MAO either does not perform the isoselective insertion of 1-butene, or tacticity is lost through racemization of the installed stereocenter during chain isomerization. Since the formation of iso-4,2-poly(1-butene) requires chain isomerization, the isotactic enchainment of 1-butene in this work represents a rare example of stereoretentive chain walking.

Scheme 3. Deuterated Propylene Study Using 1/MAO. no chain walk +

LNi-P

CD 3 1,2-ins.

cat. 1

P

+

LNi

P

+

LNi

observed

D 3C D

D 3C full chain walk

+

LNi

P D D

not observed

In summary, we performed the iso- and regioselective polymerization of 1-butene using cationic α-diimine nickel(II) complexes to produce a new polymer, isotactic-4,2-poly(1-butene). The methyl substituents on the cumyl-groups of the ligand are crucial for the isoselective insertion and stereocontrolled chain walking. Furthermore, rigidifying the catalyst through π-stacking interactions with the ligand backbone was beneficial for improving tacticity. This system allows access to isotactic polymers from a simple, inexpensive α-olefin. Additional catalyst development is currently in progress to further increase the selectivity of this unique polymerization.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Experimental procedures and spectral data (PDF) and crystallographic data (CIF)

AUTHOR INFORMATION Corresponding Author E-mail: [email protected]

Author Contributions

ACS Paragon Plus Environment

Journal of the American Chemical Society ‡

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

Authors contributed equally to this work.

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENTS Funding was provided by the Center for Sustainable Polymers, a National Science Foundation (NSF)-supported Center for Chemical Innovation (CHE-1413862). This research made use of the NMR facility at Cornell University and is supported, in part, by the NSF under the award number CHE-1531632.

REFERENCES (1) (a) Takai, T.; Mochizuki, D.; Umeno, M. Method of Producing Propylene Containing Biomass-Origin Carbon. Patent WO2007/55361, 2008. (b) Dürre, P. Biobutanol: An Attractive Biofuel. Biotechnol. J. 2007, 2, 1525–1534. (c) Kohse-Höinghaus, K.; Oßwald, P.; Cool, T. A.; Kasper, T.; Hansen, N.; Qi, F.; Westbrook, C. K.; Westmoreland, P. R. Biofuel Combustion Chemistry: From Ethanol to Biodiesel. Angew. Chem., Int. Ed. 2010, 49, 3572–3597. (d) Komon, Z. J. A.; Bu, X.; Bazan, G. C. Synthesis of Butene-Ethylene and HexeneButene-Ethylene Copolymers from Ethylene via Tandem Action of WellDefined Homogeneous Catalysts. J. Am. Chem. Soc. 2000, 122, 1830–1831. (2) Freeman, A.; Mantell, S. C.; Davidson, J. H. Mechanical Performance of Polysulfone, Polybutylene, and Polyamide 6/6 in Hot Chlorinated Water. Solar Energy 2005, 79, 624–637. (3) (a) De Rosa, C.; Auriemma, F.; Resconi, L. Metalloorganic Polymerization Catalysis as a Tool to Probe Crystallization Properties of Polymers: The Case of Isotactic Poly(1-Butene). Angew. Chem., Int. Ed. 2009, 48, 9871– 9874. (b) De Rosa, C.; Auriemma, F.; Ruiz de Ballesteros, O.; Esposito, F.; Laguzza, D.; Di Girolamo, R.; Resconi, L. Crystallization Properties and Polymorphic Behavior of Isotactic Poly(1-Butene) from Metallocene Catalysts: The Crystallization of Form I from the Melt. Macromolecules 2009, 42, 8286– 8297. (4) (a) Johnson, L. K.; Killian, C. M.; Brookhart, M. New Pd(II)- and Ni(II)-Based Catalysts for Polymerization of Ethylene and α-Olefins. J. Am. Chem. Soc. 1995, 117, 6414–6415. (b) Killian, C. M.; Tempel, D. J.; Johnson, L. K.; Brookhart, M. Living Polymerization of α-Olefins Using NiII-α- Diimine Catalysts. Synthesis of New Block Polymers Based on α-Olefins. J. Am. Chem. Soc. 1996, 118, 11664–11665. (c) Guan, Z.; Cotts, P. M.; McCord, E. F.; McClain, S. J. Chain Walking: A New Strategy to Control Polymer Topology. Science 1999, 283, 2059–2062. (d) Gates, D. P.; Svejda, S. A.; Oñate, E.; Killian, C. M.; Johnson, L. K.; White, P. S.; Brookhart, M. Synthesis of Branched Polyethylene Using (α-Diimine)nickel(II) Catalysts: Influence of Temperature, Ethylene Pressure, and Ligand Structure on Polymer Properties. Macromolecules 2000, 33, 2320–2334. (e) Ittel, S. D.; Johnson, L. K.; Brookhart, M. Late-Metal Catalysts for Ethylene Homo- and Copolymerization. Chem. Rev. 2000, 100, 1169–1203. (f) Tempel, D. J.; Johnson, L. K.; Huff, R. L.; White, P. S.; Brookhart, M. Mechanistic Studies of Pd(II)-α-Diimine-Catalyzed Olefin Polymerizations. J. Am. Chem. Soc. 2000, 122, 6686–6700. (g) Leatherman, M. D.; Svejda, S. A.; Johnson, L. K.; Brookhart, M. Mechanistic Studies of Nickel(II) Alkyl Agostic Cations and Alkyl Ethylene Complexes: Investigations of Chain Propagation and Isomerization in (α-diimine)Ni(II)-Catalyzed Ethylene Polymerization. J. Am. Chem. Soc. 2003, 125, 3068–3081. (h) Subramanyam, U.; Rajamohanan, P. R.; Sivaram, S. A Study of the Structure of Poly(Hexene-1) Prepared by Nickel(α-diimine)/MAO Catalyst Using High Resolution NMR Spectroscopy. Polymer 2004, 45, 4063–4076. (i) Bomfim, J. A. S.; Dias, M. L.; Filgueiras, C. A. L.; Peruch, F.; Deffieux, A. The Effect of Polymerization Temperature on the Structure and Properties of Poly(1-Hexene) and Poly(1Decene) Prepared with a Ni(II)-diimine Catalyst. Catal. Today 2008, 133– 135, 879–885. (j) Camacho, D. H.; Guan, Z. Designing Late-Transition Metal Catalysts for Olefin Insertion Polymerization and Copolymerization. Chem. Commun. 2010, 46, 7879–7893. (k) Sun, G.; Guan, Z. Cascade Chain-Walking Polymerization to Generate Large Dendritic Nanoparticles. Macromolecules 2010, 43, 4829–4832. (l) Liu, J.; Chen, D.; Wu, H.; Xiao, Z.; Gao, H.; Zhu, F.; Wu, Q. Polymerization of α-Olefins Using a Camphyl α-Diimine Nickel Catalyst at Elevated Temperature. Macromolecules 2014, 47, 3325–3331. (m) Leone, G.; Mauri, M.; Bertini, F.; Canetti, M.; Piovani, D.; Ricci, G. Ni(II) αDiimine-Catalyzed α-Olefins Polymerization: Thermoplastic Elastomers of Block Copolymers. Macromolecules 2015, 48, 1304–1312. (n) Guo, L.; Dai,

S.; Sui, X.; Chen, C. Palladium and Nickel Catalyzed Chain Walking Olefin Polymerization and Copolymerization. ACS Catal. 2016, 6, 428–441. (o) Dai, S.; Sui, X.; Chen, C. Synthesis of High Molecular Weight Polyethylene Using Iminopyridyl Nickel Catalysts. Chem. Commun. 2016, 52, 9113–9116. (5) McLain, S. J.; Feldman, J.; McCord, E. F.; Gardner, K. H.; Teasley, M. F.; Coughlin, E. B.; Sweetman, K. J.; Johnson, L. K.; Brookhart, M. Addition Polymerization of Cyclopentene with Nickel and Palladium Catalysts. Macromolecules 1998, 31, 6705–6707. (6) (a) Okada, T.; Park, S.; Takeuchi, D.; Osakada, K. Pd-Catalyzed Polymerization of Dienes that Involves Chain-Walking Isomerization of the Growing Polymer End: Synthesis of Polymers Composed of Polymethylene and Five-Membered-Ring Units. Angew. Chem., Int. Ed. 2007, 46, 6141–6143. (b) Okada, T.; Takeuchi, D.; Shishido, A.; Ikeda, T.; Osakada, K. Isomerization Polymerization of 4-Alkylcyclopentenes Catalyzed by Pd Complexes: Hydrocarbon Polymers with Isotactic-Type Stereochemistry and Liquid-Crystalline Properties. J. Am. Chem. Soc. 2009, 131, 10852–10853. (c) Okada, T.; Takeuchi, D.; Osakada, K. Cyclopolymerization of Monoterminal 1,6-Dienes Catalyzed by Pd Complexes. Macromolecules 2010, 43, 7998–8006. (d) Park, S.; Okada, T.; Takeuchi, D.; Osakada, K. Cyclopolymerization and Copolymerization of Functionalized 1,6-Heptadienes Catalyzed by Pd Complexes: Mechanism and Application to Physical-Gel Formation. Chem. Eur. J. 2010, 16, 8662–8678. (e) Takeuchi, D.; Osakada, K. Controlled Isomerization Polymerization of Olefins, Cycloolefins, and Dienes. Polymer 2016, 82, 392–405. (7) Mei, T.-S.; Werner, E. W.; Burckle, A. J.; Sigman, M. S. Enantioselective Redox-Relay Oxidative Heck Arylations of Acyclic Alkenyl Alcohols using Boronic Acids. J. Am. Chem. Soc. 2013, 135, 6830–6833. (b) Mei. T.-S.; Patel, H. H.; Sigman, M. S. Enantioselective Construction of Remote Quaternary Stereocentres. Nature 2014, 508, 340–344. (8) (a) Li, H.; Rojas, G.; Wagener, K. B. Precision Long-Chain Branched Polyethylene via Acyclic Diene Metathesis Polymerization. ACS Macro Lett. 2015, 4, 1225–1228. (b) Rojas, G.; Berda, E. B.; Wagener, K. B. Precision Polyolefin Structure: Modeling Polyethylene Containing Alkyl Branches. Polymer 2008, 49, 2985–2995. (c) Sworen, J. C.; Smith, J. A.; Berg, J. M.; Wagener, K. B. Modeling Branched Polyethylene: Copolymers Possessing Precisely Placed Ethyl Branches. J. Am. Chem. Soc. 2004, 126, 11238–11246. (9) Cherian, A. E.; Lobkovsky, E. B.; Coates, G. W. Chiral Anilines: Development of C2-Symmetric, Late-Transition Metal Catalysts for Isoselective 2Butene Polymerization. Chem. Commun. 2003, 2566–2567. (10) We previously reported the polymerization of 1-butene using complex 1 activated by PMAO-IP and observed amorphous 4,2-poly(1-butene) with low tacticity: Rose, J. M.; Cherian, A. E.; Coates, G. W. Living Polymerization of αOlefins with an α-Diimine Ni(II) Catalyst: Formation of Well-Defined Ethylene-Propylene Copolymers through Controlled Chain-Walking. J. Am. Chem. Soc. 2006, 128, 4186–4187. (11) (a) Cherian, A. E.; Rose, J. M.; Lobkovsky, E. B.; Coates, G. W. A C2Symmetric, Living α-Diimine Ni(II) Catalyst: Regioblock Copolymers from Propylene. J. Am. Chem. Soc. 2005, 127, 13770–13771. (b) Rose, J. M.; Deplace, F.; Lynd, N. A.; Wang, Z.; Hotta, A.; Lobkovsky, E. B.; Kramer, E. J.; Coates, G. W. C2-Symmetric Ni(II) α-Diimines Featuring Cumyl-Derived Ligands: Synthesis of Improved Elastomeric Regioblock Polypropylenes. Macromolecules 2008, 41, 9548–9555. (12) Detailed studies on ion paring in late transition metal complexes are limited, but ion pairing in early transition metal complexes are well documented. For a review on this topic, please refer to: Macchioni, A. Ion Pairing in Transition-Metal Organometallic Chemistry. Chem. Rev. 2005, 105, 2039–2074. (13) (a) Starck, P.; Rajanen, K.; Löfgren, B. Comparative Studies of Ethylene-α-Olefin Copolymers by Thermal Fractionations and TemperatureDependent Crystallinity Measurements. Thermochim. Acta 2003, 395, 169– 181. (b) Isasi, J. R.; Haigh, J. A.; Graham, J. T.; Mandelkern, L.; Alamo, R. G. Some Aspects of the Crystallization of Ethylene Copolymers. Polymer 2000, 41, 8813–8823. (14) Analogously, it has been shown that a small percentage (