You are on page 1of 14

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/306090629

Elucidating the Key Role of Phosphine


−Sulfonate Ligands in Palladium-Catalyzed
Ethylene Polymerization: Effect of...

Article in ACS Catalysis · August 2016


DOI: 10.1021/acscatal.6b00911

CITATIONS READS

3 27

8 authors, including:

Lung Wa Chung Shingo Ito


Southern University of Science and Technology The University of Tokyo
50 PUBLICATIONS 1,300 CITATIONS 50 PUBLICATIONS 1,426 CITATIONS

SEE PROFILE SEE PROFILE

Keiji Morokuma
Fukui Institute for Fundamental Chemistry
469 PUBLICATIONS 15,842 CITATIONS

SEE PROFILE

All content following this page was uploaded by Lung Wa Chung on 23 August 2016.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Research Article

pubs.acs.org/acscatalysis

Elucidating the Key Role of Phosphine−Sulfonate Ligands in


Palladium-Catalyzed Ethylene Polymerization: Effect of Ligand
Structure on the Molecular Weight and Linearity of Polyethylene
Ryo Nakano,† Lung Wa Chung,‡ Yumiko Watanabe,§ Yoshishige Okuno,§ Yoshikuni Okumura,∥
Shingo Ito,† Keiji Morokuma,*,⊥ and Kyoko Nozaki*,†

Department of Chemistry and Biotechnology, Graduate School of Engineering, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku,
Tokyo 113-8656, Japan

Department of Chemistry, South University of Science and Technology of China, Shenzhen 518055, People’s Republic of China
§
Computational Science and Technology Information Center, Showa Denko K.K., 1-1-1 Ohnodai, Midori-ku, Chiba, Chiba 267-0056,
Japan

Institute for Advanced and Core Technology, Showa Denko K.K., 2 Nakanosu, Oita, Oita 870-1809, Japan

Fukui Institute for Fundamental Chemistry, Kyoto University, 34-4 Takano Nishihiraki-cho, Sakyo, Kyoto 606-8103, Japan
*
S Supporting Information

ABSTRACT: The mechanism of linear polyethylene formation catalyzed by palladium/phosphine−sulfonate and the effect of
the ligand structure on the catalytic performance, such as linearity and molecular weight of the polyethylene, were reinvestigated
theoretically and experimentally. We used dispersion-corrected density functional theory (DFT-D3) to study the entire
mechanism of polyethylene formation from (R2PC6H4SO3)PdMe(2,6-lutidine) (R = Me, t-Bu) and elucidated the key steps that
determine the molecular weight and linearity of the polyethylene. The alkylpalladium ethylene complex is the key intermediate
for both linear propagation and β-hydride elimination from the growing polymer chain. On the basis of the key species, the
effects of substituents on the phosphorus atom (R = t-Bu, i-Pr, Cy, Men, Ph, 2-MeOC6H4, biAr) were further investigated
theoretically to explain the experimental results in a comprehensive manner. Thus, the experimental trend of molecular weights
of polyethylene could be correlated to the ΔΔG⧧ value between (i) the transition state of linear propagation and (ii) the
transition state of the path for ethylene dissociation leading to β-hydride elimination. Moreover, the experimental behavior of the
catalysts under varied ethylene pressure was well explained by our computation on the small set of key species elucidated from
the entire mechanism. In our additional experimental investigations, [o-Ani2PC6H4SO3]PdH[P(t-Bu)3] catalyzed a hydrogen/
deuterium exchange reaction between ethylene and MeOD. The deuterium incorporation from MeOD into the main chain of
polyethylene, therefore, can be explained by the incorporation of deuterated ethylene formed by a small amount of Pd−H
species. These insights into the palladium/phosphine−sulfonate system provide a comprehensive understanding of how the
phosphine−sulfonate ligands function to produce linear polyethylene.
KEYWORDS: phosphine sulfonate, palladium catalysts, ethylene polymerization, linearity, molecular weight,
suppression of β-hydride elimination

1. INTRODUCTION compatibility, especially with polar materials.3 Although early-


Polyolefins such as polyethylene and polypropylene comprise transition-metal catalysts used in the industrial production of
one of the largest classes of chemicals today.1,2 A longstanding polyolefins are heavily deactivated by polar functional groups,
challenge in polyolefin synthesis is the development of efficient
methods to introduce polar functional groups into the polymer Received: March 30, 2016
structure. These polar groups greatly improve material Revised: July 5, 2016
properties such as adhesion, dye retention, printability, and

© XXXX American Chemical Society 6101 DOI: 10.1021/acscatal.6b00911


ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

recently emerged late-transition-metal complexes have been The observation of 1-octene isomerization into internal octenes
revealed to act as catalysts for the copolymerization of polar by Claverie and co-workers also supported the chain-walking
monomers directly with ethylene.4,5 Using α-diimine ligated reactions by palladium/phosphine−sulfonate catalysts.23 From
palladium complexes, coordination−insertion copolymerization a theoretical viewpoint, in contrast, Ziegler and co-workers
of ethylene with methyl acrylate to produce highly branched reported that the β-hydride elimination is slower in comparison
copolymers was first accomplished by Brookhart and co- to the polymer-chain growth in the catalyst system.24 In order
workers in 1996.6 In 2002, Drent et al. reported the synthesis of to consolidate this discrepancy between the experimental and
linear copolymers of ethylene and alkyl acrylates using computational insights, we previously studied the mechanism of
palladium/phosphine−sulfonate catalysts.7 Subsequently, the linear polyethylene formation. 25 On the basis of the
palladium/phosphine−sulfonate system has been proven to experimental observation of long-alkyl intermediates during
tolerate a variety of functional groups,8 which successfully polymerization and the detailed theoretical calculations on the
enabled the unique copolymerizations of ethylene with polar mechanism of both linear and branched propagation pathways,
monomers such as acrylates,7,9 acrylic acid,10 acrylonitrile,11 we proposed that the chain propagation dominates over β-
vinyl halides,12 vinyl ethers,13 vinyl acetate,14 and allyl hydride elimination in the presence of sufficient ethylene
monomers.15 pressure.
The differences in the polymer microstructures of poly- Since our mechanistic studies in 2009, further mechanistic
ethylenes prepared by palladium/α-diimine and palladium/ understanding of the palladium/phosphine−sulfonate system
phosphine−sulfonate catalysts are noteworthy (Figure 1). has been accumulated. The mechanism of cis/trans isomer-
ization, which is the key step prior to the ethylene insertion
reaction, was examined by Jordan and co-workers.26 After the
isolation of cis and trans isomers of alkylpalladium complexes
and kinetic control experiments, they concluded that the
isomerization process does not require neutral ligands such as
pyridine and 2,6-lutidine, as suggested in our previous
computational study. The influence of the fourth monodentate
ligand was scrutinized by Mecking and co-workers.27 Their
study revealed little effect of the fourth ligands on the polymer
microstructure and the ceiling of activity by modification of
such ligands. Moreover, they established an effective access to a
fourth-ligand free active species by in situ activation of the
sodium chloride adduct Na[(R2PC6H4SO3)PdMeCl] with
AgBF4.
Figure 1. Differences in polyethylene microstructures obtained by
The steric effect from the substituents on the phosphorus
palladium/α-diimine and palladium/phosphine−sulfonate catalysts. atom plays an important role in modulating the molecular
weight of the obtained polyethylene (Figure 2), although the

Palladium/α-diimine catalysts are susceptible to repetitive β-


hydride elimination and reinsertion reactions, commonly
referred to as chain-walking,16 and generally form highly
branched polyolefins. On the other hand, phosphine−sulfonate
ligated palladium complexes produce highly linear polyethylene
in which the linear versus branched ratio in a main chain is
usually more than 1000. In addition to palladium and nickel17
catalysts bearing phosphine−sulfonate ligands, group 10 metal
catalysts that can achieve a comparable level of linear selectivity
are still limited. Nickel/salicylaldiminate18 and SHOP-type
nickel/R2PC6H4O complexes19 with elaborate catalyst tuning
produced highly linear polyethylene. Recently, we also
developed palladium complexes bearing bis(phosphine mon-
oxide) ligands (BPMO)20 or imidazo[1,5-a]quinolin-9-olate-1-
ylidene (IzQO)21 ligands, which produce almost linear
polyethylene with minimum branching.
The unique linear selectivity in polyethylene formation by Figure 2. Substituents on the phosphorus atom effective in increasing
the palladium/phosphine−sulfonate system prompted research- the molecular weight of polyethylene.
ers to study the reaction mechanism of ethylene polymerization
both experimentally22,23 and theoretically.24,25 From the exact role of the substituents remains unclear. Systematic
experimental side, Jordan and co-workers reported their variations of the substituent on the phosphorus atom were
attempt to copolymerize 6-chloro-1-hexene with ethylene reported by Claverie,28 Jordan,29 Mecking,27c,30 Nozaki,31
using palladium/phosphine−sulfonate catalysts. However, the Rieger,32 and their co-workers, independently. Generally, the
microstructure analysis of the obtained polymers suggested the molecular weight of polyethylene was increased by the steric
occurrence of β-chloride elimination from the ω-chloroalkyl- protection of the axial orientation of the palladium center. The
palladium species, which implied that chain-walking also highest molecular weight of polyethylene of more than 50000
proceeded with the palladium/phosphine−sulfonate system.22 was accomplished with catalysts bearing Men (menthyl)31a or
6102 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

biaryl (2-[2′,6′-(MeO)2C6H3]C6H4) substituents on the depth computational study on the entire mechanism of
phosphorus atom (Figure 2). This trend has been interpreted ethylene polymerization using dispersion-corrected density
on the basis of the inhibition of the associative chain-transfer functional theory (DFT-D3) and a realistic catalyst model
process by the steric protection, which enhances the molecular was performed. The study allowed us to determine the key
weight of polyethylene, in parallel to the aforementioned steps/factors affecting the molecular weight and microstructure
palladium/α-diimine complexes in which the blocking of the of polyethylene. In the second part, the key steps in the first
axial positions effectively hampers the chain transfer, leading to part were further examined both experimentally and theoret-
the production of polyethylene of high molecular weights ically with various substituents on the phosphorus atom. The
(Scheme 1).33,34 Our previous computational studies,25 which effect of the substituents was reasonably explained by the
relative energy levels of the key species. In the third part, a new
Scheme 1. Protection of the Axial Orientation of a Palladium plausible mechanism of D incorporation into the polyethylene
Center Bearing an α-Diimine Ligand, Leading to the chain from MeOD was proposed, on the basis of control
Suppression of Chain Transfer and the Production of High- experiments and mechanistic considerations.
Molecular-Weight Polyethylene
2. RESULTS AND DISCUSSION
2.1. Entire Free Energy Surface with Improved
Computational Methods and Models. In this computa-
tional study, several points were improved from our previous
computations.25
(1) Instead of the hybrid functional B3LYP, the dispersion-
corrected functional B3LYP-D3 was employed for better
focused on highly simplified model complexes bearing Me- or
estimation of the coordination energy.37 The solvent effect
Ph-substituted phosphine−sulfonate ligands, were unsuccessful
(toluene) was also included for all of the species by using the
in providing a mechanistic rationale for the steric influence.
SMD method.38
Recently, Rezabal and co-workers studied the steric effect of
(2) Several new transition states and intermediates involved
Me-, Ph-, and ferrocenyl (Fc)-substituted phosphine−sulfonate
ligands.35 They explained the steric effect on the chain length of in the β-hydride elimination step were identified.
polyethylene obtained by probing a pyridine dissociation step (3) In our previous study,25 the palladium/phosphine−sul-
prior to β-hydride elimination. fonate complexes were simplified by replacement of the
In addition, deuterium-labeling experiments suggested the substituents on the phosphorus atom by Me groups and 2,6-
involvement of reversible β-hydride elimination during ethylene lutidine by pyridine, as a model for comprehensive calculation.
polymerization,36 which was not in accordance with the results In addition to the (Me2PC6H4SO3)Pd(n-propyl)(2,6-lutidine)
of our previous computation.25 Nozaki and co-workers complex, we also carried out calculations with a realistic model
observed deuterium incorporation into the main chain of of the complex (t-Bu2PC6H4SO3)Pd(n-propyl)(2,6-lutidine).
polyethylene from MeOD as the deuterium source, when The results of the theoretical calculation for polyethylene
(Cy2PC6H4SO3)PdMe(2,6-lutidine) or [(o-Ani)2PC6H4SO3]- formation by (R2PC6H4SO3)Pd(n-propyl)(2,6-lutidine) com-
PdMe(2,6-lutidine) was heated in a mixture of MeOD and plexes (R = Me, t-Bu) are summarized in Figures 3 and 4,
toluene under ethylene pressure (Scheme 2).36 The proposed respectively. In the following discussion, the intermediates are
numbered according to our previous report25 for ease of
Scheme 2. Deuterium Incorporation into the Main Chain of comparison. Due to the electronic asymmetry of the
Polyethylene from MeOD phosphine−sulfonate ligands which consist of a strong donor
motif (phosphine) and a weak donor motif (sulfonate), all the
intermediates and the transition states have a pair of cis and
trans isomers such as 10 and 10′. Substituents on the
phosphorus atom are shown as subscripts such as 10Me and
10t‑Bu, when discrimination is required. The ethylene-
coordinated intermediate 10 was set as the starting complex
according to our previous paper.
Grimme’s dispersion correction (D3) relatively stabilized the
mechanism for the deuterium incorporation was as follows: the intermediates and transition states bearing neutral ligands, such
palladium hydride species formed after β-hydride elimination as 2,6-lutidine and ethylene (see Scheme S1 and Table S1 in
from the Pd alkyl species can undergo hydride/deuteride the Supporting Information for a comparison). On the basis of
exchange at the palladium center and return to the polymer- Claverie’s experimental observation of high activity for ethylene
ization cycle. This proposed mechanism including a reversible homopolymerization,28d t-Bu substituents highly destabilized
β-hydride elimination reaction during ethylene polymerization resting intermediates 8t‑Bu and 10t‑Bu by a combination of high
was not compatible with our previous conclusion based on steric repulsion and electron-donating ability.
mechanistic studies claiming that the chain propagation In the following sections, the mechanisms of linear
dominates over β-hydride elimination in the presence of propagation, β-hydride elimination, and subsequent branch
ethylene pressure. formation and chain transfer are discussed.
In the present study, we report a comprehensive mechanistic 2.1.1. Linear Propagation. Linear polyethylene formation
rationale for polyethylene formation by the palladium/ preferentially proceeds along ethylene-coordinated intermedi-
phosphine−sulfonate system, which can explain the aforemen- ate 10, pentacoordinated cis/trans isomerization via TS(10-
tioned newly reported observations. In the first part, an in- 10′), and then ethylene insertion via TS(10′-11′) (blue
6103 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

Figure 3. Free energy profile of palladium/dimethylphosphine sulfonate catalyst (B3LYP-D3/6-31G*, lanl2dz for palladium) in toluene.

Figure 4. Free energy profile of palladium/di-tert-butylphosphine sulfonate catalyst (B3LYP-D3/6-31G*, lanl2dz for palladium) in toluene.

pathway in Figures 3 and 4), as discussed in our previous 2,6-lutidine is comparable with that of ethylene, considering
work.25 that the concentration of ethylene is much higher than that of
2.1.2. β-Hydride Elimination. 2.1.2.1. Formation of cis-P− 2,6-lutidine under sufficient ethylene pressure. The mole
Pd−H Species 12a′ and 12′. In the dispersion-corrected free fraction of ethylene in toluene under 3.0 MPa of ethylene
energy surfaces, the most favorable pathway for β-hydride pressure at 90 °C was reported to be 0.20.39 Then, under
elimination proceeds along the formation of β-agostic typical conditions (0.10 mmol/L of catalyst in toluene and 3.0
intermediate 9′ and TS(9′-12a′) (Figures 3 and 4). For this MPa of ethylene pressure at 80 °C), the ratio between ethylene
route, the rates of all the paths for the formation of and 2,6-lutidine (=catalyst) in the solvent mixture reaches more
intermediate 9′ are much slower than that of the following β- than 2400. When the activity difference between ethylene and
hydride elimination via TS(9′-12a′). 2,6-lutidiene is ignored, the ratio corresponds to ca. 5 kcal/mol
Three possible types of formation pathways were found for difference. Thus, TS(9′-10′) is the key transition state before β-
the formation of intermediate 9′: (1) direct cis/trans hydride elimination under a high ethylene pressure.40
isomerization from β-agostic intermediate 9 to 9′ via TS(9- The contribution of pathway 3, which is discussed in this
9′), (2) the fourth ligand associated cis/trans isomerization via work for the first time, depends upon the steric pressure from
TS(8-8med)/TS(8med-8′) or TS(10-10′) and subsequent substituents on the phosphorus atom. With the less hindered
associative ligand exchange between the β-C−H bond and Me-substituted ligand, the relative free energy of intermediate
ethylene or 2,6-lutidine (via TS(8′-9′) or TS(9′-10′), 9NA′Me is higher than that of TS(9′-10′)Me by ca. 2 kcal/mol,
respectively), and (3) the fourth ligand associated cis/trans and then the contribution of the type 3 pathway should be
isomerization, ligand dissociation to three-coordinated inter- negligible under sufficient ethylene pressure. On the other
mediate 9NA′ (9′ with no agostic interaction), and following hand, with the more sterically demanding t-Bu-substituted
isomerization to 9′. The contribution of pathway 1 via TS(9- ligand, three-coordinated intermediate 9NA′t‑Bu is more stable
9′) should be negligible, since it requires much higher than TS(9′-10′)t‑Bu by ca. 5 kcal/mol, suggesting that this
activation energy than pathways 2 and 3 (Figures 3 and 4). pathway can contribute even under high pressure of ethylene.
For pathway 2, the key dissociation steps via TS(8′-9′) and 2.1.2.2. Formation of trans-P−Pd−H Species 12a and 12.
TS(9′-10′) are associative ligand exchange reactions in which Other β-hydride elimination/reinsertion pathways via TS(9-
the fourth neutral ligand (2,6-lutidine or ethylene) is replaced 12a), TS(9a-12), TS(12-13a), and TS(12a-13) were then
by a β-agostic interaction via five-coordinated transition states. discussed for the first time. The difference between 12a and 12
Therefore, the steric hindrance on the axial orientation of the lies in the orientation of the coordinating propylene (see
palladium center provided by substituents on the phosphorus Figures S28 and S30 in the Supporting Information), and 12a is
atom can retard these processes. The acceleration effect with more easily accessible. The thermodynamic instability of
6104 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

Figure 5. Possible pathways for chain-transfer reaction: (a) R = Me; (b) R = t-Bu.

intermediates 12a and 12 can be attributed to the strong trans tendency indicates that, once the β-hydride elimination does
influence from the phosphine donor. The activation barriers of take place, there is a better chance of methyl branch formation
TS(9-12a) and TS(9a-12) for forming the trans-P−Pd−H if it does not proceed to chain transfer.
species are higher than that for the main pathway of 2.1.3.2. Linear Propagation and Methyl-Branch Forma-
polyethylene formation via TS(10′-11′) (shown in blue) by tion from trans-P−Pd−H Species 12 or 12a. Due to the
4−6 kcal/mol due to the thermodynamic instability of the thermodynamic inaccessibility of 12 and 12a (trans-P−Pd−H
products and are comparable to that for TS(9′-10′) (which is species), further reactions from those species are less likely to
the rate-determining step for the formation of the cis-P−Pd−H happen. The possible pathways from 12 and 12a should be
species). Therefore, the formation of the trans-P−Pd−H discussed separately, since we could not find the direct
species should be less plausible, unless under extremely low interconversion transition states between intermediates 12
concentration of ethylene. and 12a. When the ethylene is placed coplanar with the
2.1.3. Possible Pathways after β-Hydride Elimination. palladium plane, reinsertion of the olefin easily proceeds to re-
2.1.3.1. Linear Propagation, Chain Transfer, and Methyl- form 9 and 9a.
Branch Formation from cis-P−Pd−H Species 12a′ or 12′. From the more accessible intermediate 12a, further
There are three possible pathways from intermediates 12a′ and conversion to methyl branches or chain transfer is less
12′ (cis-P−Pd−H species), which are formed after β-hydride plausible. The isomerization to branched intermediate 13 via
elimination from 9′ or 9a′ and in fast equilibrium via TS(12a′- TS(12a-13) required much greater activation energy in
12′): reinsertion to linear-chain formation (the left side of comparison to the pathway to the linear intermediate 9 via
Figures 3 and 4), methyl-branch formation (the right side of TS(9-12a) by more than 5 kcal/mol for both of the
Figures 3 and 4), and chain-transfer reactions (Figure 5). When substituents on the phosphorus atom. For an associative
the three possibilities are compared, the chain-transfer reactions chain-transfer reaction, five-coordinated intermediate 18Me was
are more favored than the other two pathways in the cases of the intermediate for the transition between 12Me and 17Me
both the Me- and t-Bu-substituted phosphine−sulfonate (Figure 5a). On the other hand, with t-Bu substituents, only the
ligands. The Me-substituted palladium/phosphine−sulfonate direct transition state TS(12a-17)t‑Bu was found for the
favors an associative chain transfer via five-coordinated TS(12′- transition between 12t‑Bu and 17t‑Bu (Figure 5a). Given that
17′)Me, while the sterically crowded t-Bu-substituted palla- the dissociation of each alkene from 18Me requires 1−2 kcal/
dium/phosphine−sulfonate favors a dissociative chain transfer mol activation energy, we treated both 18Me and TS(12a-
via the three-coordinated intermediate 19′t‑Bu (Figure 5). 17)t‑Bu as the barrier for the associative chain-transfer reaction.
Here, the selectivity of linear chain propagation versus While the associative chain-transfer reaction via 18Me seems
branch formation after β-hydride elimination is discussed. Since feasible with the Me-substituted phosphine−sulfonate ligand
the β-hydride elimination/reinsertion (via TS(9′-12a′)/TS- (Figure 5a), the chain-transfer reactions via TS(12a-17)t‑Bu or
(12′-13′)) and propylene rotation (via TS(12a′-12′)) are via 19 (14-electron intermediate) are highly hindered and
relatively fast, the recoordination of ethylene or 2,6-lutidine inaccessible with the t-Bu-substituted ligand (Figure 5b).
toward the β-agostic intermediate 9′ (linear) or 13′ (branched) From the less accessible intermediate 12, on the other hand,
determines the selectivity between linear repropagation and methyl branch formation via TS(12-13a) is faster than the
branch formation. In the cases of both Me and t-Bu substituents preceding β-hydride elimination via TS(9a-12), suggesting that
on the phosphorus atom, branch formations via TS(13′-14′) isomerization via 12 cannot be excluded under low pressure of
and TS(13′-20′) are slightly lower in energy than the ethylene.
corresponding linear pathways via TS(9′-10′) and TS(8′-9′). 2.1.4. Consideration I: Linearity of the Polyethylene. In
The relative stabilities of intermediates 9′ and 13′ (13′ is more order to summarize the first part, the key transition states for
stable than 9′ by ca. 3 kcal/mol) agree with this trend. This linear propagation, β-hydride elimination (BHE), subsequent
6105 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

Scheme 3. Selected Key Intermediates and Transition States for Linear Propagation, β-Elimination, Chain-Transfer Reactions,
and Methyl-Branch Formation

methyl-branch formation, and chain-transfer reactions are When the branch formation via trans-P−PD−hydride species
shown in Scheme 3. The linear chain of polyethylene is 12 is negligible, the linearity should reflect the relative rate of
formed by the linear PE formation cycle involving intermediates linear propagation via TS(10′-11′) versus frequency of the
9, 10, 10′, and 11′ (left cycle). Both of the side reactions, entry to the BHE valley, since branch formation and chain-
methyl-branch formation and chain-transfer reaction, arise from transfer reactions would be dominant over linear chain
the BHE valley involving intermediates 9′, 9a′, 12′, 12a′, 13′, propagation from the BHE valley involving 9′. Therefore, in
and 13a′. Due to the electronic asymmetry of the the presence of enough ethylene which would diminish the role
phosphine−sulfonate ligands, entrance to the BHE valley is of 2,6-lutidine-mediated TS(8′-9′) and ethylene-dissociated
facilitated by a fourth neutral ligand such as ethylene via 9NA′, the linearity should be in principle proportional to
TS(10′-9′) or 2,6-lutidine via TS(8′-9′). In the case of a t-Bu ΔΔG⧧ between TS(10′-11′) and TS(9′-10′). Under such
substituent on the phosphorus atom, spontaneous ethylene conditions, the linearity of polyethylene does not depend on
dissociation via 9NA′ should be considered for the entry to the the ethylene concentration, since both pathways start from the
BHE valley as well. After entry into the BHE valley, methyl- same intermediate 10′. On the other hand, when the
branch formation (right) and chain-transfer reactions (middle) contribution of the 2,6-lutidine-mediated transition state and
are feasible pathways. The catalyst goes back to the linear PE spontaneous ethylene dissociation to enter the BHE valley
formation cycle after methyl-branch formation (right), and becomes significant, the ratios of the methyl-branch formation
intermediate 17′ formed after chain-transfer reactions (middle) and chain-transfer reactions may be affected by the ethylene
can also reinitiate the linear PE formation. In addition to the pressure.
pathways from the BHE valley, branch forming via trans- 2.1.5. Consideration II: Molecular Weight of the Poly-
P−PD−hydride species 12 (upper pathway) cannot be ethylene. The molecular weight of polyethylene is determined
excluded, especially with the t-Bu substituent on the by the frequency of entering the BHE valley and the ratio of the
phosphorus atom. subsequent olefin reinsertion vs chain transfer. The favorable
Here the linearity of polyethylene prepared by a catalyst is mechanism of chain transfer depends on the size of the
defined as substituent on the phosphorus atom, such as Me or t-Bu (vide
infra); some catalysts favor an ethylene-mediated associative
chain transfer via TS(12′-17′), and others favor a dissociative
linearity = ln(DP of linear chain in polyethylene)
chain transfer via 19′. If the chain-transfer reactions dominate
DP of polyethylene over the linear and branch propagation pathways from the BHE
= ln
1(=chain end per chain) + branches per chain valley, the molecular weight of the obtained polyethylene
without methyl branches is simultaneously determined by the
where DP is the degree of polymerization. linearity of the obtained polyethylene in the presence of the
6106 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

Table 1. Effect of Substituents on the Phosphorus Atom on Ethylene Polymerization Activity, Molecular Weight, and
Microstructure of Polyethylenea

catalyst conditions result of polymerization


[cat.] ethylene temp activityb methyl branches on methyl branches per
entry R L (mM) (MPa) (°C) (kg mol−1 h−1) Mnc (103) PDIc 1000Cd chain linearitye
1f t-Bu lut 0.10 3.0 80 1860 6.2 4.1 1.0 0.48 5.1
2 t-Bu lut 0.10 1.0 80 663 4.5 6.2 1.2 0.38 4.8

3f i-Pr lut 0.10 3.0 80 641 6.7 2.7 0 0 5.5


4 i-Pr lut 0.10 1.0 80 497 8.4 2.6 0 0 5.7

5f Cy lut 0.10 3.0 80 1150 11 2.4 0 0 6.0


6 Cy lut 0.10 1.0 80 280 12 2.8 0 0 6.1
7g Cy lut 0.10 3.0 80 121 12 2.9 0 0 6.1

8f Men lut 0.10 3.0 80 205 170 1.5 0 0 8.7


9 Men lut 0.10 1.0 80 123 29 3.7 0 0 7.0

10 Ph lut 0.10 3.0 80 533 5.0 2.2 0 0 5.2


11 Ph lut 0.10 1.0 80 167 5.1 2.1 0 0 5.2
12h Ph NaCl 0.046 0.5 80 1820 1.5j 2.2 0.6 0.06 3.9
13 o-Ani lut 0.10 3.0 80 631 14 2.1 0.55 0.53 5.8
14 o-Ani lut 0.10 1.0 80 229 13 2.6 0.83 0.79 5.6
15h o-Ani NaCl 0.039 0.5 80 2750 12j 2.1 3 2.5 4.8

16i biAr py 0.050 2.1 85 1040 230k 3.9 n.r. n.r. (9.0)
17h biAr NaCl 0.034 0.5 80 5330 29j 2.8 0 0 6.9
a
A mixture of catalyst 1 (10 μmol) in toluene (100 mL) was stirred under an ethylene atmosphere (3.0−1.0 MPa) in a 300 mL autoclave for 1.0 h at
80 °C, unless otherwise noted. Abbreviations: Cy, cyclohexyl; Men, (−)-menthyl; o-Ani, 2-MeOC6H4; biAr, 2-[2′,6′-(MeO)2C6H3]C6H4; lut, 2,6-
lutidine; py, pyridine; n.r., not reported. bBased on isolated yields after precipitation with methanol. cMolecular weights determined by size-exclusion
chromatography (SEC) using polystyrene standards and corrected by universal calibration. dThe number of methyl branches determined by
quantitative 13C NMR analyses. eln(DP of linear chain in polyethylene) = ln(MW of polyethylene) − ln(1 + branches per chain) − ln(MW of
ethylene). See also section 2.1 for the definition. Values in parentheses represent that no data for the branch ratio were available. fFrom ref 31. g9.0
equiv of additional lutidine (90 μmol) was added. hFrom ref 27c. iFrom ref 28d. jMolecular weights determined by SEC using polyethylene
standards. kMolecular weights determined by light scattering analysis.

catalysts. On the other hand, catalysts that can form methyl reported results, we performed the polymerization reaction
branches from the BHE valley can form higher molecular under 1.0 MPa of ethylene pressure, in order to probe the
weight polyethylene in comparison to that estimated from their influence of ethylene pressure on the architecture of the
linearity. obtained polyethylene.
2.2. Effect of Various Substituents on the Phosphorus In Table 1, entries 1 and 2, with t-Bu substituents on the
Atom on the Key Steps. In order to confirm the theoretical phosphorus atom, a slight decrease in molecular weight was
results obtained in section 2.1, the effects of the substituents observed upon lowering the ethylene pressure. Interestingly, a
present on the phosphorus atom on the molecular weight and slight increase in the branches per chain was observed under a
linearity of the obtained polyethylene were experimentally higher pressure of ethylene (0.48 under 3.0 MPa of ethylene;
examined with varying ethylene pressure. The observed 0.38 under 1.0 MPa of ethylene). On the other hand, the
substituent trend in the experimental study was further studied catalysts bearing i-Pr (entries 3 and 4) and Cy (entries 5−7)
theoretically by computing the energetics for the key substituents on the phosphorus atom showed little response to
intermediates and transition states depicted in Scheme 3. ethylene pressure, in terms of the molecular weight and
2.2.1. Experimental Part. The experimental data of ethylene microstructure of polyethylene without methyl branches. It is
polymerization with palladium/phosphine−sulfonate catalysts noteworthy that the presence of an additional 9 equiv of 2,6-
bearing the alkyl substituents t-Bu, i-Pr, cyclohexyl (Cy), and lutidine which corresponds to a 10 times higher concentration
menthyl (Men) and aryl substituents Ph, 2-MeOC6H4 (o-Ani), of 2,6-lutidine did not change the structure of polyethylene
and 2-[2′,6′-(MeO)2C6H3]C6H4 (biAr) on the phosphorus obtained using the palladium catalyst bearing the Cy
atom are summarized in Table 1. The results in entries 1, 3, 5, substituent on the phosphorus atom (compare entries 5 and
7, 8, 12, and 15−17 were taken from the literature (see the 7). Given that the increase of 2,6-lutidine concentration
footnotes in Table 1 for the data sources). In addition to the corresponds to a relative decrease of ethylene concentration,
6107 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

the effect of 2,6-lutidine on the microstructure of polyethylene on the phosphorus atom did not form methyl-branched
may be negligible under the experimental conditions in Table 1 polyethylene even under low ethylene pressure. The Ph-
(3.0−0.5 MPa of ethylene with 0.10 mM of catalyst). Using substituted catalysts exhibited trace branch formation under
Men substituents on the phosphorus atom (entries 8 and 9), low ethylene pressure (0.5 MPa), but the branch number per
which can dramatically increase the molecular weight of chain was as low as 0.06. These experimentally observed
polyethylene,31a the molecular weight and linearity of the features of each substituent were studied theoretically by
obtained polyethylene were decreased by lowering the ethylene evaluating the energetics of the key species shown in Scheme 3.
pressure, whereas no detectable methyl branch was formed in 2.2.2. Theoretical Part. The results of the computational
both cases. study on the key transition states and intermediates listed in
In the case of ligands bearing aryl substituents, plentiful Scheme 3 are summarized in Table 2. In this theoretical study,
experimental data were provided by many researchers, and even the propagating chain end is modeled by n-pentyl instead of n-
fourth ligand free conditions were established by Mecking and propyl for a better estimation of the steric bulkiness from a
co-workers.27b With Ph substituents on the phosphorus atom wide range of substituents on the phosphorus atom at positions
(Table 1, entries 10−12), the reported molecular weights and farther away from the palladium center. For a linear PE
calculated linearity differed in the studies by different authors, formation cycle, the ethylene insertion reaction via TS(10′-11′)
whereas our experiments (entries 10 and 11) suggest no was the turnover-limiting step with all of the examined
dependence on the ethylene pressure. At 5 bar of ethylene substituents, not the cis/trans isomerization via TS(10−10′).
pressure (entry 12), low linearity and formation of a slight After entry to the BHE valley, all the substituents examined
amount of methyl branches were observed by Mecking and co- seem to favor branch formation or chain transfer in comparison
workers,27c which suggests the contribution of non-ethylene- to linear repropagation, which means that the entry to the BHE
mediated pathways for methyl-branch formation. The intro- valley should result in the decrease of linearity by methyl-
duction of o-Ani substituents on the phosphorus atom (entries branch formation or chain termination (compare the rows of
13−15) increased the methyl branches in polyethylene even entry to BHE valley, chain transfer, and branch formation in
under a high pressure of ethylene. The branch-forming nature Table 2). In the following sections, the effect of the substituents
was pronounced under lower pressures of ethylene, which led present on the phosphorus atom on the microstructure of
to the decrease in the linearity of polyethylene but not the polyethylene classified in Figure 6 is discussed on the basis of
molecular weight. The biAr substituent (entries 16 and 17), the mechanistic features obtained from Table 2.
which effectively increases the molecular weight of poly- 2.2.2.1. Substituent Effect on the Linearity. Although the
ethylene, showed an effect similar to that of the Men computational methods employed here do not reach a
substituent; the molecular weight and linearity of the obtained quantitative level, the observed experimental linearity of
polyethylene were decreased by lowering the ethylene pressure. polyethylene (see the columns of linearity in Table 1)
In a summary of the experimental study in Table 1, the correlated well with the calculated ΔΔG⧧ values between
substituents on the phosphorus atom are classified according to TS(10′-11′) and TS(9′-10′) (see the row of linearity in Table
the linearity and the experimental response to varied ethylene 2); the computational trend of linearity is in excellent
pressure (Figure 6). The experimental order of linearity was Ph agreement with the experimental trend.
experimental: Ph ≈ t ‐Bu < o‐Ani ∼ i‐Pr < Cy
< Men ≈ biAr

computed: Ph < t ‐Bu < o‐Ani < i‐Pr < Cy < Men
< biAr
2.2.2.2. Substituent Effect on the Molecular Weight
Dependence on Ethylene Pressure. The molecular weight
dependence on the ethylene pressure used in the polymer-
ization, which varied with the substituent on the phosphorus
atom (see also Table 1 and Figure 6), could be explained by the
free energy difference between the pathways for the entrance to
the BHE valley, namely ethylene-associated TS(9′-10′) and
Figure 6. Classification of the substituents on the phosphorus atom spontaneously dissociated 9NA′, and the free energy difference
based on the experimental response to varied ethylene pressure. between the associative and dissociative chain transfers. For the
associative chain transfer, we found not only TS(12′-17′) for
≈ t-Bu < o-Ani ≈ i-Pr < Cy < Men < biAr. In terms of the the conversion between 12′ and 17′ but also the pentacoordi-
molecular weight dependence on the ethylene pressure, i-Pr-, nated intermediate 18′ structurally similar to TS(12′-17′).
Cy-, Ph-, and o-Ani-substituted catalysts showed no significant Given that the dissociation of each alkene from 18′ requires 1−
dependence on the ethylene pressure around 1.0−3.0 MPa. On 2 kcal/mol activation energy, we treated both TS(12′-17′) and
the other hand, the molecular weights of polyethylene formed 18′ as the barriers for the associative chain-transfer reaction,
by the catalysts bearing t-Bu, Men, and biAr substituents on the just as we treated 19′ as the barrier for the dissociative chain
phosphorus atom were affected by the ethylene pressure even transfer. (see also the description about 18Me in section
at 1.0 MPa. For the branch-forming tendency, only o-Ani and t- 2.1.3.2).
Bu substituents can facilitate the formation of an observable The catalysts that showed no response to varied ethylene
degree of branches even under a high pressure of ethylene, concentration (R = i-Pr, Cy, Ph, o-Ani) favor the ethylene-
whereas catalysts bearing i-Pr, Cy, Men, and biAr substituents associated TS(9′-10′) for entrance to the BHE valley and 18′
6108 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

Table 2. ΔG, ΔG⧧, and ΔΔG⧧ Values of the Key Transition States and Intermediates from the Corresponding 10 (kcal/mol)a

a
B3LYP-D3/6-31G* and lanl2dz for palladium in toluene. The propagating chain end is modeled by n-pentyl with all-anti configuration, instead of
n-propyl for a better estimation of steric bulk at a position farther from the substituents on the phosphorus atom. Abbreviations: E, ethylene; BHE, β-
hydride elimination. Legend: (*) 18′; (**) coordination of ipso carbon was observed.

for chain transfer, rather than dissociative 9NA′ and 19′ (see 14′ to 13NA′. Actually, in the case of the t-Bu substituent, a
the rows of linearity dependence on ethylene pressure, slight increase in the methyl branches per chain was
associative vs dissociative chain transfer in Table 2). Therefore, experimentally observed under higher pressures of ethylene.
all the steps that determine the molecular weight of polymers, 2.2.2.3.2. i-Pr, Cy, and Ph. The catalysts bearing i-Pr, Cy,
namely chain propagation via TS(10′-11′), the entrance to the and Ph substituents on the phosphorus atom produced highly
BHE valley via TS(9′-10′), and the chain transfer via 18′, linear polyethylene and showed no response to variations in
should show a first-order dependence on ethylene concen- ethylene pressure (entries 3−7 and 10−12 in Table 1). On the
tration, and thereby the effect of ethylene activity is canceled. basis of the computation in Table 2, the formation of trans-P−
On the other hand, the catalysts that produce lower Pd−H species via TS(9a-12) and entry to the BHE valley via
molecular weight polyethylene at decreased ethylene pressure 9NA′ are less likely, since those required a larger activation
(R = t-Bu, Men, biAr) apparently favor dissociative entry to the energy in comparison to the linear propagation via TS(10′-11′)
BHE valley via 9NA′ rather than TS(9′-10′) with associated and entry to the BHE valley via TS(9′-10′). From the BHE
ethylene (see the row of linearity dependence on ethylene valley, these substituents facilitate both associative chain
pressure in Table 2). Therefore, at a lower ethylene pressure transfer via TS(12′-17′) and dissociative path via 19′ over
that decelerates the linear propagation cycle, the frequency of the branch-forming pathway via TS(13′-14′) and 13NA′.
chain transfer can be pronounced. Hence, little branch formation occurs from the BHE valley.
2.2.2.3. Substituent Effect on Methyl-Branch Formation. 2.2.2.3.3. o-Ani. The catalyst bearing o-Ani substituents on
The trend of each catalyst to form methyl branches is discussed the phosphorus atom can produce slightly branched poly-
separately on the basis of the classification in Figure 6. ethylene even under high ethylene pressure, and the branches
2.2.2.3.1. t-Bu. The catalyst bearing t-Bu substituents on the per chain was increased at lower ethylene pressures (entries
phosphorus atom produced slightly branched polyethylene 13−15 in Table 1). From the computational results in Table 2,
even under high ethylene pressures (entries 1 and 2 in Table a characteristic of this catalyst is the low-energy TS(9a-12),
1). On the basis of the computation in Table 2, the formation comparable to TS(9′-10′), the favored entry to the BHE valley.
of trans-P−Pd−H species via TS(9a-12) is less likely, since it Therefore, the branch formation with the catalyst possibly
required a larger activation energy than the linear propagation occurs via the trans-P−Pd−H species 12 (zeroth order to
via TS(10′-11′) and the entry to the BHE valley via TS(9′-10′) ethylene pressure) and not via the BHE valley through 9′/9a′
or via 9NA′. From the BHE valley, ethylene coordination to and 12′/12a′. This explanation is consistent with the
13NA′ to give 14′ proceeds with little activation barrier and is experimental observation of increased branches per chain at
in competition with associative and dissociative chain transfers. lowered ethylene pressure.
Therefore, a higher concentration of ethylene may increase the 2.2.2.3.4. Men and biAr. In the case of Men and biAr
branches per chain, due to a relative decrease in the rate of substituents, while the formation of methyl branches was not
dissociative chain transfer via 19′ and reverse conversion from experimentally detectable (entries 8, 9, 16, and 17 in Table 1),
6109 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

the computational results suggest that methyl-branch formation Scheme 5. Plausible Mechanism of Deuterium Incorporation
after β-hydride elimination is feasible under high pressures of from MeOD into Ethylene42
ethylene. This discrepancy may arise from some overlooked
factors in the small set of calculations, such as conformational
sampling or entropic effects, or uncovered advent effects from
these huge substituents on the phosphorus atom toward chain-
transfer process.
2.2.3. Summary of the Substituent Effect on the
Polyethylene Formation. In section 2.2, the effect of the
substituents present on the phosphorus atom of the palladium/
phosphine−sulfonate catalysts on the mechanism of linear
polyethylene formation was experimentally and theoretically
studied. Experimentally, several polymerization reactions were
carried out under varied ethylene pressures, in order to
investigate the mechanism of linear polyethylene formation.
Theoretically, the key transition states and intermediates in
Scheme 3 were examined with various substituents on the
phosphorus atom. The experimental behaviors of each catalyst,
including linearity, tendency to form methyl branches, and
response to varied ethylene pressure, were qualitatively 23 and the following C−C bond rotation and β-hydride (not
described by the relative free energies of the key species. deuteride) elimination affords 25, having deuterated ethylene.
Therefore, the key species proposed in Scheme 3 are very likely Finally, an ethylene exchange reaction releases the deuterated
to constitute the main catalytic pathways for the linear ethylene and regenerates the starting complex 17′-H.
propagation, β-hydride elimination, chain transfer, and This proposal also leads to an alternative mechanism of
methyl-branch-forming reactions. deuterium incorporation into polyethylene from MeOD, which
2.3. Mechanism of Deuterium Incorporation from is compatible with the mechanism of linear polyethylene
MeOD into Polyethylene: H/D Exchange Reaction of 1- formation detailed in sections 2.1 and 2.2. The palladium−
Alkenes Catalyzed by Palladium Hydride Species. One of hydride species 17′ formed after β-elimination and chain
the questions remaining in our proposed mechanism in Scheme transfer may be able to catalyze deuteration of ethylene as
2 is the mechanism of deuterium incorporation into the main proposed in Scheme 4, before 23 and 24 reinitiate linear
chain of polyethylene from MeOD used as a deuterium source polyethylene formation. Therefore, the deuterium inclusion
(Scheme 2).36 Although reversible β-hydride elimination and into polyethylene from MeOD can be possible via the
H/D exchange reactions during polymerization were proposed formation and incorporation of deuterated ethylene into linear
in the previous paper,36 our computation and experimental polyethylene formation.
observation in sections 2.1 and 2.2 suggest that, once β-hydride When 1-eicosene was used as the substrate for the hydrogen/
elimination occurred, the chain transfer or the branch deuterium exchange reaction, the C1 position of the 1-alkene
formation dominates over the linear propagation. In this part, was slowly but selectively deuterated (Scheme 6). The
we propose an alternative mechanism of deuterium incorpo-
ration, on the basis of the experimental observation of a
Scheme 6. Deuteration of C1 Position of 1-Alkene by a Pd/
hydrogen/deuterium exchange reaction of 1-alkenes catalyzed
Phosphine−Sulfonate Complex
by palladium hydride species.
When a solution of [o-Ani2PC6H4SO3]PdH[P(t-Bu)3]41 (22-
H) in benzene was treated with atmospheric pressure of
ethylene and MeOD, catalytic deuteration of ethylene was
confirmed by quantitative NMR analyses (Scheme 4). On the

Scheme 4. Deuteration of Ethylene by a Palladium/


Phosphine−Sulfonate Complex

selectivity for the C1 position reached ca. 30-fold, which


could originate from the free energy surface of the BHE valley.
We propose that the insertion of propylene into palladium
hydride species 12a′ or 12′ favors the 2,1-fashion to give 13′,
because 13′ is more stable than 9′ by 2.1 kcal/mol mainly due
basis of the partial deuteration of the palladium hydride species to steric effects (Scheme 7). In addition, the 2,1-insertion via
(Scheme 4), as previously observed by Mecking and co- TS(12′-13′) is lower in energy than the 1,2-insertion via
workers,12b we propose a plausible mechanism for the TS(9′-12a′) by 1.0−1.5 kcal/mol for the (R2PC6H4SO3)PdPr
deuteration of ethylene (Scheme 5). The starting palladium complexes (R = Me, t-Bu) (Figures 3 and 4).
hydride complex 22-H undergoes a hydride/deuterium
exchange reaction with MeOD,12b and the phosphine ligand 3. CONCLUSION
in 22-D can be replaced by the excess amount of alkenes In this study, we theoretically and experimentally investigated
present. Once 17′-D was formed, insertion of ethylene to give the mechanism of linear polyethylene formation by palladium/
6110 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

Scheme 7. Plausible Explanation for the C1 Selectivity Scheme 8. Role of the Phosphine−Sulfonate Ligand on
Palladium-Catalyzed Linear Polyethylene Formation

phosphine−sulfonate catalysts. In the first part, dispersion-


corrected DFT studies on the entire mechanism of poly-
ethylene formation starting from (R2PC6H4SO3)PdMe(2,6-
lutidene) (R = Me, t-Bu) were carried out. Alkylpalladium
ethylene complex 10′ is the key intermediate for both linear
propagation and β-hydride elimination from the growing
polymer chain: ethylene insertion to 10′ for linear propagation
takes place via TS(10′-11′), and β-hydride elimination
proceeds from 9′, which is formed from 10′ either associatively
via five-coordinated TS(9′-10′) or dissociatively via three-
coordinated intermediate 9NA′. Once β-hydride elimination
took place via TS(9′-12a′), associative and dissociative chain-
transfer reactions (via TS(12′-17′) and 19′, respectively)
dominate over the branch formation via TS(13′-14′) and
TS(14′-15′), for both Me- and t-Bu-substituted phosphine−-
sulfonate ligands.
In section 2.2, the effects of various substituents on the
phosphorus atom (R = Me, t-Bu, i-Pr, Cy, Men, Ph, o-Ani,
biAr) were investigated theoretically and experimentally. The
effect of the substituents on the phosphorus atom was classified
by experimentally observed linearity and the response to varied
ethylene pressure. The obtained relative free energies of the key
species proposed in section 2.1 reasonably explained the
experimental behavior of the corresponding catalysts (Figure and a weak σ donor promotes linear polyethylene formation
6). The linearity of polyethylene was correlated to the relative after cis/trans isomerization, suppressing (1) the formation of
activation energy difference between TS(10′-11′) for ethylene trans-P−Pd−H species by a strong trans influence from the
insertion and TS(9′-10′) for ethylene dissociation, both after phosphine donor (electronic inhibition) and (2) the formation of
cis/trans isomerization. The mechanism of entering the BHE cis-P−Pd−H species by retarding the prior ethylene dissocia-
valley leading to the chain-transfer reaction and branch tion (steric inhibition). In the case of the related palladium/α-
formation differed by steric repulsion from substituents on diimine catalysts, steric coverage of the axial orientation of the
the phosphorus atom. The catalysts that produce constant palladium center was proposed to hinder the associative chain-
molecular weights of polyethylene under varied ethylene transfer process (Scheme 1).33 On the other hand, in the case
pressures (i-Pr, Cy, Ph, and o-Ani substituents) were computa- of the palladium/phosphine−sulfonate catalysts, steric bulki-
tionally suggested to favor the ethylene-associated entry to the ness provided by substituents on the phosphorus atom can
BHE valley via TS(10′-11′) and undergo associative chain retard not only the associative chain-transfer process but also
transfer. In contrast, the catalysts that showed a dependence on ethylene dissociation via TS(9′-10′) prior to the β-hydride
the ethylene pressures (t-Bu, Men, and biAr substituents) elimination reaction. The importance of the ethylene
favored entry into the BHE valley by the spontaneous dissociation transition state for linear polyethylene formation
dissociation of ethylene via 9NA′. Furthermore, the tendency was recently proposed by Jensen in the case of nickel/
of slight methyl-branch formation of t-Bu- and o-Ani- salicylaldiminate catalysts.40 Our computation revealed that the
substituted catalysts was also explained by the energies of a transition states can be controlled by steric inhibition, and the
small set of key transition states. linearity of polyethylene can be an experimental indicator of the
In section 2.3, we found that [o-Ani2PC6H4SO3]PdH[P(t- transition states. The occurrence of the dissociative ethylene
Bu) 3] catalyzed the C-1 selective hydrogen/deuterium dissociation process via 9NA′, when bulky Men- and biAr-type
exchange reaction of 1-alkenes using MeOD as the deuterium substituents are present on the phosphorus atom, represents an
source. On the basis of this observation, a new mechanism has intrinsic limitation of the efforts to block associative ligand
been proposed for the incorporation of deuterium into ethylene exchange processes but also indicates the possibility of
and eventual incorporation of deuterated ethylene, formed by a variability of molecular weights of polyethylene and ethylene/
small amount of Pd−H species, into the main chain of polar monomer copolymers by tuning reaction conditions.
polyethylene. We believe that the insights obtained here can be the basis
The correspondence of computations and experiments with for further in silico modification of the phosphine−sulfonate
various phosphine−sulfonate ligands shows how the unsym- ligands and can also help in designing brand-new unsymmetric
metric nature of the phosphine−sulfonate ligands affects linear ligands for olefin polymerization. Sufficient electronic dis-
polyethylene formation with a group 10 metal center (Scheme symmetry as well as construction of a steric environment
8). The combination of a strong and sterically bulky σ donor around the strong donor is essential for linear polymerization of
6111 DOI: 10.1021/acscatal.6b00911
ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

simple olefins, which is shared in the design of our recent works (8) For a review, see: Nakamura, A.; Anselment, T. M. J.; Claverie, J.;
on palladium/BPMO20 and palladium/IzQO21 catalysts. Goodall, B.; Jordan, R. F.; Mecking, S.; Rieger, B.; Sen, A.; van
Leeuwen, P. W. N. M.; Nozaki, K. Acc. Chem. Res. 2013, 46, 1438−
4. COMPUTATIONAL METHODS 1449.
(9) (a) Guironnet, D.; Roesle, P.; Rünzi, T.; Göttker-Schnetmann, I.;
All calculations were performed using the Gaussian 09 Mecking, S. J. Am. Chem. Soc. 2009, 131, 422−423. (b) Guironnet, D.;
packages.43 The DFT (B3LYP-D3) methods chosen were Caparaso, L.; Neuwald, B.; Göttker-Schnetmann, I.; Cavaro, L.;
used primarily with 6-31G* basis sets44 for the light atoms and Mecking, S. J. Am. Chem. Soc. 2010, 132, 4418−4426.
Lanl2DZ basis sets and effective core potential for palladium.45 (10) (a) Rünzi, T.; Fröhlich, D.; Mecking, S. J. Am. Chem. Soc. 2010,
All local minima and saddle points were confirmed by their 132, 17690−17691. (b) Daigle, J.-C.; Piche, L.; Claverie, J. P.
vibrational frequency calculations (with zero and one imaginary Macromolecules 2011, 44, 1760−1762.
frequencies, respectively). The saddle points found were (11) (a) Kochi, T.; Noda, S.; Yoshimura, K.; Nozaki, K. J. Am. Chem.
confirmed to be the correct ones by visualizing the vibrational Soc. 2007, 129, 8948−8949. (b) Nozaki, K.; Kusumoto, S.; Noda, S.;
mode of the imaginary frequency with GaussView. To include Kochi, T.; Chung, L. W.; Morokuma, K. J. Am. Chem. Soc. 2010, 132,
solvation effects (toluene as the solvent), the SMD solvation 16030−16042.
model38 was used. (12) (a) Weng, W.; Shen, Z.; Jordan, R. F. J. Am. Chem. Soc. 2007,


129, 15450−15451. (b) Leicht, H.; Göttker-Schnetmann, I.; Mecking,
ASSOCIATED CONTENT S. Angew. Chem., Int. Ed. 2013, 52, 3963−3966.
(13) Luo, S.; Vela, J.; Lief, G. R.; Jordan, R. F. J. Am. Chem. Soc. 2007,
* Supporting Information
S
129, 8946−8947.
The Supporting Information is available free of charge on the (14) Ito, S.; Munakata, K.; Nakamura, A.; Nozaki, K. J. Am. Chem.
ACS Publications website at DOI: 10.1021/acscatal.6b00911. Soc. 2009, 131, 14606−14607.
Experimental procedures and full characterization data of (15) (a) Ito, S.; Kanazawa, M.; Munakata, K.; Kuroda, J.; Okumura,
polymers (PDF) Y.; Nozaki, K. J. Am. Chem. Soc. 2011, 133, 1232−1235. (b) Jian, Z.;
Cartesian coordinates of optimized species and compar- Mecking, S. Angew. Chem., Int. Ed. 2015, 54, 15845−15849.
ison of computational methods (PDF) (16) (a) Guan, Z.; Cotts, P. M.; McCord, E. F.; McLain, S. J. Science


1999, 283, 2059−2062. (b) Guo, L.; Dai, S.; Sui, X.; Chen, C. ACS
Catal. 2016, 6, 428−441.
AUTHOR INFORMATION (17) (a) Nowack, R. J.; Hearley, A. K.; Rieger, B. Z. Anorg. Allg.
Corresponding Authors Chem. 2005, 631, 2775−2781. (b) Guironnet, D.; Rünzi, T.; Göttker-
*E-mail for K.M.: morokuma.keiji.3a@kyoto-u.ac.jp. Schnetmann, I.; Mecking, S. Chem. Commun. 2008, 40, 4965−4967.
*E-mail for K.N.: nozaki@chembio.t.u-tokyo.ac.jp. (c) Zhou, X.; Bontemps, S.; Jordan, R. F. Organometallics 2008, 27,
4821−4824. (d) Noda, S.; Kochi, T.; Nozaki, K. Organometallics 2009,
Notes
28, 656−658. (e) Zhang, D.; Wang, J.; Yue, Q. J. Organomet. Chem.
The authors declare no competing financial interest.


2010, 695, 903−908. (f) Perrotin, P.; McCahill, J. S. J.; Wu, G.; Scott,
S. L. Chem. Commun. 2011, 47, 6948−6950. (g) Ito, S.; Ota, Y.;
ACKNOWLEDGMENTS Nozaki, K. Dalton Trans. 2012, 41, 13807−13809.
This work was partially supported by CREST, JST. The (18) (a) Younkin, T. R.; Connor, E. F.; Henderson, J. I.; Friedrich, S.
theoretical calculations were performed using computational K.; Grubbs, R. H.; Bansleben, D. A. Science 2000, 287, 460−462.
resources provided by the Research Center for Computational (b) Hicks, F. A.; Brookhart, M. Organometallics 2001, 20, 3217−3219.
Science, National Institutes of Natural Sciences, Okazaki, Japan. (c) Jenkins, J. C.; Brookhart, M. J. Am. Chem. Soc. 2004, 126, 5827−
R.N. is grateful to the Japan Society for the Promotion of 5842. (d) Bastero, A.; Göttker-Schnetmann, I.; Röhr, C.; Mecking, S.
Science (JSPS) for a Research Fellowship for Young Scientists. Adv. Synth. Catal. 2007, 349, 2307−2316. (e) Berkefeld, A.; Drexler,


M.; Möller, H. M.; Mecking, S. J. Am. Chem. Soc. 2009, 131, 12613−
REFERENCES 12622. (f) Berkefeld, A.; Mecking, S. J. Am. Chem. Soc. 2009, 131,
1565−1574. (g) Soshnikov, I. E.; Semikolenova, N. V.; Zakharov, V.
(1) (a) Jeremic, D. In Ullmann’s Encyclopedia of Industrial Chemistry;
A.; Möller, H. M.; Ö lscher, F.; Osichow, A.; Göttker-Schnettmann, I.;
Wiley-VCH: Weinheim, Germany, 2014. (b) Gahleitner, M.; Paulik, C.
In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Mecking, S.; Talsi, E. P.; Bryliakov, K. P. Chem. - Eur. J. 2013, 19,
Weinheim, Germany, 2014. 11409−11417. (h) Osichow, A.; Göttker-Schnetmann, I.; Mecking, S.
(2) (a) Boor, J., Jr. Ziegler-Natta Catalysts and Polymerizations; Organometallics 2013, 32, 5239−5242. (i) Osichow, A.; Rabe, C.;
Academic Press: New York, 1979. (b) McMillan, F. M. The Chain Vogtt, K.; Narayanan, T.; Harnau, L.; Drechsler, M.; Ballauff, M.;
Straighteners; The MacMillan Press: London, 1979. Mecking, S. J. Am. Chem. Soc. 2013, 135, 11645−11650.
(3) (a) Boaen, N. K.; Hillmyer, M. A. Chem. Soc. Rev. 2005, 34, 267− (19) (a) Kuhn, P.; Sémeril, D.; Matt, D.; Chetcuti, M. J.; Lutz, P.
275. (b) Chung, T. C. Functionalization of Polyolefins; Academic Press: Dalton Trans. 2007, 515−528. (b) Shimizu, F.; Shin, S.; Tanna, A.;
London, 2002. Goromaru, S.; Matsubara, Y. WO 2010/050256, 2010.
(4) Boffa, L. S.; Novak, B. M. Chem. Rev. 2000, 100, 1479−1493. (20) (a) Carrow, B. P.; Nozaki, K. J. Am. Chem. Soc. 2012, 134,
(5) (a) Ittel, S. D.; Johnson, L. K.; Brookhart, M. Chem. Rev. 2000, 8802−8805. (b) Mitsushige, Y.; Carrow, B. P.; Ito, S.; Nozaki, K.
100, 1169−1203. (b) Gibson, V. C.; Spitzmesser, S. K. Chem. Rev. Chem. Sci. 2016, 7, 737−744. (c) Sui, X.; Dai, S.; Chen, C. ACS Catal.
2003, 103, 283−315. (c) Sen, A.; Borkar, S. J. Organomet. Chem. 2007, 2015, 5, 5932−5937.
692, 3291−3299. (d) Nakamura, A.; Ito, S.; Nozaki, K. Chem. Rev. (21) (a) Nakano, R.; Nozaki, K. J. Am. Chem. Soc. 2015, 137, 10934−
2009, 109, 5215−5244. (e) Ito, S.; Nozaki, K. Chem. Rec. 2010, 10, 10937. See also: (b) Tao, W.-j; Nakano, R.; Ito, S.; Nozaki, K. Angew.
315−325. (f) Mu, H.; Pan, L.; Song, D.; Li, Y. Chem. Rev. 2015, 115, Chem., Int. Ed. 2016, 55, 2835−2839.
12091−12137. (22) Vela, J.; Lief, G. R.; Shen, Z.; Jordan, R. F. Organometallics 2007,
(6) (a) Johnson, L. K.; Mecking, S.; Brookhart, M. J. Am. Chem. Soc. 26, 6624−6635.
1996, 118, 267−268. (b) Mecking, S.; Johnson, L. K.; Wang, L.; (23) Skupov, K. M.; Piche, L.; Claverie, J. P. Macromolecules 2008, 41,
Brookhart, M. J. Am. Chem. Soc. 1998, 120, 888−899. 2309−2310.
(7) Drent, E.; van Dijk, R.; van Ginkel, R.; van Oort, B.; Pugh, R. I. (24) Haras, A.; Anderson, G. D. W.; Michalak, A.; Rieger, B.; Ziegler,
Chem. Commun. 2002, 744−745. T. Organometallics 2006, 25, 4491−4497.

6112 DOI: 10.1021/acscatal.6b00911


ACS Catal. 2016, 6, 6101−6113
ACS Catalysis Research Article

(25) Noda, S.; Nakamura, A.; Kochi, T.; Chung, L. W.; Morokuma, ethylene coordination to 23 with TS(13′-14′), a comparison between
K.; Nozaki, K. J. Am. Chem. Soc. 2009, 131, 14088−14100. the rows of associative chain transfer via TS(12′-17′) or 18′ and
(26) (a) Conley, M. P.; Jordan, R. F. Angew. Chem., Int. Ed. 2011, 50, ethylene coordination via TS(13′-14′) in Table 2 could be informative
3744−3746. (b) Zhou, X.; Lau, K.-C.; Petro, B. J.; Jordan, R. F. for the staying period of 17′-H in the H/D exchange cycle. The
Organometallics 2014, 33, 7209−7214. catalyst bearing an o-Ani-substituted ligand easily proceeds to branch
(27) (a) Friedberger, T.; Wucher, P.; Mecking, S. J. Am. Chem. Soc. formation from the BHE valley, since the chain transfer via 18′ is
2012, 134, 1010−1018. (b) Neuwald, B.; Ö lscher, F.; Göttker- slightly favored over ethylene coordination via TS(13′-14′) by 3.1
Schnetmann, I.; Mecking, S. Organometallics 2012, 31, 3128−3137. kcal/mol. On the other hand, in the case of a Cy-substituted ligand,
(c) Neuwald, B.; Falivene, L.; Caporaso, L.; Cavallo, L.; Mecking, S. the preference to the chain transfer via 18′ over ethylene coordination
Chem. - Eur. J. 2013, 19, 17773−17788. via TS(13′-14′) is more pronounced by 6.0 kcal/mol in comparison to
(28) (a) Skupov, K. M.; Marella, P. R.; Simard, M.; Yap, G. P. A.; that of the o-Ani substituted ligand by 3.1 kcal/mol. Therefore, the
Allen, N.; Conner, D.; Goodall, B. L.; Claverie, J. P. Macromol. Rapid palladium hydride species with Cy substituents on the phosphorus
Commun. 2007, 28, 2033−2038. (b) Skupov, K. M.; Hobbs, J.; atom can stay longer in the H/D exchange cycle in comparison to that
Marella, P.; Conner, D.; Golisz, S.; Goodall, B. L.; Claverie, J. P. with o-Ani substituents, leading to the more efficient deuteration of
Macromolecules 2009, 42, 6953−6963. (c) Piche, L.; Daigle, J.-C.; Poli, ethylene.
R.; Claverie, J. P. Eur. J. Inorg. Chem. 2010, 4595−4601. (d) Piche, L.; (43) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Daigle, J.-C.; Rehse, G.; Claverie, J. P. Chem. - Eur. J. 2012, 18, 3277− Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
3285. B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H.
(29) (a) Shen, Z.; Jordan, R. F. J. Am. Chem. Soc. 2010, 132, 52−53. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
(b) Shen, Z.; Jordan, R. F. Macromolecules 2010, 43, 8706−8708. Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima,
(c) Contrella, N. D.; Jordan, R. F. Organometallics 2014, 33, 7199− T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.;
7208. Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin,
(30) Jian, Z.; Wucher, P.; Mecking, S. Organometallics 2014, 33, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.;
2879−2888. Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega,
(31) (a) Ota, Y.; Ito, S.; Kuroda, J.; Okumura, Y.; Nozaki, K. J. Am. N.; Millam, M. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.;
Chem. Soc. 2014, 136, 11898−11901. See also: (b) Ota, Y.; Ito, S.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.;
Kobayashi, M.; Kitade, S.; Sakata, K.; Tayano, T.; Nozaki, K. Angew. Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.;
Chem., Int. Ed. 2016, 55, 7505−7509. Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.;
(32) (a) Anselment, T. M. J.; Anderson, C. E.; Rieger, B.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .;
Boeddinghaus, M. B.; Fässler, T. F. Dalton Trans. 2011, 40, 8304− Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09,
8313. (b) Anselment, T. M. J.; Wichmann, C.; Anderson, C. E.; Revision D.01; Gaussian, Inc., Wallingford, CT, 2009.
Herdtweck, E.; Rieger, B. Organometallics 2011, 30, 6602−6611. (44) Ditchfield, R.; Hehre, W. J.; Pople, J. A. J. Chem. Phys. 1971, 54,
(33) (a) Johnson, L. K.; Killian, C. M.; Brookhart, M. J. Am. Chem. 724−728.
Soc. 1995, 117, 6414−6415. (b) Camacho, D. H.; Salo, E. V.; Ziller, J. (45) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270−283.
W.; Guan, Z. Angew. Chem., Int. Ed. 2004, 43, 1821−1825.
(c) Popeney, C. S.; Levins, C. M.; Guan, Z. Organometallics 2011,
30, 2432−2452. (d) Allen, K. E.; Campos, J.; Daugulis, O.; Brookhart,
M. ACS Catal. 2015, 5, 456−464. (e) Dai, S. Y.; Sui, X. L.; Chen, C. L.
Angew. Chem. Int. Ed. 2015, 54, 9948−9953.
(34) (a) Musaev, D. G.; Svensson, M.; Morokuma, K.; Stromberg, S.;
Zetterberg, K.; Siegbahn, P. E. M. Organometallics 1997, 16, 1933−
1945. (b) Froese, R. D. J.; Musaev, D. G.; Morokuma, K. J. Am. Chem.
Soc. 1998, 120, 1581−1587.
(35) Rezabal, E.; Asua, J. M.; Ugalde, J. M. Organometallics 2015, 34,
373−380.
(36) Kanazawa, M.; Ito, S.; Nozaki, K. Organometallics 2011, 30,
6049−6052.
(37) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. J. Chem. Phys.
2010, 132, 154104.
(38) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. J. Phys. Chem. B
2009, 113, 6378−6396.
(39) Sato, Y.; Hosaka, N.; Inomata, H.; Kanaka, K. Fluid Phase
Equilib. 2013, 344, 112−116.
(40) The importance of the ethylene-dissociation TS was proposed
in related nickel catalysts. See: Heyndrickx, W.; Occhipinti, G.; Jensen,
V. R. Chem. - Eur. J. 2014, 20, 7962−7978.
(41) Rünzi, T.; Tritschler, U.; Roesle, P.; Göttker-Schnetmann, I.;
Möller, H. M.; Caporaso, L.; Poater, A.; Cavallo, L.; Mecking, S.
Organometallics 2012, 31, 8388−8406.
(42) According to the proposed mechanism for the formation of
deuterated ethylene, the higher D incorporation ratio by Cy-
substituted palladium/phosphine−sulfonate in comparison to that by
the o-Ani-substituted species can be explained as follows. Since the
molecular weights of polyethylene formed by these catalysts are
comparable, the frequency to generate 17′-H should be similar
between these catalysts. Then the degree of deuteration will be
determined by the rate of H/D exchange reaction and the staying
period of 17′-H in the H/D exchange cycle. Because of the similarities
of the alkene exchange TS(25-17′) with TS(12′-17′)/18′ and the

6113 DOI: 10.1021/acscatal.6b00911


ACS Catal. 2016, 6, 6101−6113

View publication stats

You might also like