You are on page 1of 12

pubs.acs.

org/accounts Article

Toward the Copolymerization of Propylene with Polar Comonomers


Stephen L. J. Luckham and Kyoko Nozaki*

Cite This: https://dx.doi.org/10.1021/acs.accounts.0c00628 Read Online

ACCESS Metrics & More Article Recommendations


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV OF NEW ENGLAND on November 28, 2020 at 09:39:02 (UTC).

CONSPECTUS: Polyolefins are produced in vast amounts and are found in so many consumer products that the two most
commonly produced forms, polyethylene (PE) and polypropylene (PP), fall into the rather sparse category of molecules that are
likely to be known by people worldwide, regardless of their occupation. Although widespread, the further upgrading of their
properties (mechanical, physical, aesthetic, etc.) through the formation of composites with other materials, such as polar polymers,
fibers, or talc, is of huge interest to manufacturers. To improve the affinity of polyolefins toward these materials, the inclusion of
polar functionalities into the polymer chain is essential. The incorporation of a functional group to trigger controlled polymer
degradation is also an emerging area of interest. Currently practiced methods for the incorporation of polar functionalities, such as
post-polymerization functionalization, are limited by the number of compatible polar monomers: for example, grafting maleic
anhydride is currently the sole method for practical functionalization of PP. In contrast, the incorporation of fundamental polar
comonomers into PE and PP chains via coordination insertion polymerization offers good control, making it a highly sought-after
process. Early transition metal catalysts (which are commonly used for the production of PE and PP) display poor tolerance toward
the functional groups within polar comonomers, limiting their use to less-practical derivatives. As late transition metal catalysts are
less-oxophilic and thus more tolerant to polar functionalities, they are ideal candidates for these reactions. This Account focuses on
the copolymerization of propylene with polar comonomers, which remains underdeveloped as compared to the corresponding
reaction using ethylene. We begin with the challenges associated with the regio- and stereoselective insertion of propylene, which is a
particular problem for late transition metal systems because of their propensity to undergo chain walking processes. To overcome
this issue, we have investigated a range of metal/ligand combinations. We first discuss attempts with group 4 and 8 metal catalysts
and their limitations as background, and then focus on the copolymerization of propylene with methyl acrylate (MA) using Pd/
imidazolidine−quinolinolate (IzQO) and Pd/phosphine−sulfonate (PS) precatalysts. Each generated regioregular polymer, but
while the system featuring an IzQO ligand did not display any stereocontrol, that using the chiral PS ligand did. A further difference
was found in the insertion mode of MA: the Pd/IzQO system inserted in a 1,2 fashion, while in the Pd/PS system a 2,1 insertion was
observed. We then move onto recent results from our lab using Pd/PS and Pd/bisphosphine monoxide (BPMO) precatalysts for the
copolymerization of propylene with allyl comonomers. These P-stereogeneic precatalysts generated the highest isotacticity values
reported to date using late transition metal catalysts. This section closes with our work using Earth-abundant nickel catalysts for the
reaction, which would be especially desired for industrial applications: a Ni/phosphine phenolate (PO) precatalyst yielded
regioregular polypropylene with the incorporation of some allyl monomers into the main polymer chain. The installation of a chiral
menthyl substituent on the phosphine allowed for moderate stereoselectivity to be achieved, though the applicable polar monomers
currently remain limited. The Account concludes with a discussion of the factors that affect the insertion mode of propylene and
polar comonomers in copolymerization reactions, beginning with our recent computational study, and finishing with work from
ourselves and others covering both comonomer and precatalyst steric and electronic profiles with reference to the observed
regioselectivity.

■ KEY REFERENCES
• Nakano, R.; Nozaki, K. Copolymerization of Propylene
Received: September 29, 2020

and Polar Monomers Using Pd/IzQO Catalysts. J. Am.


Chem. Soc. 2015, 137, 10934−10937.1 The first report
on the copolymerization of propylene with polar
monomers to give regioregular, atactic copolymers.

© XXXX American Chemical Society https://dx.doi.org/10.1021/acs.accounts.0c00628


A Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

Figure 1. Selection of fundamental polar monomers that can be used in copolymerization reactions.

Scheme 1. Coordination Insertion Reaction Overview, Highlighting the Benefits Possible with an Optimized System

• Ota, Y.; Ito, S.; Kobayashi, M.; Kitade, S.; Sakata, K.; has often been overlooked until recently is the improved
Tayano, T.; Nozaki, K. Crystalline Isotactic Polar prospect of controlled polymer degradation following a
Polypropylene from the Palladium-Catalyzed Copoly- product’s useful lifecycle if the polymer contains polar
merization of Propylene and Polar Monomers. Angew. functionalities because of the incorporation of heteroatoms.8,9
Chem., Int. Ed. 2016, 55, 7505−7509.2 The use of a Pd The most commonly practiced method for the introduction
complex bearing a chiral phosphine sulfonate ligand of polar functionalities into homopolymers (such as PE and
allows for the formation of a highly regioregular, PP) is postpolymerization functionalization.10 While appealing
moderately isotactic copolymer with methyl acrylate. as no modifications to existing processes to produce
• Seidel, F. W.; Tomizawa, I.; Nozaki, K. Expedient polyolefins are required, the upgrading of alkanes is notoriously
Synthetic Identification of a Novel P-Stereogenic Ligand difficult because of their inert C−H bonds. Consequently, side
Motif for the Palladium-Catalyzed Preparation of reactions such as cross-linking and chain-scission are hard to
Isotactic Polar Polypropylenes. Angew. Chem., Int. Ed. prevent due to the harsh conditions that these reactions
2020, 10.1002/anie.202009027.3 Our most recent work, require.11 This also leads to a lack of control over the final
using palladium complexes bearing two different classes structure of the functionalized polymer product. A more
of P-stereogenic ligands. In addition to the formation of elegant solution is the direct copolymerization of the olefin
regioregular, isotactic polymers, further computational (for example, ethylene or propylene) with a polar comonomer,
experiments allow for a rational explanation as to why some common examples of which are displayed in Figure 1.
steric bulk is important on the P moiety. These fundamental polar monomers have a functional group
• Konishi, Y.; Tao, W.; Yasuda, H.; Ito, S.; Oishi, Y.; directly attached to the olefinic double bond and are readily
Ohtaki, H.; Tanna, A.; Tayano, T.; Nozaki, K. Nickel- available as they are already used on a large scale within
Catalyzed Propylene/Polar Monomer Copolymeriza- industry.12,13 Copolymerization reactions can proceed either
tion. ACS Macro Lett. 2018, 7, 213−217.4 A chiral Ni/ via a radical or a coordination insertion mechanism. The only
phosphine phenolate complex allows for the copoly- systems currently capable of copolymerization reactions using
merization of propylene with nonfundamental polar polar monomers on a commercial scale are radical based.14
comonomers to give regioregular, moderately isotactic However, these processes require elevated temperatures and
polymers. pressures (of up to 375 °C and 300 MPa, respectively),
resulting in high costs, in addition to producing highly
1. INTRODUCTION branched copolymer products with broad molecular weight
Since its emergence in the early twentieth century, the distributions and low crystallinity.10 In contrast, the single site
development of polymer chemistry has resulted in significant transition metal catalysts used in coordination insertion
advances in both science and engineering. This in turn has polymerization reactions have the potential to offer a level of
directly improved the lifestyles of consumers worldwide, with control that makes them highly desirable from an industrial
applications found in the packaging, automotive, construction, perspective (Scheme 1), as by limiting branch formation,
and medical sectors, among others.5 The two most commonly highly crystalline polymers can be formed.
used polymers in the world today are polyethylene (PE) and As they are well-defined on a molecular level, targeted
polypropylene (PP), as has been the case for almost half a catalyst design and subtle alterations therein could potentially
century.6 While PE and PP serve as excellent materials, give rise to systems in which polymers are prepared with
industrial and academic groups have long sought to further desired proportions of comonomer incorporation, at lower
improve their properties through the controlled incorporation temperatures and pressures than radical polymerizations.
of polar functionalities into the otherwise nonpolar polymer Industrially, the transition metal catalyzed copolymerization
structure. In doing so, the surface properties of the polymer are of ethylene with nonpolar monomers is known; however, no
enhanced, which makes the synthesis of composites containing commercial process currently exists for the copolymerization of
materials such as fibers and talc possible, resulting in improved olefins with polar monomers,14 due to a series of issues termed
and tunable product performance.7 A further advantage that the “polar monomer problem”.15,16 Academically, there have
B https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

Figure 2. Different insertion modes of propylene within a growing polymer chain.

been significant breakthroughs in the copolymerization of Starting from intermediate 1, in which a metal center is π-
ethylene with fundamental polar comonomers (such as those coordinated by a molecule of propylene and σ coordinated by
displayed in Figure 1) that will not be covered here; readers are the polymer chain (the last propylene unit of which has
instead directed to recent literature.17,18 This Account will undergone a 1,2 insertion), the insertion of propylene into the
instead cover the comparatively underdeveloped copolymer- metal−carbon bond in either a 1,2 or a 2,1 mode is possible. In
ization of propylene with polar comonomers. Although this is a scenario A, propylene inserts in a 1,2 fashion to give 2, as does
more challenging reaction than the corresponding one with each subsequent unit of propylene, resulting in a head-to-tail
ethylene, it can elicit polymers with a defined tacticity and polymer with minimal branching. Following each insertion of
regiochemistry, increasing the number of potential final propylene, however, competitive β-hydride elimination and
applications. A recent review from Marks covers early reinsertion in the opposite manner is possible, for example
transition metal catalyzed systems;19 the focus of this Account from 2 to 4. This leads to the formation of a branch along the
will therefore be on late transition metal catalyzed systems,20 polymer chain in a process known as chain walking, as shown
with an emphasis on work from our laboratory. in scenario B. (Here, a single branched unit is shown due to
space constraints, but it is also possible for the branch to
2. PROPYLENE INSERTION AND SELECTIVITY consist of many units, and for branches to form along
To properly frame both the following work and the challenges branches.)
that remain in the area, a short overview of aspects regarding Alternatively, a 2,1 insertion of propylene can occur from 1
the polymerization of propylene is first presented. As to give 3, in which the metal is now bonded to a secondary
compared to ethylene, the extra carbon present within carbon center. Following the coordination and 1,2 insertion of
propylene means that insertion into a metal−carbon bond in propylene, a head to tail defect is found within the chain
either a 1,2 (carbon 1, the terminal carbon, bonds to the metal; (scenario C). Complex 3 is also able to undergo chain walking,
carbon 2, the internal carbon, to the growing polymer chain) which in this case results in chain straightening after a formal
or 2,1 fashion (vice versa) is possible.21 Although 1,2 insertion 1,3 insertion (scenario D). While chain walking is possible
is generally favored in both early and late transition metal after either a 1,2 or 2,1 insertion, it is less likely to take place
systems,22,23 2,1 insertions can occur,24,25 resulting in following a 1,2 insertion (from complex 2) because of the
variations in the regiochemistry of the resultant polymer kinetic and thermodynamic instability of the tertiary alkene
chain (Figure 2). that is formed as a result of β-H elimination from 2. In
C https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

Figure 3. Common tacticities found within polypropylene chains.

Scheme 2. Copolymerization of Propylene with a Polar Comonomers Using Group 4 Metal Precatalyst33

contrast, following a β-H elimination from the 2,1 insertion and allylamine, masked by aluminum when used in the
product (complex 3) a primary alkene is formed, which is a presence of excess amount of alkylaluminum reagents.31,32 A
comparatively more favorable process and thus more readily recent example from Marks does not require the use of excess
facilitates chain walking. Lewis acid for the highly isospecific copolymerization of
Following each insertion of propylene in either a 1,2 or 2,1 propylene with an amino olefin bearing a bulky substituent, but
mode, a new chiral center is formed in the growing polymer several methylene spacers were required between the amino
chain. The stereoregularity of the resulting polymer can be and olefin groups, limiting the practicality of the system
calculated using quantitative 13C NMR spectroscopy, as the (Scheme 2).33
chemical shift of each methyl carbon in the chain is highly Our laboratory became interested in the use of less oxophilic
sensitive to the stereochemistry of its neighboring carbon metal catalysts to avoid the limitations associated with group 4
atoms (up to 5 monomer units either side of the carbon α to metal chemistry as described above. The iron bis(imino)-
itself). It is, therefore, possible to quantify the degree of
pyridine (PDI) complex, 5, in combination with MMAO,
isotacticity present within the polymer as the probability of
originally reported by Small and Brookhart to give isotactic
finding consecutive chiral centers with the same (meso, m) or
opposite (racemo, r) absolute configuration. In highly polypropylene with excellent tacticity control,22 was used for
stereospecific polymers, this can be done to the pentad level the copolymerization of propylene with allyl monomers.
(the chance of finding five consecutive m stereogeneic centers, Amorphous polymers with comonomer incorporation of
mmmm). For less stereoregular polymerizations, the shorter 0.3−2.1% were formed (Scheme 3).34 The polymerization
triad or diad counts (mm and m, respectively) can be used. activity was significantly lowered by the presence of polar
This allows for the identification of several different polymer monomers, however, and the applicable polar monomers were
tacticities, some of which are displayed in Figure 3. limited to silylated alcohols.
From the above regio- and stereoinsertion modes, there is a To date, the group 10 metal catalysts reported for propylene
strong industrial preference for high molecular weight polymerization are inferior to their early transition metal
crystalline polymers, meaning that highly active catalysts
capable of giving the isotactic (and less commonly, Scheme 3. Copolymerization of Propylene and Allyl
syndiotactic) form with minimal branching are sought.26 Monomers Using an Iron(PDI) Complex
Early transition metal catalysts (typically d0 metals from
groups 3 and 4) can do this efficiently and have been widely
exploited for the homopolymerization of propylene.27 In
contrast, late transition metal catalysts (typically square planar
d8 nickel or palladium complexes) readily undergo chain
walking and straightening processes unless the reaction
temperature is extremely low (vide infra),28 which have so
far limited their use for the polymerization of propylene.29

3. COPOLYMERIZATION OF PROPYLENE WITH


POLAR COMONOMERS
As early transition metal d0 catalysts (which are highly active in
homopolymerization reactions) possess a low-lying LUMO,30
they are highly oxophilic and, therefore, easily poisoned by
polar functionalities. The comonomers that have been
demonstrated to be compatible with early transition metal
catalysts are protected polar monomers, such as allyl alcohol
D https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

Figure 4. Metal/ligand combinations that have been successfully employed for the copolymerization of propylene with polar comonomers. M =
Pd/Ni; R, R′ = aryl/alkyl/silyl. BArF4 = [B(3,5-(CF3)2C6H3)4]−. Although the Ni/PS example has not been used with propylene, it is worth noting
due to its high activity in copolymerizations with ethylene.35

Table 1. Copolymerization of Propylene with Allyl and Acetate Comonomers Using Either a Pd/IzQO or Pd/PS Precatalyst

entry catalyst (μmol) X (mmol) yield (g) activity (mol/mol h) Mn (×10−3) Mw/Mn incorp. (%) [mm] regiodefects (mol %)
a ,b
1 6 (2) 0.20 808 16 2.2 1.2
2a,c 7 (2) 2.38 235 24 2 0.57 0.9
3b 6 (2) CH2OAc (0.1) 0.08 30.9 3.9 2.2 2.0 1.6
4c 7 (2) CH2OAc (0.46) 0.18 17.8 8.9 2.7 1.7 0.55 1.0
5b,d 6 (10) CO2Me (0.11) 0.04 16.4 3.0 2.2 1.5 - 0.9
6c 7 (10) CO2Me (2.77) 0.20 19.7 4.6 2.2 1.6 0.56 0.9
a
No comonomer added. b0.23 mol of propylene, 20 mL of toluene, 100 °C, 6 h. c0.14 mol of propylene, 10 mL of toluene, 50 °C, 12 h. d3 h.

counterparts regarding tacticity control and activity. They nickel are known as Brookhart catalysts) have gone on to be
have, however, shown the requisite tolerance toward polar widely exploited across late transition metal catalyzed
functionalities to enable the successful copolymerization of polymerizations; however, there are no further reports using
ethylene with fundamental polar comonomers, to yield polar diimine ligands for the copolymerization of propylene. When
functionalized copolymers. Taken together, this suggests that taken together with more recent work from Brookhart on the
with the correct catalyst design, the selective copolymerization copolymerization of ethylene with vinylsilanes to give high
of propylene with polar comonomers using late transition molecular weight copolymers with minimal branching,37 this
metal catalysts should be possible. The remainder of this suggests that the selective insertion of propylene remains
Account covers our efforts toward developing and investigating challenging with complexes bearing these ligands because of
these reactions and how the challenges that remain in the area their predisposition toward chain walking (β-hydride elimi-
may be confronted. nation followed by reinsertion).28 Indeed, the formation of
The copolymerization of propylene with polar comonomers highly isotactic polypropylene using a Brookhart catalyst was
using late transition metal catalysts has so far been reported shown to be possible by Coates,38 yet required a reaction
using the ligand classes displayed in Figure 4. temperature of −78 °C, rendering polar commoner incorpo-
The first report in the area came from Brookhart, who used ration unfeasible.
cationic palladium complexes bearing α-diimine ligands to In contrast, palladium complexes bearing an asymmetric
successfully incorporate either MA or fluorinated octyl acrylate anionic bidentate ligand have been shown to effectively
(FOA) into the resultant polymer.36 The amorphous polymers suppress β-hydride elimination from alkylpalladium species.
formed were highly branched and contained chain straightened For example, palladium complexes bearing a phosphine−
units due to the formal 1,3 insertion of propylene (scenarios B sulfonate (PS) ligand (originally reported by Drent in 2002)39
and D in Figure 2). MA was found to insert in a 2,1 mode and have been used as catalysts to afford highly linear polyethylene
was mainly located at the branch ends, indicating that its and to copolymerize ethylene with polar comonomers.40
insertion was followed by chain termination or chain walking. Accordingly, we thought that if by using these palladium
Neutral α-diimine ligands (which complexed with palladium or catalysts the insertion mode of propylene could be controlled
E https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

in a 1,2 fashion, it would be possible to obtain regiocontrolled BPMO system (8, [mm] = up to 0.75), while the Pd/PS
polypropylene since scenario B in Figure 2 would be avoided. system was also moderately stereoselective (9, [mm] = up to
While examining a variety of unsymmetric bidentate anionic 0.67). Both systems were also able to copolymerize propylene
ligands, we found that palladium complexes bearing an with allyl comonomers while maintaining their stereoselectivity
imidazolium-quinolinolate (IzQO) ligand (complex 6, Table (Scheme 5). It should not be overlooked that 8 and 9 are
1) were able to generate regiocontrolled polypropylene, the phosphinite complexes. When an analog of 9 with ArO = 2,6-
first example of a group 10 metal catalyst to do so. By Ph2C6H3O is compared with its phosphine counterpart (with
employing a bulky substituent on the nitrogen atom, a higher
2,6-Ph2C6H3CH2), the tacticity control was about the same
molecular weight polymer with less regio defects was formed,
though the polymer was completely atactic.1 Recently, the use while the phosphine was five times more active than the
of a Pd/aIzPy precatalyst was reported by Hong to be active phosphinite.3 It is of interest that the copolymerization
for the regioselective polymerization of propylene, though no reactions using the Pd/BPMO precatalyst 8 in fact displayed
details regarding stereoselectivity were provided.41 a greater stereoselectivity than the homopolymerized counter-
In parallel, we examined the effect of altering the bulky part. This “comonomer enhanced stereocontrol” was also
substituents of the phosphine−sulfonate (PS) ligand in their noted recently by Marks in Zr-catalyzed copolymerization
corresponding palladium complexes. With the introduction of reactions,33 and further experiments are ongoing in our
two menthyl groups on the phosphorus atom,2 (complex 7, laboratory to establish the rationale behind this surprising
Table 1), regiocontrolled polypropylene was obtained with a effect.
moderate isotacticity of [mm] = 0.57. The effective protection While good progress has been made in the palladium-
of one of the axial sites at the palladium center by the ligand catalyzed copolymerization of olefins with polar monomers,
was suggested to be the reason for the observed regiocontrol, the corresponding reaction using nickel catalysts is more
in addition to the suppression of chain transfer.20 As a attractive because of the cheap and abundant nature of nickel
representative example, the number of polymer chains per compared to palladium. Thus, intensive efforts have been
catalyst was 10 and the incorporation of polar monomer per
devoted toward this end.42,43 It should be noted, however, that
chain was 3.5 units (entry 4, Table 1). With these two classes
of catalysts, the copolymerization of propylene with several until 2018 nickel catalysts were hardly applicable to propylene
polar monomers, including methyl acrylate (MA), a polymerization due to the low regioselectivity observed. In
fundamental polar monomer, became possible as shown in 2017, Shimizu and co-workers reported a Ni/SHOP type
Table 1. catalyst for the copolymerization of ethylene and polar
As with the homopolymerization, no stereoselectivity was monomers;44 Pan and co-workers have since expanded the
observed using 6, though 7 was able to produce copolymers scope of polar comonomers tolerated by the reaction.45
with a value of [mm] of up to 0.56. A 1,2 insertion of both Unfortunately, the SHOP catalysts used in their research,
propylene and allyl acetate was observed using each bearing a triarylphosphine−phenolate ligand, afforded only
precatalyst; however, a different insertion mode of MA was short oligomers when propylene was employed instead of
observed in each case: 1,2 insertion for 6 and 2,1 insertion for ethylene. We speculated that the low molecular weight
7 (Scheme 4). Notably, the 2,1 MA insertion when 7 was used originated from rapid chain-transfer following a 2,1 insertion
results in a head-to-head structure at the junction of propylene of propylene with the Ni/SHOP catalyst. We envisioned that
and MA. the bulky alkyl substituent on the P atom within our PS ligand
may be applicable to the SHOP system and could, therefore,
Scheme 4. Differing Regioinsertion Modes of MA Using Pd/
allow for a further improvement regarding the 1,2 regio-control
IzQO and Pd/PS Precatalysts
of propylene insertion. Accordingly, with the well-defined
nickel precatalyst 10, which bears an alkylphosphine-phenolate
(PO) ligand, crystalline copolymers with minimal 1,3 insertion
defects were formed (the less bulky derivative 11 did not
furnish an active system). Comonomer incorporation (up to
1.4%) into the main polypropylene chain was possible, and the
reaction was carried out at a practical temperature of 50 °C
(Table 2).4 A nickel complex bearing a bisphosphine
monoxide (BPMO) ligand in combination with MMAO was
also applicable for the copolymerization of propylene with allyl
This differing regioselectivity is interesting, though it is not acetate, although the copolymerization suffered from poor
immediately apparent why it is observed. This subject will be activity and low comonomer incorporation.46
returned to in the final section of this Account along with a Despite affording a highly regioselective system, the
discussion on the factors that dictate which regioinsertion
precatalyst 10 was not able to impart any stereoregularity
mode is formed (vide infra).
Very recently, our group has developed two classes of within the resulting polymer (values of [mm] below 0.20 were
asymmetric bidentate ligands, each containing a stereogenic obtained). To improve the stereocontrol of the system, a
phosphorus atom.3 Using a rapid synthetic protocol, a large precatalyst bearing a ligand with chiral menthyl groups on its
family of Pd/BPMO and Pd/PS catalysts were screened. The phosphine moiety (12, Table 2) was synthesized. This allowed
highest isotacticity value among late transition metal catalysts for the production of moderately isotactic polymers and
(disregarding the examples38 which require a very low copolymers ([mm] values of up to 0.61 obtained) due to
temperature, for example, −78 °C) was found using the Pd/ increased enantiomorphic site control.
F https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

Scheme 5. Homopolymerization of Propylene and the Copolymerization of Propylene with Allyl Monomers Using Pd/PS and
Pd/BPMO Precatalystsa

a
BArF4 = [B(3,5-(CF3)2C6H3)4]−. L = 2,6 lutidine or pyridine.

Table 2. Copolymerization of Propylene and Allyl Monomers Using Nickel/Alkylphosphine−phenolate Complexesa

entry Ni (μmol) R (mmol) activity (mol/mol h) Mn (×10−3) Mw/Mn incorp. (mol %) [mm] regiodefects (mol %)
b,c
1 10 (30) 2614 17 1.8 0.19 0.61
2b 11 (30)
3b 12 197 16 1.8 0.60 7.70
4 10 (20) CH2OH (6.97) 7.12 5.2 1.9 1.4 0.18 0.36
5 10 (10) NHCO2(tBu) (0.76) 11.9 11 1.9 0.39 0.19 0.39
6 10 (20) CH2OAc (1.59) 5.46 7.1 1.3 0.91 0.19 0.80
7 12 (20) CH2OAc (0.81) 42.8 9.1 1.5 0.49 0.61 9.20
a
Conditions: Propylene (237 mmol) toluene (10 mL), 50 °C, reaction time = 12 h. bNo comonomer added, reaction time = 3 h. cThe use of 7
instead of 6 for the homopolymerization of propylene resulted in no polymer product being formed.

4. REGIOINSERTION MODES OF PROPYLENE AND can give polymers with different structures and properties. In
FUNDAMENTAL POLAR COMONOMERS the case of propylene 1,2 insertion dominates, but the degree
The final section of this Account will discuss the factors that of regiocontrol depends on the ligand used for the polymer-
affect the insertion of propylene and polar monomers in ization. Moreover, the choice of ligand can alter the insertion
copolymerization reactions between the two. Whether a 1,2 or mode of polar monomers (for example, see the different
2,1 insertion takes place is an important consideration in insertion modes of MA observed using the Pd/PS and Pd/
copolymerization reactions with propylene (and any other IzQO systems displayed in Table 1). Our recent work using P-
nonsymmetric alkene) as the insertion mode of the monomer stereogenic Pd/BPMO and Pd/PS catalysts (8 and 9, Scheme
G https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

5, and shown again in Figure 5 for clarity) has allowed us to dimethoxyphenyl group aligned parallel to the Pd square
more clearly define the reasons why, based on their steric planar coordination plane; Figure 6) and their corresponding
profiles, these ligands are able to deliver stereo- and SamVca 2.1 plots (not shown here). In each case, one face of
regioselective polymers.3 the coordination plane is effectively shielded by one of the 2,6-
dimethoxyphenyl groups, rendering it more sterically con-
gested than the other. The growing polymer chain is proposed
to occupy the less sterically congested face (the purple area on
Figure 6), while forcing the methyl group on the propylene to
face the 2,6-dimethoxyphenyl moiety (the green orientation on
Figure 6). Other arrangements of the propylene (as shown by
the red orientations) are less favorable, which we tentatively
suggest to be due to reduced London dispersion interactions.
Figure 5. Our recently reported Pd/BPMO and Pd/PS precatalysts, Together therefore, the steric bulk on the P atom ensures chain
used as computational models to help understand why they furnish propagation occurs both regio- and stereoselectively.
active systems. It has previously been noted for stereoselective acrylate
insertion using Pd/PS complexes that there exists a “window”
Both classes of catalyst were found to be under within the steric spectrum that allows for regio- and
enantiomorphic site control, meaning that the coordination stereoselectivity to be achieved, and that excessive bulk near
mode of propylene prior to insertion dictates the resulting to the Pd center is detrimental.50 This is pertinent in the case
tacticity of the polymer. As the propagating polymer chain of 8 and 9, where (as shown in Figure 6 and discussed above)
must be trans to the P moiety to be sufficiently activated to there is a delicate balance between the coordination modes of
allow for further propagation,47 the coordinated propylene propylene within the catalytic pocket. The presence of the
must be cis to the P atom in the square planar geometry. Thus, phosphinite oxygen atom in 8 and 9 may allow the ligand to be
the methyl group of the propylene is sterically repelled by the positioned so as to provide moderate steric protection close to
bulk of the two R groups on the P center, enforcing a the Pd center, while still being able to affect the regio- and
regioselective 1,2 insertion. It can also be said that these stereoinsertion of propylene.
ligands display a certain electronic preference for 1,2 insertion, Regarding the polar comonomers used in these reactions, it
as the use of less bulky derivatives did not result in a large has been demonstrated computationally that for palladium-
increase in the proportion of regiodefects (excluding an outlier, alkyl complexes a 2,1 insertion is favored on electronic grounds
below 5% was recorded in all cases). when electron-deficient monomers (such as acrylates) are
The stereoselectivity can be explained through analogy to used, whereas for electron-rich monomers (such as vinyl
group IV metallocene catalysts, which are proposed to enforce ethers) a 1,2 insertion is preferred.51,52 A recent report by
the methyl group on the propylene to align anti with respect to Gordon, Copéret, and co-workers demonstrated that for group
the growing polymer chain.48,49 Furthermore, these studies 10 metal catalysts the insertion of an olefin into a metal alkyl
suggest that the relative orientation of the propylene monomer bond involves the nucleophilic attack of an occupied
is less dictated by the steric environment provided by the antibonding orbital of π-symmetry of the metal alkyl onto
ligand itself but rather by the growing polymer chain (whose the π*-orbital of the bound olefin, which eloquently describes
orientation is enforced by the sterics of the ligand). We the reason for this differing selectivity.30 In practice, both
suggested the same anti orientation to be present in our PS and insertion modes often take place, and it would be remiss to
BPMO systems, based on the crystal structures of the three apply a one size fits all approach to any system. Stoichiometric
best performing precatalysts (each of which contained a 2,6- reactions between well-defined complexes and the polar

Figure 6. ORTEP structure of Pd/PS precatalyst 9, viewed along two different axes. The purple shading represents the least sterically encumbered
area around the palladium center according to the SamVca 2.1 plot (to be occupied by the growing polymer chain). The green orientation of
propylene is proposed to be more favorable than the red due to less steric pressure from the ligand and growing polymer chain.3

H https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

Figure 7. Different insertion modes of (i) methyl acrylate (MA; Brookhart) and methyl methacrylate (MMA; Sen) into the same cationic
palladium α-diimine complex and (ii) MA into cationic palladium BPMO complexes, with either alkyl or aryl substituents (Nozaki).

monomer function as a useful way of elucidating which is


favored ex situ.
The potentially capricious behavior of polar monomers is
highlighted in Figure 7, where (i) using the same Pd α-diimine
complex, a different insertion mode is observed depending on
whether methyl acrylate (MA) or methyl methacrylate (MMA)
is used36,53 and (ii) altering the ligand shrubbery of Pd
bisphosphine monoxide (BPMO) complexes is shown to give a
different insertion mode of MA.54
While for Figure 7i, it is clear that the bulkier monomer
MMA is prevented from undergoing a 2,1 insertion because of
the methyl substituent, it is less obvious why the differing
regioselectivity is observed in the case of Figure 7ii due to the
similar steric profiles of each ligand. Mecking has more clearly
demonstrated that for palladium complexes bearing diazaphos- Figure 8. Different insertion modes of methyl acrylate (MA) into
pholidine-sulfonate ligands, increasing the bulk of the ligand palladium diazaphospholidine-sulfonate complexes. L = 2,6 lutidine.
leads to an alteration of the regioselectivity of MA insertion
from the preferred 2,1 to a 1,2 mode (Figure 8).55
Although the complexes do not function as active
precatalysts, Figure 8 provides strong evidence that an increase
in the steric bulk of the ligand results in a change in the
regioselectivity of acrylate insertion from 2,1 to 1,2. It is
important to note that either a 1,2 or a 2,1 insertion of polar
monomers can furnish an active catalytic system, as
demonstrated by Brookhart who used palladium and nickel
α-diimine catalysts for the copolymerization of ethylene with
vinylalkoxysilane (Figure 9).37 In each case, active systems
were reported following a predominant 2,1 (nickel) or 1,2
(palladium) insertion of the silane into the metal alkyl bond.
We can now return to the differing regioselectivity observed Figure 9. Work from Brookhart demonstrating a 1,2 or 2,1 insertion
with respect to MA insertion when the precatalysts 6 and 7 are of a vinylsilane into a metal−carbon bond. L = Et2O.
used in copolymerization reactions. As discussed above, the
regioselectivity of MA insertion into palladium-alkyl bonds can fashion. It is tempting to suggest that the difference in
be altered depending on the ligand used, with bulky ligands regioselectivity demonstrated by 6 and 7 is due to an increased
leading to a preferential insertion of polar monomers in a 1,2 steric bulk of the IzQO ligand vis-à-vis the PS. While it is
I https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

correct that the IzQO ligand present in 6 is indeed bulky, we Funding


have also shown that menthyl-substituted PS ligands, such as This work was supported by JST CREST Grant Number
that present in 7, possess significant steric bulk,56 suggesting JPMJCR1323, Japan. A part of this work was conducted at the
that the observed regioselectivity is more nuanced than ligand Research Hub for Advanced Nano Characterization, The
steric arguments alone (a 1,2 insertion would otherwise be University of Tokyo, supported by MEXT, Japan.
expected using 7). It would therefore appear that the electronic Notes
composition of the PS ligand dictates the preferential 2,1
The authors declare no competing financial interest.
insertion of MA when 7 is used as a precatalyst in the
copolymerization reaction with propylene (Table 1).40 This Biographies
highlights the delicate balance that exists between sterics and
Stephen Luckham was born in Manchester, England, in 1991. He
electronics when the ligand and monomer within a system are received his M.Chem. degree from the University of Leicester in 2014,
altered, and serves as a reminder that we do not yet possess followed by an M.Res. degree (Catalysis) from Cardiff University in
sufficient understanding to confidently predict which regioin- 2015. He then moved to the University of Bristol where he undertook
sertion product will form without a thorough mechanistic his doctoral studies under the guidance of Prof. Robin Bedford,
investigation. followed by a postdoctoral position in collaboration with Victrex plc.
In 2019, he joined the group of Prof. Kyoko Nozaki at the University
5. CONCLUSION AND PERSPECTIVE of Tokyo as a postdoctoral fellow.
As this Account has demonstrated, the copolymerization of Kyoko Nozaki was born in Osaka, Japan, and graduated from Kyoto
propylene with polar monomers has been accomplished by University with a B.Sc. degree in 1986. She received her Ph.D. in 1991
palladium and nickel complexes of finely tuned asymmetric from the same university under the guidance of Prof. Kiitiro Utimoto.
bidentate ligands. Further catalyst development should be During her Ph.D. studies, she joined Prof. Clayton H. Heathcock’s
targeted toward improving the activity and tacticity control group at the University of California, Berkeley, as an exchange student
offered by these systems, as they remain far behind group 4 for 1 year in 1988. Since 1991, she has been a faculty member at
metal catalysts in these important areas. Considering its price, Kyoto University, moved to the University of Tokyo in 2002, and has
nickel should be the metal of choice for industrial applications, been a Professor at the University of Tokyo since 2003. Her research
while further studies on palladium complexes can be used to interests are focused on the development of homogeneous catalysts
for polymer synthesis and organic synthesis.


further develop our fundamental understanding of these
processes. For industrial applications, the studies on materials
properties (viscosity, rheology, surface, mechanical, etc.) are ACKNOWLEDGMENTS
also critical. The authors are grateful to Prof. Shingo Ito (Nanyang
Although much information was provided by DFT Technological University, Singapore), Dr. Yusuke Ota (ETH,
calculations when we studied the effect of the ligand on the Zürich), and Mr. Falk W. Seidel (The University of Tokyo) for
suppression of β-hydride elimination,45 the design of catalysts careful proof reading and valuable suggestions.
capable of inducing high isotacticity in propylene polymer-
ization is less predictable. There are large number of close
lying, energetically accessible isomeric structures that are
■ REFERENCES
(1) Nakano, R.; Nozaki, K. Copolymerization of Propylene and
responsible for the observed polymer tacticities. In this regard, Polar Monomers Using Pd/IzQO Catalysts. J. Am. Chem. Soc. 2015,
machine learning or a combinatorial approach may better fit 137, 10934−10937.
(2) Ota, Y.; Ito, S.; Kobayashi, M.; Kitade, S.; Sakata, K.; Tayano, T.;
for further catalyst development.


Nozaki, K. Crystalline Isotactic Polar Polypropylene from the
Palladium-Catalyzed Copolymerization of Propylene and Polar
AUTHOR INFORMATION Monomers. Angew. Chem., Int. Ed. 2016, 55, 7505−7509.
(3) Seidel, F. W.; Tomizawa, I.; Nozaki, K. Expedient Synthetic
Corresponding Author Identification of a Novel P-Stereogenic Ligand Motif for the
Kyoko Nozaki − Department of Chemistry and Biotechnology, Palladium-Catalyzed Preparation of Isotactic Polar Polypropylenes.
School of Engineering, The University of Tokyo, 113-8656 Angew. Chem., Int. Ed. 2020, DOI: 10.1002/anie.202009027.
Tokyo, Japan; orcid.org/0000-0002-0321-5299; (4) Konishi, Y.; Tao, W.; Yasuda, H.; Ito, S.; Oishi, Y.; Ohtaki, H.;
Tanna, A.; Tayano, T.; Nozaki, K. Nickel-Catalyzed Propylene/Polar
Email: nozaki@chembio.t.u-tokyo.ac.jp
Monomer Copolymerization. ACS Macro Lett. 2018, 7, 213−217.
Author (5) Millet, H.; Vangheluwe, P.; Block, C.; Sevenster, A.; Garcia, L.;
Antonopoulos, R. The Nature of Plastics and Their Societal Usage. In
Stephen L. J. Luckham − Department of Chemistry and Plastics and the Environment; The Royal Society of Chemistry, 2019;
Biotechnology, School of Engineering, The University of pp 1−20.
Tokyo, 113-8656 Tokyo, Japan; orcid.org/0000-0003- (6) Stürzel, M.; Mihan, S.; Mülhaupt, R. From Multisite Polymer-
3732-0668 ization Catalysis to Sustainable Materials and All-Polyolefin
Composites. Chem. Rev. 2016, 116, 1398−1433.
Complete contact information is available at: (7) Goodall, B. L. Late Transition Metal Catalysts for the
https://pubs.acs.org/10.1021/acs.accounts.0c00628 Copolymerization of Olefins and Polar Monomers. In Metal Catalysts
in Olefin Polymerization; Guan, Z., Ed.; Springer: Berlin, 2009; pp
Author Contributions 159−178.
(8) Wang, X.; Seidel, F. W.; Nozaki, K. Synthesis of Polyethylene
The manuscript was written through contributions of all with In-Chain α,β-Unsaturated Ketone and Isolated Ketone Units:
authors. All authors have given approval to the final version of Pd-Catalyzed Ring-Opening Copolymerization of Cyclopropenone
the manuscript. with Ethylene. Angew. Chem., Int. Ed. 2019, 58, 12955−12959.

J https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

(9) Morgen, T. O.; Baur, M.; Göttker-Schnetmann, I.; Mecking, S. (30) Bumberger, A. E.; Gordon, C. P.; Trummer, D.; Copéret, C.
Photodegradable Branched Polyethylenes from Carbon Monoxide C−H Activation and Olefin Insertion in D8 and D0 Complexes: Same
Copolymerization under Benign Conditions. Nat. Commun. 2020, 11, Elementary Steps, Different Electronics. Helv. Chim. Acta 2020, 103,
3693. e1900278.
(10) Franssen, N. M. G.; Reek, J. N. H.; de Bruin, B. Synthesis of (31) Imuta, J.; Kashiwa, N.; Toda, Y. Catalytic Regioselective
Functional ‘Polyolefins’: State of the Art and Remaining Challenges. Introduction of Allyl Alcohol into the Nonpolar Polyolefins:
Chem. Soc. Rev. 2013, 42, 5809−5832. Development of One-Pot Synthesis of Hydroxyl-Capped Polyolefins
(11) Boaen, N. K.; Hillmyer, M. A. Post-Polymerization Function- Mediated by a New Metallocene IF Catalyst. J. Am. Chem. Soc. 2002,
alization of Polyolefins. Chem. Soc. Rev. 2005, 34, 267−275. 124, 1176−1177.
(12) Nakamura, A.; Ito, S.; Nozaki, K. Coordination−Insertion (32) Hagihara, H.; Tsuchihara, K.; Sugiyama, J.; Takeuchi, K.;
Copolymerization of Fundamental Polar Monomers. Chem. Rev. Shiono, T. Copolymerization of Propylene and Polar Allyl Monomer
2009, 109, 5215−5244. with Zirconocene/Methylaluminoxane Catalyst: Catalytic Synthesis of
(13) Monomers in which there are one or more methylene spacer Amino-Terminated Isotactic Polypropylene. Macromolecules 2004, 37,
groups between the olefin and the functional group are also used in 5145−5148.
copolymerization reactions and can possess useful material properties (33) Huang, M.; Chen, J.; Wang, B.; Huang, W.; Chen, H.; Gao, Y.;
in the polymer product but are less industrially relevant. Marks, T. J. Polar Isotactic and Syndiotactic Polypropylenes by
(14) Jeremic, D. Polyethylene. In Ullmann’s Encyclopedia of Industrial Organozirconium-Catalyzed Masking-Reagent-Free Propylene and
Chemistry; American Cancer Society, 2014; pp 1−42. Amino−Olefin Copolymerization. Angew. Chemie Int. Ed. 2020, 59
(15) Jordan, R. F. Progress on the Polar Monomer Problem. Polym. (46), 20522−20528.
Prepr. 2007, 48, 196−197. (34) Kawakami, T.; Ito, S.; Nozaki, K. Iron-Catalysed Homo- and
(16) Keyes, A.; Basbug Alhan, H. E.; Ordonez, E.; Ha, U.; Beezer, D. Copolymerisation of Propylene: Steric Influence of Bis(Imino)-
B.; Dau, H.; Liu, Y.-S.; Tsogtgerel, E.; Jones, G. R.; Harth, E. Olefins Pyridine Ligands. Dalt. Trans. 2015, 44, 20745−20752.
and Vinyl Polar Monomers: Bridging the Gap for Next Generation (35) Chen, M.; Chen, C. Rational Design of High-Performance
Materials. Angew. Chem., Int. Ed. 2019, 58, 12370−12391. Phosphine Sulfonate Nickel Catalysts for Ethylene Polymerization
(17) Dall’Anese, A.; Rosar, V.; Cusin, L.; Montini, T.; Balducci, G.; and Copolymerization with Polar Monomers. ACS Catal. 2017, 7,
D’Auria, I.; Pellecchia, C.; Fornasiero, P.; Felluga, F.; Milani, B. 1308−1312.
Palladium-Catalyzed Ethylene/Methyl Acrylate Copolymerization: (36) Johnson, L. K.; Mecking, S.; Brookhart, M. Copolymerization
Moving from the Acenaphthene to the Phenanthrene Skeleton of α- of Ethylene and Propylene with Functionalized Vinyl Monomers by
Diimine Ligands. Organometallics 2019, 38, 3498−3511. Palladium(II) Catalysts. J. Am. Chem. Soc. 1996, 118, 267−268.
(18) Tan, C.; Chen, C. Emerging Palladium and Nickel Catalysts for (37) Chen, Z.; Brookhart, M. Exploring Ethylene/Polar Vinyl
Copolymerization of Olefins with Polar Monomers. Angew. Chem., Int. Monomer Copolymerizations Using Ni and Pd α-Diimine Catalysts.
Ed. 2019, 58, 7192−7200. Acc. Chem. Res. 2018, 51, 1831−1839.
(19) Marks, T. J.; Chen, J.; Gao, Y. Early Transition Metal Catalysis (38) Cherian, A. E.; Rose, J. M.; Lobkovsky, E. B.; Coates, G. W. A
for Olefin-Polar Monomer Copolymerization. Angew. Chem., Int. Ed. C2-Symmetric, Living α-Diimine Ni(II) Catalyst: Regioblock
2020, 59, 14726−14735. Copolymers from Propylene. J. Am. Chem. Soc. 2005, 127, 13770−
(20) Ito, S. Palladium-Catalyzed Homo- and Copolymerization of 13771.
Polar Monomers: Synthesis of Aliphatic and Aromatic Polymers. Bull. (39) Drent, E.; van Dijk, R.; van Ginkel, R.; van Oort, B.; Pugh, R. I.
Chem. Soc. Jpn. 2018, 91, 251−261. Palladium Catalysed Copolymerisation of Ethene with Alkylacrylates:
(21) The inductive effect of the methyl group as compared to a Polar Comonomer Built into the Linear Polymer Chain. Chem.
hydrogen atom also alters the energies of the HOMO and LUMO Commun. 2002, 744−745.
orbitals. (40) Nakamura, A.; Anselment, T. M. J.; Claverie, J.; Goodall, B.;
(22) Kawamura-Kuribayashi, H.; Koga, N.; Morokuma, K. An Ab Jordan, R. F.; Mecking, S.; Rieger, B.; Sen, A.; van Leeuwen, P. W. N.
Initio MO Study on Ethylene and Propylene Insertion into the M.; Nozaki, K. Ortho-Phosphinobenzenesulfonate: A Superb Ligand
Titanium-Methyl Bond in CH3TiCl2+ as a Model of Homogeneous for Palladium-Catalyzed Coordination−Insertion Copolymerization
Olefin Polymerization. J. Am. Chem. Soc. 1992, 114, 2359−2366. of Polar Vinyl Monomers. Acc. Chem. Res. 2013, 46, 1438−1449.
(23) Michalak, A.; Ziegler, T. Stochastic Simulations of Polymer (41) Park, D.-A.; Byun, S.; Ryu, J. Y.; Lee, J.; Lee, J.; Hong, S.
Growth and Isomerization in the Polymerization of Propylene Abnormal N-Heterocyclic Carbene−Palladium Complexes for the
Catalyzed by Pd-Based Diimine Catalysts. J. Am. Chem. Soc. 2002, Copolymerization of Ethylene and Polar Monomers. ACS Catal.
124, 7519−7528. 2020, 10, 5443−5453.
(24) Small, B. L.; Brookhart, M. Polymerization of Propylene by a (42) Mu, H.; Pan, L.; Song, D.; Li, Y. Neutral Nickel Catalysts for
New Generation of Iron Catalysts: Mechanisms of Chain Initiation, Olefin Homo- and Copolymerization: Relationships between Catalyst
Propagation, and Termination. Macromolecules 1999, 32, 2120−2130. Structures and Catalytic Properties. Chem. Rev. 2015, 115, 12091−
(25) Hustad, P. D.; Tian, J.; Coates, G. W. Mechanism of Propylene 12137.
Insertion Using Bis(Phenoxyimine)-Based Titanium Catalysts: An (43) Mecking, S.; Schnitte, M. Neutral Nickel(II) Catalysts: From
Unusual Secondary Insertion of Propylene in a Group IV Catalyst Hyperbranched Oligomers to Nanocrystal-Based Materials. Acc.
System. J. Am. Chem. Soc. 2002, 124, 3614−3621. Chem. Res. 2020, 53, 2738.
(26) Because of their faster rates of crystallization during injection (44) Xin, B. S.; Sato, N.; Tanna, A.; Oishi, Y.; Konishi, Y.; Shimizu,
molding, the isotactic form is preferred in industry. F. Nickel Catalyzed Copolymerization of Ethylene and Alkyl
(27) Resconi, L.; Cavallo, L.; Fait, A.; Piemontesi, F. Selectivity in Acrylates. J. Am. Chem. Soc. 2017, 139, 3611−3614.
Propene Polymerization with Metallocene Catalysts. Chem. Rev. 2000, (45) Zhang, Y.; Mu, H.; Pan, L.; Wang, X.; Li, Y. Robust Bulky
100, 1253−1346. [P,O] Neutral Nickel Catalysts for Copolymerization of Ethylene with
(28) Guo, L.; Dai, S.; Sui, X.; Chen, C. Palladium and Nickel Polar Vinyl Monomers. ACS Catal. 2018, 8, 5963−5976.
Catalyzed Chain Walking Olefin Polymerization and Copolymeriza- (46) Jung, J.; Yasuda, H.; Nozaki, K. Copolymerization of Nonpolar
tion. ACS Catal. 2016, 6, 428−441. Olefins and Allyl Acetate Using Nickel Catalysts Bearing a Methylene-
(29) It has recently been demonstrated that the insertion of an Bridged Bisphosphine Monoxide Ligand. Macromolecules 2020, 53,
alkene into a metal carbon bond, as takes place during chain 2547−2556.
propagation within polymerization reactions, must be considered to (47) Nakano, R.; Chung, L. W.; Watanabe, Y.; Okuno, Y.; Okumura,
occur via fundamentally different mechanisms for d0 and d8 systems. Y.; Ito, S.; Morokuma, K.; Nozaki, K. Elucidating the Key Role of
(see ref 30). Phosphine−Sulfonate Ligands in Palladium-Catalyzed Ethylene

K https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research pubs.acs.org/accounts Article

Polymerization: Effect of Ligand Structure on the Molecular Weight


and Linearity of Polyethylene. ACS Catal. 2016, 6, 6101−6113.
(48) Cavallo, L.; Guerra, G.; Vacatello, M.; Corradini, P. A Possible
Model for the Stereospecificity in the Syndiospecific Polymerization
of Propene with Group 4a Metallocenes. Macromolecules 1991, 24,
1784−1790.
(49) Dahlmann, M.; Erker, G.; Nissinen, M.; Fröhlich, R. Direct
Experimental Observation of the Stereochemistry of the First Propene
Insertion Step at an Active Homogeneous Single-Component
Metallocene Ziegler Catalyst. J. Am. Chem. Soc. 1999, 121, 2820−
2828.
(50) Neuwald, B.; Caporaso, L.; Cavallo, L.; Mecking, S. Concepts
for Stereoselective Acrylate Insertion. J. Am. Chem. Soc. 2013, 135,
1026−1036.
(51) Szabo, M. J.; Jordan, R. F.; Michalak, A.; Piers, W. E.; Weiss, T.;
Yang, S.-Y.; Ziegler, T. Polar Copolymerization by a Palladium−
Diimine-Based Catalyst. Influence of the Catalyst Charge and Polar
Substituent on Catalyst Poisoning and Polymerization Activity. A
Density Functional Theory Study. Organometallics 2004, 23, 5565−
5572.
(52) von Schenck, H.; Strömberg, S.; Zetterberg, K.; Ludwig, M.;
Åkermark, B.; Svensson, M. Insertion Aptitudes and Insertion
Regiochemistry of Various Alkenes Coordinated to Cationic (σ-
R)(Diimine)Palladium(II) (R = − CH3, − C6H5). A Theoretical
Study. Organometallics 2001, 20, 2813−2819.
(53) Borkar, S.; Yennawar, H.; Sen, A. Methacrylate Insertion into
Cationic Diimine Palladium(II)−Alkyl Complexes and the Synthesis
of Poly(Alkene-Block-Alkene/Carbon Monoxide) Copolymers. Orga-
nometallics 2007, 26, 4711−4714.
(54) Mitsushige, Y.; Carrow, B. P.; Ito, S.; Nozaki, K. Ligand-
Controlled Insertion Regioselectivity Accelerates Copolymerisation of
Ethylene with Methyl Acrylate by Cationic Bisphosphine Monoxide−
Palladium Catalysts. Chem. Sci. 2016, 7, 737−744.
(55) Wucher, P.; Roesle, P.; Falivene, L.; Cavallo, L.; Caporaso, L.;
Göttker-Schnetmann, I.; Mecking, S. Controlled Acrylate Insertion
Regioselectivity in Diazaphospholidine-Sulfonato Palladium(II) Com-
plexes. Organometallics 2012, 31, 8505−8515.
(56) Ota, Y.; Ito, S.; Kuroda, J.; Okumura, Y.; Nozaki, K.
Quantification of the Steric Influence of Alkylphosphine−Sulfonate
Ligands on Polymerization, Leading to High-Molecular-Weight
Copolymers of Ethylene and Polar Monomers. J. Am. Chem. Soc.
2014, 136, 11898−11901.

L https://dx.doi.org/10.1021/acs.accounts.0c00628
Acc. Chem. Res. XXXX, XXX, XXX−XXX

You might also like