Next Article in Journal
Genome-Wide and Expression Pattern Analysis of the DVL Gene Family Reveals GhM_A05G1032 Is Involved in Fuzz Development in G. hirsutum
Next Article in Special Issue
Transient Behavior and Control of Polyethylene Production in a Fluidized Bed Reactor Utilizing Population Balance Model
Previous Article in Journal
SumVg: Total Heritability Explained by All Variants in Genome-Wide Association Studies Based on Summary Statistics with Standard Error Estimates
Previous Article in Special Issue
Theoretical and Experimental Study of Different Side Chains on 3,4-Ethylenedioxythiophene and Diketopyrrolopyrrole-Derived Polymers: Towards Organic Transistors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Synthesis of Non-Alternating Polyketone Generated by Copolymerization of Carbon Monoxide and Ethylene

by
Xieyi Xiao
,
Handou Zheng
,
Heng Gao
,
Zhaocong Cheng
,
Chunyu Feng
,
Jiahao Yang
and
Haiyang Gao
*
School of Materials Science and Engineering, PCFM Lab, GD HPPC Lab, Sun Yat-sen University, Guangzhou 510275, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(2), 1348; https://doi.org/10.3390/ijms25021348
Submission received: 29 December 2023 / Revised: 19 January 2024 / Accepted: 20 January 2024 / Published: 22 January 2024
(This article belongs to the Special Issue Synthesis of Advanced Polymer Materials 2.0)

Abstract

:
The copolymers of carbon monoxide (CO) and ethylene, namely aliphatic polyketones (PKs), have attracted considerable attention due to their unique property and degradation. Based on the arrangement of the ethylene and carbonyl groups in the polymer chain, PKs can be divided into perfect alternating and non-perfect alternating copolymers. Perfect alternating PKs have been previously reviewed, we herein focus on recent advances in the synthesis of PKs without a perfect alternating structure including non-perfect alternating PKs and PE with in-chain ketones. The chain structure of PKs, catalytic copolymerization mechanism, and non-alternating polymerization catalysts including phosphine–sulfonate Pd, diphosphazane monoxide (PNPO) Pd/Ni, and phosphinophenolate Ni catalysts are comprehensively summarized. This review aims to enlighten the design of ethylene/CO non-alternating polymerization catalysts for the development of new polyketone materials.

1. Introduction

Polyketones (PKs) are a kind of polymer containing the carbonyl groups in the main chain, which are usually made by the copolymerization of carbon monoxide (CO) and olefins [1,2,3]. According to the olefin comonomer type, PKs are divided into aliphatic (aliphatic alkene) [1,4,5], aromatic (vinyl arene) [6,7,8,9,10,11,12], and cyclic PKs (cycloolefin) [13,14,15,16]. Among these three kinds of PKs, aliphatic polymers generated by the copolymerization of olefins and CO have attracted considerable attention over the last decades, especially as a cheap and plentiful ethylene monomer [17,18]. In this review, we focus on aliphatic PKs generated from the copolymerization of ethylene and CO, and aliphatic PKs are called PKs in the following text.
As a kind of engineering plastic, PKs show unusual properties such as conspicuous chemical resistance, impermeability to hydrocarbons, high impact strength, and good thermostability [17,18]. Various modifications [19,20,21,22,23,24,25,26,27] and post-functionalizations of PKs widely expand their potential applications in biomedicine, electronic devices, fluorescent materials, and textile fibers [17,28,29,30,31]. Most importantly, PKs are a kind of degradable polymeric material, and their degradation can be induced by UV light irradiation through the well-known Norrish I and II processes [32,33,34,35,36,37]. Undoubtedly, PKs are highly attractive polymeric materials for increasing white plastic pollution. On the one hand, the synthesis of PKs needs to consume polluting gas CO stock. On the other hand, degradable PKs are environmentally friendly polymeric materials.
The first example of the metal-catalyzed synthesis of PKs was reported in 1951, in which CO/ethylene copolymerization was catalyzed by K2Ni[CN]4 to produce low-melting oligomers [38]. In the subsequent decades, various nickel and palladium catalysts have been developed for alternating copolymerizing ethylene and CO to synthesize aliphatic PKs [39,40,41,42,43,44,45,46,47]. A breakthrough is the diphosphine palladium catalysts discovered by Drent in 1991, which are highly efficient for alkene/CO copolymerization [48]. In 1996, a CO/ethylene/propylene terpolymer called “Carilon” was successfully commercialized by Shell company, but production of “Carilon” was closed in 2000 due to strategic changes [2]. Hyosung company began to produce the same PKs in 2013, and the current annual production of PKs has reached 50,000 tons [49]. These engineering plastics are commercially available for food contact, industry, and medical equipment.
Because the insertion of CO into the five-membered Pd-acyl-chelates is kinetically favored over ethylene insertion and successive insertion of CO is thermodynamically prohibited, perfect alternating PKs are often produced by Ni/Pd catalyzed copolymerization of ethylene and CO. In the early period, perfect alternating PKs received considerable attention in both academic and industrial fields (Figure 1). Recently, PKs without a perfect alternating structure, including non-perfect alternating PKs and polyethylene (PE) with in-chain ketones, have also attracted increasing interest (Figure 1). In 2002, non-alternating CO/ethylene copolymerization was first realized using phosphine–sulfonate Pd catalysts developed by Drent [50]. With improved flexibility, non-alternating PKs have potential applications in many industrial fields such as textile fibers and packing films. PE with in-chain ketones was recently developed and exhibited similar mechanical properties to commercial high-density polyethylene (HDPE) and better degradability and hydrophily [51,52]. Introducing a small amount of CO in polyolefins provides a promising strategy to balance the demand for polyolefins and expectations of environmental friendliness, and PE with in-chain ketones can be understood as degradable HDPE.
Perfect alternating PKs have been systematically summarized in previous reviews [1,3,4,5,53,54,55,56,57]. In this review covering the literature published until the end of 2023, we focus on recent advances in the synthesis of PKs without a perfect alternating structure including non-perfect alternating PKs and PE with in-chain ketones. The chain structure of PKs without a perfect alternating structure, catalytic copolymerization mechanism, and Ni/Pd catalysts are summarized.

2. Chain Structure of PKs without a Perfect Alternating Structure

PKs without a perfect alternating structure have sequential ethylene units in a polymer chain, which originates from sequential ethylene insertion during the copolymerization of ethylene and CO. The resulting microstructure and melting temperatures (Tm) of three kinds of PKs are shown in Table 1.
Based on the arrangement of the ethylene and carbonyl groups, the structural fragments and corresponding NMR shifts of PKs without a perfect alternating structure are classified in Table 2.
Carbonyl content and extra ethylene (C2) insertion are the two most important structural parameters of PKs, defined as total mol % CO units relative to all monomer units [59] and total mol % ethylene of copolymer involved in non-alternation [50], respectively. Carbonyl content can be calculated by 1H NMR according to the formula reported by Sen [59]:
C a r b o n y l   c o n t e n t = I 2 + I 4 + I 7 + I 11 2 × I 2 + I 4 + I 7 + I 11 + I 5 + I 8 + I 12 + ( I 9 + I 13 ) × 100 %
I2, I4, I5, I7, I8, I9, I11, I12, and I13 are the integral of 1H NMR signal of corresponding methylene.
Extra ethylene (C2) insertion can be calculated by 13C NMR (signal of methylene) according to the formula reported by Chen [58]:
e x t r a   e t h y l e n e   ( C 2 )   i n s e r t i o n = I 4 + I 7 + I 11 I 2 + I 4 + I 7 + I 11 × 100 %
I2, I4, I7, and I11 are the integral of 13C NMR signal of corresponding methylene.
In the case of PE with in-chain ketones, in addition to carbonyl content, the proportion of different carbonyl groups (1, 3, 6, and 14) in PKs also becomes important because the carbonyl content is too low to display obvious differences [51,52,60,61,62]. The 13C NMR signals of different carbonyl groups (1, 3, 6, and 14) and adjacent methylene (2, 4, 7, and 15) almost relatively isolate each other; therefore, the proportion of each kind of carbonyl can be calculated. Due to the low carbonyl content, the replacement of 13CO stock is used to enhance the sensitivity of the characteristic carbonyl [51,52]. Owing to the unique and intense signal of carbonyl, IR spectroscopy provides another practical measurement of the carbonyl content of PE with in-chain ketones, and a series of PKs with known carbonyl contents is need to draw a standard curve [52,63,64].
The relationship between the properties and microstructure of the copolymer has been systematically studied by Müller and Mecking [65,66]. Properties of PKs without a perfect alternating structure, such as Tm, thermal decomposition temperature, and melt viscosity, for the microstructure, have a strong dependence on the copolymer, the CO content, and molecular weight, and the concentration of methyl branches along the chain directly determines the properties of the copolymer. In general, the properties of the copolymer are close to that of PE when the carbonyl content is low.
When the carbonyl content is high (>35%), the Tm of the copolymer decreases with the decrease of the carbonyl content because the multiple ethylene insertion destroys the dipolar interaction between the carbonyl groups in the alternating structure [50,67,68,69]. However, at a low carbonyl content (<30%), the properties of the copolymer are close to PE, and the decrease in carbonyl content usually leads to an increase in crystallinity, so the Tm of the copolymer decreases with the decrease in carbonyl content [59].

3. Catalytic Copolymerization Mechanism

Successive CO insertion is thermodynamically prohibited in CO/ethylene copolymerization [56]. Therefore, the key to obtain alternating or non-alternating copolymers is whether the ethylene sequential insertion with unfavorable kinetics can be achieved by the improvement of the catalyst system and the availability of copolymerization conditions.
Phosphine–sulfonate ligand palladium catalyst systems achieved the first example of non-alternating copolymerization of CO and ethylene, of which the catalytic mechanisms have been investigated scientifically [59,70,71,72]. Therefore, take the example of typical bidentate [P, P] palladium complexes (alternating copolymerization catalysts) and neutral phosphine–sulfonate ligated Pd complexes (Scheme 1), the origins of sequential insertions of ethylene are proved to be (1) destabilization of Pd-acyl β-chelates A (phosphine-sulfonate); (2) facile decarbonylation of six-membered Pd-acyl chelates D [67,68,70]. As shown in Scheme 1, five-membered Pd-acyl β-chelates A existing in CO/ethylene copolymerization catalyzed by bidentate [P, P] and phosphine–sulfonate palladium complexes, can accept insertions from ethylene or CO [67]. The stability of β-chelates A is lower in the case of phosphine–sulfonate ligands; therefore, the chelate cycle is easily opened by ethylene [68,71]. The binding affinities of CO and ethylene to β-chelates A have been measured by Sen in 2009 [59]. Although the ratio of binding affinities determined credibly by NMR seems reliable enough to explain the multiple insertions of ethylene, Sen believes decarbonylation also plays a crucial part in the non-alternation [59]. Thus, the second reason involving decarbonylation of six-membered Pd-acyl chelates D can be considered as a supplement to the first reason, which together constitute the reasons for the non-alternating copolymerization ability of neutral phosphine–sulfonate ligated palladium complexes. The relatively facile decarbonylation is also related to the instability of the corresponding Pd-acyl chelates D of phosphine–sulfonate ligands [59,71].
Although the destabilization of Pd-acyl β-chelates and decarbonylation could partly explain why non-alternating copolymerization occurs, most mechanisms cannot escape the scope of phenomenological theory. In fact, it is extremely difficult to find a direct connection between the ability of non-alternating copolymerization and the structure of ligands or catalysts. Therefore, there is still no perfect theory to predict new catalyst systems that can achieve non-alternating copolymerization of CO and ethylene.
However, some researchers have made outstanding contributions to exploring the nature of the origin of non-alternating copolymerization [72,73,74,75,76]. In different unique polymerizations realized by phosphine sulfonate palladium catalysts, similar cis/trans isomerization (Scheme 2) of complexes and consistency in the high stability of cis σ-complex intermediates are useful to unravel the deep mechanism of non-alternating copolymerization [72,73].
Cis/trans isomerization is the connecting bridge of ethylene coordination and ethylene insertion. The cis σ-complex and trans π-complex are under equilibrium, and isomerization takes place readily in almost all cases. [72,73] Equilibrium with a high kinetic constant cannot directly determine the kinetically favorable step, because CO/ethylene non-alternating copolymerization is a result of kinetic regulation. However, the resulting concentration of the trans π-complex influences the kinetic trend of the monomer insertion to some extent [73]. In the case of CO/ethylene non-alternating copolymerization, sequential ethylene insertion must occur in the trans π-complex, and the concentration of the trans π-complex depends on the cis/trans equilibrium constant due to the stability of the cis σ-complex and trans π-complex and the structure of catalysts. More importantly, cis/trans isomerization is also assumed in recent novel examples of CO/ethylene non-alternating copolymerization [63,64], which implies that it might be a concordant insight into the non-alternating pathway.
Charge effects and orbital interactions were also proposed to explain the particularity of phosphine sulfonate ligands, and complexes with neutral charge and atoms with lone pairs covalent bonding Pd promote the multiple insertions of ethylene [2,73,74]. Orbital interactions specifically refer to an appropriate stereoelectronic effect of ligands and metal centers. The interactions involving lone pairs in p orbital on the ligated oxygen and dπ orbital on palladium atom in the case of phosphine sulfonate ligands palladium catalyst systems (Figure 2). Generally, the SO3 group withdraws the π-electrons of the adjacent group. When an adjacent group connects directly with an oxygen atom, repulsion between one pair on the oxygen atom and π-electrons on Pd possibly causes an increase in the electron density of palladium. Resultantly, the back-donation from palladium to the antibonding orbital of ethylene is enhanced and the binding affinity of ethylene and metal may be strengthened, thereby enhancing the stabilization of the corresponding trans π-complex [53,74].
Notably, similar stereoelectronic structures exist in both diphosphazane monoxide and phosphinophenolate systems (Figure 3), which are two important catalysts for CO/ethylene non-alternating copolymerization (see Section 4). The striking similarity in stereoelectronic structure is the most critical factor in CO/ethylene non-alternating copolymerization. Orbital interactions provide a new approach to catalyst design and are helpful in developing new CO/ethylene non-alternating copolymerization catalysts.

4. Non-Alternating Copolymerization Catalysts

Due to the unfavorable kinetics of ethylene sequential insertion, designing catalysts for the synthesis of PKs without a perfect alternating structure is extremely difficult. Until recently, few examples of catalysts achieved ethylene sequential insertion. Phosphine–sulfonate Pd catalyst systems, diphosphazane monoxide (PNPO) Pd/Ni catalysts, and phosphinophenolate Ni catalysts represent the three main catalyst systems for CO/ethylene non-alternating copolymerization [50,51,58]. In addition, phosphanylferrocene-carboxylate ligand Pd catalysts [77] and zwitterionic Ni catalysts [78] were also reported to synthesize non-alternating PKs with low activity (<15 g PK·mmol−1·h−1).

4.1. Phosphine–Sulfonate Pd Catalysts

In 2002, Drent first reported the neutral palladium catalysts generated in situ from phosphine–sulfonate ligands [79]. The first example of a non-alternating linear CO/ethylene copolymer with successive ethylene incorporation was also reported by Drent by using a phosphine–sulfonate palladium catalyst system [50].
A new series of [P,O] palladium catalysts Pd1~Pd3 (Figure 4) with ortho-alkoxy substituents generated in situ from Pd(OAc)2 and phosphine–sulfonate ligands were employed for CO/ethylene copolymerization. Despite the activities of copolymerization with these neutral catalysts (49~190 g PK·mmol·Pd−1·h−1) being low compared to classical 1,3-bis(diphenylphosphino)propane (dppp)-based palladium systems, these neutral catalysts produced non-alternating copolymers with 2.4~18.3% non-alt insertions (extra C2 insertion) [50]. The non-alternating insertions increase with the steric hindrance of the ortho-alkoxy substituents in the palladium catalytic system (OiPr > OEt > OMe). Furthermore, the melting points of non-alternating copolymers made by catalysts from Pd1 (234 °C), Pd2 (231 °C) and Pd3 (229 °C) reduce, reflecting the impact of steric hindrance in the palladium catalytic system on the microstructure of the resulting PKs.
The non-alternating CO/ethylene copolymerization using phosphine–sulfonate palladium systems is attributed to the fact that ethylene insertion can take place from the most stable chelate complex “cis σ-complex” via the intermediate “trans π-complex” (see the “Catalytic copolymerization mechanism” section) [73]. By examining the kinetic and thermodynamic parameters, Sen discovered that the catalyst exhibits a remarkably minor disparity in its binding affinities for ethylene and carbon monoxide [59]. However, the difference in monomer binding affinity alone is not sufficient to explain the degree of non-alternation actually observed in this system. Thus, decarbonylation plays a significant role in the non-alternation. It has been observed that the steric bulk prevents the formation of chelate bonds between the carbonyl oxygens of the PKs chain and the metal center, which facilitates the decarbonylation of an acyl group. This process leads to successive ethylene incorporation and ultimately results in the formation of non-alternating units [59,70,71].
Although successive CO insertions are thermodynamically impossible, CO exhibits a stronger binding affinity to the active palladium species and tends to insert more readily into the Pd-alkyl bond than ethylene. Consequently, obtaining PKs with low CO content is greatly challenging. Sen and coworkers successfully achieved the formation of non-perfect alternating PKs with as low as 10 mol% CO in the Pd1 (Figure 4) catalytic system, although the yields were relatively low [69]. Sen and coworkers systematically controlled the monomer feed ratio and reaction temperature, and non-perfect alternating PKs with CO incorporation ranging from 0 to 50 mol% were produced [59].
Muller successfully prepared low-molecular-weight non-alternating PKs (<6 kg·mol−1) with varying CO contents (0~50 mol%) in the Pd1 (Figure 4) [65]. Segments with high CO density result in domains with strong intermolecular dipole–dipole interactions, which play a dominant role in the phase change behavior of PKs. Additionally, the melt viscosity decreases rapidly with decreasing molecular weight of PKs. The crystallinity of the non-perfect alternating PKs is directly related to the -(C2H4)n- segment length. Additionally, the presence of methyl branches along the chain significantly affects the material properties in the solid state, resulting in reduced crystallinity and lower enthalpy of melting.
The unique properties of phosphine–sulfonate palladium catalysts have sparked widespread research interest in the synthesis of non-alternating PKs. Rieger and coworkers reported the palladium catalysts Pd4~Pd6 (Figure 4) with modification of both the benzene ring and phosphine moieties for non-alternating CO/ethylene copolymerizations [67]. Pd4 does not produce any significant amounts of polymer under the conditions used, and Pd5 and Pd6 can yield high molecular weight PKs (Mw up to 370 kg·mol−1) with 2.4~18.3% extra C2 insertions and a melting temperature of 220~230 °C.
Rieger further reported the palladium catalysts Pd7~Pd9 (Figure 5) and two monosubstituted palladium catalysts, Pd10 and Pd11 (Figure 5), for CO/ethylene copolymerizations [67]. Additionally, Pd7 does not produce PK, but Pd8 and Pd9 exhibit high polymerization activities (2.5~106 g PK·mmol·Pd−1·h−1) and thermal stability compared to the corresponding in situ generated palladium catalysts. Pd9 also exhibits substantial copolymerization activity of 75 g PK·mmol·Pd−1·h−1 at 200 °C. However, the obtained PK has lower extra C2 insertion (<4.9 mol%) [67]. When 2 wt% B(C6F5)3 was used as an activator, the monosubstituted palladium Pd10 and Pd11 (Figure 5) showed increased copolymerization activities (up to 292 g PK·mmol·Pd−1·h−1) to produce PKs with remarkably high extra ethylene insertions (up to 26.4%) [67].
In addition to modifying the phosphine moiety, Bianchini and Oberhauser conducted a comparative study on the impact of the backbones of phosphine–sulfonate ligands on the performance of palladium catalysts [68]. They investigated the non-alternating CO/ethylene copolymerization catalyzed by palladium catalysts Pd12~Pd15 (Figure 6) with modified phosphine–sulfonate ligands. Pd12 and Pd14 with rigid phenyl backbones exhibited much higher activity (41~607 g PK·mmol·Pd−1·h−1) than Pd13 and Pd15 (0.4~5.4 g PK·mmol·Pd−1·h−1) with the more flexible ligands. The presence of benzoquinone (BQ) oxidant has no appreciable influence on extra C2 insertion while increasing productivity in copolymerization catalyzed by Pd12. Additionally, the former produces copolymers with a significantly higher molecular weight (Mw > 30 kg·mol−1). The authors think that the decreased copolymerization activities observed in Pd13 and Pd15 with flexible ligands could be mainly attributed to the faster degradation of catalytically active species. Furthermore, palladium catalysts Pd12 and Pd14 with a rigid backbone produced PKs with higher extra-ethylene insertions (up to 27.8 mol%) compared to Pd13 and Pd15 (0.2~23.2 mol%).
The ferrocene moiety provides quite a variety of possibilities for electronically modifying catalyst systems derived from chelate ligands. Erker and coworkers reported two in situ phosphine–sulfonate palladium catalysts, Pd16 and Pd17 (Figure 7), and Pd18 complexes with a ferrocene-derived ligand for CO/ethylene copolymerization [80]. At a relatively low CO/ethylene feed, these catalysts produced nearly alternating PKs. When the CO/ethylene partial pressure ratio is approximately 1:10, substantially non-alternating PKs were obtained. The overall catalyst activities decreased as the CO partial pressure decreased, but the activities were still quite remarkable (18~159 g PK·mmol·Pd−1·h−1). Pd17 exhibited higher copolymerization activity than Pd16 in spite of a lower percentage of extra ethylene insertions (<1 mol%). The addition of B(C6F5)3 as a Lewis acid activator increased the activity of Pd18 while decreased extra C2 insertion. By precisely controlling the conditions, it is possible to prepare PKs with different amounts of extra ethylene (1~27 mol%) and varying polymer melting temperatures (167~242 °C).
Nozaki also reported phosphine–sulfonate palladium catalysts Pd19~Pd22 (Figure 8) for the copolymerization of ethylene with metal carbonyls as the CO source [60]. In contrast to conventional CO/ethylene copolymerization reactions, excellent non-alternating selectivity has been achieved with moderate activity (1.2~54 g PK·mmol·Pd−1·h−1). Linear high molecular weight (44 kg·mol−1) polyethylene bearing low content isolated in-chain carbonyls (<4%) was obtained. Although the copolymer retained the properties of polyethylene, it exhibited faster degradation under UV light irradiation compared to pure PE.
Jian reported a non-alternating (>99%) terpolymerization of CO with ethylene and fundamental polar monomers using palladium catalysts Pd23~Pd26 (Figure 8) [81]. This palladium system effectively prevented the formation of undesired polar monomer chelates such as acrylate (MA, BA), acrylic acid (AA), vinyl ether (VE), and acrylonitrile (AN). High molecular weight linear PE was synthesized with both in-chain isolated carbonyl (>99%) and main-chain functional groups (0.18~2.38 mol%). The incorporation of these low-content isolated carbonyls allows for the photodegradation of the PE. Additionally, the introduction of polar functional groups significantly enhances their surface properties.
Table 3 lists a summary of representative results for non-alternating copolymerization activity, extra C2 insertion, and melting temperature in the phosphine–sulfonate palladium catalyst systems.

4.2. Diphosphazane Monoxide Pd/Ni Catalysts

In 2018, Chen discovered a new type of electronic asymmetric diphosphazane monoxide (PNPO) ligand (Figure 5) with multiple modifiable moieties for copolymerization of ethylene with polar monomers [82,83,84]. The short-bite ligand platform increases ligand rigidity, and unique electronic asymmetry achieved by a strongly σ-donating phosphine moiety and a weakly σ-donating phosphine oxide moiety is believed to be responsible for its unexpectedly excellent catalytic property. PNPO Ni/Pd catalysts for CO/ethylene non-alternating copolymerization are shown in Figure 9, and the corresponding non-alternating copolymerization results are summarized in Table 4.
Chen developed a series of catalysts Pd27~Pd40 (Figure 9) with different electronic effect substituents on phosphine moieties (Ar1, Ar2) and N-aryl moieties (Ar3) [58]. Notably, the introduction of electron-donating substituents at the N-aryl moiety (Figure 10) of PNPO ligands (Pd27~Pd28, Pd29~Pd32, Pd33~Pd35, and Pd36~Pd37) caused increasing activities and Mn. More importantly, the carbonyl content also decreased as the electron-donating ability (–CF3 < –H < –tBu < –OMe < –NMe2) of the substituents at the N-aryl moiety increased [57]. The introduction of different substituents on phosphine moiety (Pd32, Pd35, Pd37, Pd38, and Pd39) significantly affects the activity and carbonyl content in the copolymer. Pd38 with Ar3 = 2,6-OMeC6H3 exhibits the highest activity and the lowest carbonyl content, while Pd39 with stronger electron-donating Ar3 = 2-NMe2C6H4 is inactive because of the coordination mode change [57]. Catalyst Pd40 bearing isopropyl affords PK with high carbonyl content (49%) and low activity (25.5 g PK·mmol·Pd−1·h−1). The representative catalyst Pd38 is highly efficient for CO/ethylene non-alternating copolymerization. With an increasing CO/ethylene feed ratio from 1/1 to 1/20, the resultant PKs have alternating to non-alternating structures with a decreasing melting temperature (Tm) from 248 °C to 147 °C [58].
Liu has reported Pd41~Pd50 for non-alternating copolymerization of CO and ethylene [63]. Compared with Pd32, Pd41~Pd43 bearing an aliphatic substituent on the amine moiety showed non-alternating copolymerization with almost two times higher activity and higher carbonyl content. Pd44 bearing benzylamine and Ar3(2,6-OMeC6H3) (Figure 10) is highly efficient (activity up to 124.1 g PK·mmol·Pd−1·h−1) for CO/ethylene non-alternating copolymerization. Liu also found that steric hindrance and electron-donating ability of the substituent on phosphine oxide moiety are beneficial to extra ethylene insertion, thus Pd45~Pd49 afforded PKs with relatively high extra ethylene insertion (12.2%) [63]. In addition, Pd50 bearing -NMe2 on the phosphine oxide moiety has a different complexation mode, and it afforded PKs with low activity and carbonyl content.
Liu also developed a series of PNPO nickel/palladium catalysts, and PE with in-chain ketones was synthesized by Ni1~Ni8, Pd29~Pd32, and Pd38 [61]. In a very low CO/ethylene feed of 0.2%, copolymers were produced in low carbonyl content and low Tm (<140 °C). Both steric hindrance and the electronic nature of ligands have significant impacts on activity, Mn, and carbonyl content. With increasing electron-donating ability of the substituents at the N-aryl moiety from Ni1 to Ni4, the resultant copolymers’ Mn decreases, but their carbonyl content and activity exhibited uncertain regularity. Ni5 and Ni8 bearing the strongest electron-donating and electron-withdrawing substituents on phosphine moiety are inactive, while Ni6 and Ni7 afforded a copolymer with the highest activity (301 g PK·mmol·Ni−1·h−1). Palladium-based catalysts Pd29~Pd32 and Pd38 were also used to prepare PE with in-chain ketones. The electron-donating substituent on arylamine moieties (Pd29~Pd32) increased both activity and Mn [58,61]. Pd38 bearing a strong electron-donating phosphine moiety is the most efficient and afforded copolymers with the highest Mn (93.3 kg·mol−1) in 0.1% CO/ethylene feed.

4.3. Phosphinophenolate Ni Catalysts

Copolymerization of CO and ethylene catalyzed by Ni complexes is extremely challenging owing to the strong electrophilic nature of Ni. Most Ni catalysts only produced alternating PKs [85,86,87,88,89]. Phosphinophenolate nickel catalysts were first developed for the copolymerization of ethylene and acrylates [90,91]. In 2021, Mecking reported CO/ethylene non-alternating copolymerization using phosphinophenolate Ni complexes (Ni9~Ni14) to yield in-chain ketone PE materials. In 0.2 mol % CO feed, PE with in-chain ketones was obtained with a Tm low of 134 °C [51]. This material maintains the excellent mechanical and processing properties of traditional PE and also possesses the degradability of PKs [51]. Phosphinophenolate Ni catalysts for CO/ethylene non-alternating copolymerization are summarized in Figure 11, and the corresponding non-alternating copolymerization results are summarized in Table 5.
Generally, phosphinophenolate Ni complexes have a relatively complex structure and multiple modification modes [51,64]. Unlike phosphine sulfonate Pd catalysts, the rate-determining step of alternating and non-alternating steps in phosphinophenolate Ni-catalyzed CO/ethylene copolymerization were identified as the opening of the six-membered C,O-chelate by ethylene coordination and cis/trans isomerization of an alkyl-olefin intermediate, respectively (Figure 12) [64]. Short duration copolymerization experiments and theoretical studies were systematically studied, and the details of the connection between the catalyst structure and rate-determining step were unraveled. The steric effect at the aromatic rings on the phosphine moiety apically coordinated to the metal center significantly impacts the free energy barrier of cis/trans isomerization and the free energy barrier of the ethylene coordination step (Figure 12) [64]. Simultaneously, an η2-coordination of a P-bound aromatic moiety axially oriented to the metal center, which is considered as a crucial feature of catalysts, modulates the competition between alternating and non-alternating pathways.
In general, the influence of catalyst structure on competition can be summarized as follows. (1) Bulkier steric hindrance, enhanced by the η2-coordination mentioned above, impedes the inversion of the intermediate conformation during copolymerization, thereby increasing the energy barrier of cis/trans isomerization and finally weakening the tendency for non-alternating polymerization. (2) The electronic effect of ligands directly determines the electron density at the Ni center. The strong electrophilicity of Ni easily causes a very stable resting state, which prohibits both alternating and non-alternating pathways. In contrast, strong electron-donating groups lead to low electrophilicity Ni, which is favorable for alternating polymerization because of the easy opening of the six-membered C,O-chelate [51,64]. Interestingly, phenolate substitution has few effects on the microstructure of copolymers compared with phosphine substituents [64].
In addition, Nozaki also reported phosphinophenolate Ni catalysts for the synthesis of PE with in-chain ketones using metal carbonyls as a carbonyl source [62]. High selectivity to isolated carbonyl and higher activity were realized compared to the Pd catalysts reported by them.

5. Conclusions and Outlook

Aliphatic PKs generated from the copolymerization of CO and ethylene have attracted considerable attention over the last decades due to their unique properties and environmental friendliness. Alternating PKs have a high Tm and insolubility, while PKs without a perfect alternating structure have better processability. However, the synthesis of PKs without a perfect alternating structure is extremely difficult because of the unfavorable kinetics of ethylene sequential insertion. Phosphine–sulfonate Pd catalyst systems, PNPO Pd/Ni catalysts, and phosphinophenolate Ni catalysts represent the three main catalysts for CO/ethylene non-alternating copolymerization. Catalytic activity and the carbonyl content of PKs are two important parameters to evaluate catalysts. Interestingly, stereoelectronic structure and the orbital interactions involving lone pairs in the p orbital on the oxygen donor and dπ orbital on the metal atom are observed in three kinds of catalysts, which open a window for the design of non-alternating ethylene/CO copolymerization catalysts.
Despite significant breakthroughs in non-alternating copolymerization of CO in recent years, there are still several challenges for future development in this field: (1) modulating the carbonyl content of PKs by CO/ethylene feed in copolymerization often causes a rapid decrease in activity. (2) There is no widely acceptable theory to explain the relationship between the structure of catalysts and non-alternating copolymerization, therefore designing new catalysts seems to only depend on a large number of experiments. (3) Noble palladium catalyst represents the most successful catalytic system, while the replacement of palladium by nickel or other non-noble transition metals poses significant challenges.

Author Contributions

X.X.: Conceptualization, Formal analysis, Writing—original draft, and Methodology. H.Z.: Writing—original draft and Methodology. H.G. (Heng Gao): Formal analysis and Methodology. Z.C.: Methodology, Review, and Editing. C.F.: Formal analysis and Review. J.Y.: Methodology and Formal analysis. H.G. (Haiyang Gao): Supervision, Project administration, Conceptualization, Methodology, Funding acquisition, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the State Key Research Development Program of China (Grant No. 2021YFB3800701), National Natural Science Foundation of China (NSFC) (52173016), and PetroChina Scientific and Technological Projects (2022DJ6308, 2021DJ5902, and 2020-CB-02-13), which are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Drent, E.; Budzelaar, P.H. Palladium-catalyzed alternating copolymerization of alkenes and carbon monoxide. Chem. Rev. 1996, 96, 663–682. [Google Scholar] [CrossRef] [PubMed]
  2. Nakamura, A.; Ito, S.; Nozaki, K. Coordination-insertion copolymerization of fundamental polar monomers. Chem. Rev. 2009, 109, 5215–5244. [Google Scholar] [CrossRef] [PubMed]
  3. Bianchini, C.; Meli, A. Alternating copolymerization of carbon monoxide and olefins by single-site metal catalysis. Coordin. Chem. Rev. 2002, 225, 35–66. [Google Scholar] [CrossRef]
  4. Müller, G.; Rieger, B. Propene based thermoplastic elastomers by early and late transition metal catalysis. Prog. Polym. Sci. 2002, 27, 815–851. [Google Scholar] [CrossRef]
  5. Durand, J.; Milani, B. The role of nitrogen-donor ligands in the palladium-catalyzed polyketones synthesis. Coordin. Chem. Rev. 2006, 250, 542–560. [Google Scholar] [CrossRef]
  6. Xiao, Z.; Zheng, H.; Du, C.; Zhong, L.; Liao, H.; Gao, J.; Gao, H.; Wu, Q. Enhancement on alternating copolymerization of carbon monoxide and styrene by dibenzobarrelene-based α-diimine palladium catalysts. Macromolecules 2018, 51, 9110–9121. [Google Scholar] [CrossRef]
  7. Liao, G.; Xiao, Z.; Chen, X.; Du, C.; Zhong, L.; Cheung, C.S.; Gao, H. Fast and regioselective polymerization of para-alkoxystyrene by palladium catalysts for precision production of high-molecular-weight polystyrene derivatives. Macromolecules 2019, 53, 256–266. [Google Scholar] [CrossRef]
  8. Du, C.; Chu, H.; Xiao, Z.; Zhong, L.; Zhou, Y.; Qin, W.; Liang, G.; Gao, H. Alternating vinylarene–carbon monoxide copolymers: Simple and efficient nonconjugated luminescent macromolecules. Macromolecules 2020, 53, 9337–9344. [Google Scholar] [CrossRef]
  9. Wang, S.; Wang, L.; Zhong, L.; Xu, R.; Wang, X.; Kang, W.; Gao, H. C1-symmetric tert-butyl substituted pyridylamido hafnium complex for ethylene, α-olefin, and styrene polymerizations. Eur. Polym. J. 2020, 131, 109709–109719. [Google Scholar] [CrossRef]
  10. Du, C.; Cheung, C.S.; Zheng, H.; Li, D.; Du, W.; Gao, H.; Liang, G.; Gao, H. Bathochromic-shifted emissions by postfunctionalization of nonconjugated polyketones. ACS Appl. Mater. Interfaces 2021, 13, 59288–59297. [Google Scholar] [CrossRef]
  11. Xiao, Z.; Zhong, L.; Du, C.; Du, W.; Zheng, H.; Cheung, C.S.; Wang, L.; Gao, H. Unprecedented steric and positioning effects of comonomer substituents on α-diimine palladium-catalyzed vinyl arene/CO copolymerization. Macromolecules 2021, 54, 687–695. [Google Scholar] [CrossRef]
  12. Cheung, C.S.; Shi, X.B.; Pei, L.X.; Du, C.; Gao, H.; Qiu, Z.L.; Gao, H.Y. Alternating copolymerization of carbon monoxide and vinyl arenes using [N,N] bidentate palladium catalysts. J. Polym. Sci. 2022, 60, 1448–1467. [Google Scholar] [CrossRef]
  13. Abu-Surrah, A.S.; Lappalainen, K.; Kettunen, M.; Repo, T.; Leskelä, M.; Hodali, H.A.; Rieger, B. Homo- and copolymerization of strained cyclic olefins with new palladium(II) complexes bearing ethylene-bridged heterodonor ligands. Macromo. Chem. Phys. 2001, 202, 599–603. [Google Scholar] [CrossRef]
  14. Kettunen, M.; Abu-Surrah, A.S.; Repo, T.; Leskelä, M. Copolymerization of carbon monoxide with-methylenecycloalkane and dienes:: Synthesis of functionalized aliphatic polyketones. Polym. Int. 2001, 50, 1223–1227. [Google Scholar] [CrossRef]
  15. Wang, X.; Seidel, F.W.; Nozaki, K. Synthesis of polyethylene with in-chain alpha,beta-unsaturated ketone and isolated ketone units: Pd-catalyzed ring-opening copolymerization of cyclopropenone with ethylene. Angew. Chem. Int. Ed. Engl. 2019, 58, 12955–12959. [Google Scholar] [CrossRef] [PubMed]
  16. Dai, Y.; Luo, J.; Liu, T.; Jia, L. Dual-site catalysis for sustainable polymers to replace current commodity polymers—Carbonylative copolymerization of ethylene, ethylene oxide, and tetrahydrofuran. Chem. Commun. 2020, 56, 15341–15344. [Google Scholar] [CrossRef] [PubMed]
  17. Yang, Y.; Li, S.Y.; Bao, R.Y.; Liu, Z.Y.; Yang, M.B.; Tan, C.B.; Yang, W. Progress in polyketone materials: Blends and composites. Polym. Int. 2018, 67, 1478–1487. [Google Scholar] [CrossRef]
  18. Zong, Y.L.; Li, Q.K.; Mu, H.L.; Jian, Z.B. Palladium promoted copolymerization of carbon monoxide with polar or non-polar olefinic monomers. Curr. Org. Chem. 2021, 25, 287–300. [Google Scholar] [CrossRef]
  19. Nozaki, K.; Kosaka, N.; Graubner, V.M.; Hiyama, T. Methylenation of optically active γ-polyketones. Polym. J. 2002, 34, 376–382. [Google Scholar] [CrossRef]
  20. Kosaka, N.; Hiyama, T.; Nozaki, K. Baeyer-Villiger oxidation of an optically active 1,4-polyketone. Macromolecules 2004, 37, 4484–4487. [Google Scholar] [CrossRef]
  21. Kosaka, N.; Oda, T.; Hiyama, T.; Nozaki, K. Synthesis and photoisomerization of optically active 1,4-polyketones substituted by azobenzene side chains. Macromolecules 2004, 37, 3159–3164. [Google Scholar] [CrossRef]
  22. Pérez-Foullerat, D.; Hild, S.; Mücke, A.; Rieger, B. Synthesis and properties of poly (ketone-co-alcohol) materials: Shape memory thermoplastic elastomers by control of the glass transition process. Macromo. Chem. Phys. 2004, 205, 374–382. [Google Scholar] [CrossRef]
  23. Zhang, Y.C.; Broekhuis, A.A.; Stuart, M.C.A.; Picchioni, F. Polymeric Amines by chemical modifications of alternating aliphatic polyketones. J. Appl. Polym. Sci. 2008, 107, 262–271. [Google Scholar] [CrossRef]
  24. Toncelli, C.; Bouwhuis, S.; Broekhuis, A.A.; Picchioni, F. Cyclopentadiene-functionalized polyketone as self-cross-linking thermo-reversible thermoset with increased softening temperature. J. Appl. Polym. Sci. 2015, 133, 42924. [Google Scholar] [CrossRef]
  25. Ataollahi, N.; Girardi, F.; Cappelletto, E.; Vezzù, K.; Di Noto, V.; Scardi, P.; Callone, E.; Di Maggio, R. Chemical modification and structural rearrangements of polyketone-based polymer membrane. J. Appl. Polym. Sci. 2017, 134, 46549. [Google Scholar] [CrossRef]
  26. Hwang, S.Y.; Cho, Y.S.; Kim, T.; Jung, Y.S.; Hwang, T.S. Synthesis of highly durable sulfonated polyketone fibers by direct sulfonation reaction and their adsorption properties for heavy metals. Macromol. Res. 2020, 28, 336–342. [Google Scholar] [CrossRef]
  27. Luleburgaz, S.; Hizal, G.; Tunca, U.; Durmaz, H. Modification of polyketone via chlorodimethylsilane-mediated reductive etherification reaction: A practical way for alkoxy-functional polymers. Macromolecules 2021, 54, 5106–5116. [Google Scholar] [CrossRef]
  28. Agostinelli, E.; Belli, F.; Tempera, G.; Mura, A.; Floris, G.; Toniolo, L.; Vavasori, A.; Fabris, S.; Momo, F.; Stevanato, R. Polyketone polymer: A new support for direct enzyme immobilization. J. Biotechnol. 2007, 127, 670–678. [Google Scholar] [CrossRef] [PubMed]
  29. El Ghanem, H.M.; Abdul Jawad, S.D.; Al-Saleh, M.H.; Hussain, Y.A.; Abu-Surrah, A.S. Electrical impedance spectroscopic study of CNT/ethylene-alt-CO/propylene-alt-CO polyketones nanocomposite. J. Macromol. Sci. 2014, 53, 878–892. [Google Scholar] [CrossRef]
  30. Won, J.S.; Jin, D.Y.; Lee, J.E.; Lee, S.G. Effects of finishing conditions on the properties of polyketone fiber. Fiber. Polym. 2015, 16, 1908–1916. [Google Scholar] [CrossRef]
  31. Ataollahi, N.; Vezzù, K.; Nawn, G.; Pace, G.; Cavinato, G.; Girardi, F.; Scardi, P.; Di Noto, V.; Di Maggio, R. A polyketone-based anion exchange membrane for electrochemical applications: Synthesis and characterization. Electrochim. Acta 2017, 226, 148–157. [Google Scholar] [CrossRef]
  32. Mu, J.L.; Fan, W.J.; Shan, S.Y.; Su, H.Y.; Wu, S.S.; Jia, Q.M. Thermal degradation kinetics of polyketone based on styrene and carbon monoxide. Thermochim. Acta 2014, 579, 74–79. [Google Scholar] [CrossRef]
  33. Reeves, J.A.; Allegrezza, M.L.; Konkolewicz, D. Rise and Fall: Poly(phenyl vinyl ketone) photopolymerization and photodegradation under visible and uv radiation. Macromol. Rapid. Commun. 2017, 38, 1600623. [Google Scholar] [CrossRef] [PubMed]
  34. Lin, H.; Pearson, A.; Kazemi, Y.; Kakroodi, A.; Hammami, A.; Heydrich, M.; Xu, B.; Naguib, H.E. Influence of hygrothermal conditioning on the chemical structure and thermal mechanical properties of aliphatic polyketone. Polym. Degrad. Stabil. 2020, 179, 109260. [Google Scholar] [CrossRef]
  35. Morgen, T.O.; Baur, M.; Gottker-Schnetmann, I.; Mecking, S. Photodegradable branched polyethylenes from carbon monoxide copolymerization under benign conditions. Nat. Commun. 2020, 11, 3693. [Google Scholar] [CrossRef] [PubMed]
  36. Bovey, F.A.; Gooden, R.; Schilling, F.C.; Winslow, F.H. Carbon-13 NMR study of the solid-state photochemistry of poly(ethylene-co-carbon monoxide). Macromolecules 2002, 21, 938–944. [Google Scholar] [CrossRef]
  37. Xu, F.Y.; Chien, J.C.W. Photodegradation of alpha-olefin/carbon monoxide alternating copolymer. Macromolecules 2002, 26, 3485–3489. [Google Scholar] [CrossRef]
  38. Walter, R.; August, M. Production of Ketonic Bodies. U.S. Patent No. 2,577,208, 4 December 1951. [Google Scholar]
  39. Keim, W.; Maas, H. Copolymerization of ethylene and carbon monoxide by phosphinite-modified palladium catalysts. J. Organomet. Chem. 1996, 514, 271–276. [Google Scholar] [CrossRef]
  40. Bianchini, C.; Lee, H.M.; Meli, A.; Moneti, S.; Vizza, F.; Fontani, M.; Zanello, P. Copolymerization of carbon monoxide with ethene catalyzed by palladium(ii) complexes of 1,3-bis(diphenylphosphino)propane ligands bearing different substituents on the carbon backbone. Macromolecules 1999, 32, 4183–4193. [Google Scholar] [CrossRef]
  41. Doherty, S.; Eastham, G.R.; Tooze, R.P.; Scanlan, T.H.; Williams, D.; Elsegood, M.R.J.; Clegg, W. Palladium complexes of C2-, C3-, and C4-bridged bis(phospholyl) ligands:  remarkably active catalysts for the copolymerization of ethylene and carbon monoxide. Organometallics 1999, 18, 3558–3560. [Google Scholar] [CrossRef]
  42. Lindner, E.; Schmid, M.; Wegner, P.; Nachtigal, C.; Steimann, M.; Fawzi, R. Palladium(II) complexes with hemilabile etherdiphos ligands in the alternating copolymerization of carbon monoxide with olefins. Inorganica Chim. Acta. 1999, 296, 103–113. [Google Scholar] [CrossRef]
  43. Bianchini, C.; Lee, H.M.; Meli, A.; Oberhauser, W.; Vizza, F.; Brüggeller, P.; Rainer, H.B.; Langes, C. New structurally rigid palladium catalysts for the alternating copolymerization of carbon monoxide and ethene. Chem. Commun. 2000, 9, 777–778. [Google Scholar] [CrossRef]
  44. Verspui, G.; Schanssema, F.; Sheldon, R.A. A stable, conspicuously active, water-soluble pd catalyst for the alternating copolymerization of ethene and CO in water. Angew. Chem. Int. Ed. Engl. 2000, 39, 804–806. [Google Scholar] [CrossRef]
  45. Doherty, S.; Robins, E.G.; Knight, J.G.; Newman, C.R.; Rhodes, B.; Champkin, P.A.; Clegg, W. Selectivity for the methoxycarbonylation of ethylene versus CO ethylene copolymerization with catalysts based on C4-bridged bidentate phosphines and phospholes. J. Organomet. Chem. 2001, 640, 182–196. [Google Scholar] [CrossRef]
  46. Lu, C.C.; Peters, J.C. Catalytic copolymerization of CO and ethylene with a charge neutral palladium(II) zwitterion. J. Am. Chem. Soc. 2002, 124, 5272–5273. [Google Scholar] [CrossRef]
  47. Bianchini, C.; Meli, A.; Oberhauser, W.; Claver, C.; Garcia Suarez, E.J. Unraveling the o-methoxy effect in the CO/ethene copolymerization reaction by diphosphanepalladium(II) catalysis. Eur. J. Inorg. Chem. 2007, 2007, 2702–2710. [Google Scholar] [CrossRef]
  48. Drent, E.; Vanbroekhoven, J.A.M.; Doyle, M.J. Efficient palladium catalysts for the copolymerization of carbon-monoxide with olefins to produce perfectly alternating polyketones. J. Organomet. Chem. 1991, 417, 235–251. [Google Scholar] [CrossRef]
  49. Hyosung. Available online: https://www.hyosung.com/ (accessed on 28 December 2023.).
  50. Drent, E.; van Dijk, R.; van Ginkel, R.; van Oort, B.; Pugh, R.I. The first example of palladium catalysed non-perfectly alternating copolymerisation of ethene and carbon monoxide. Chem. Commun. 2002, 9, 964–965. [Google Scholar] [CrossRef] [PubMed]
  51. Baur, M.; Lin, F.; Morgen, T.O.; Odenwald, L.; Mecking, S. Polyethylene materials with in-chain ketones from nonalternating catalytic copolymerization. Science 2021, 374, 604–607. [Google Scholar] [CrossRef]
  52. Wang, C.; Xia, J.; Zhang, Y.; Hu, X.; Jian, Z. Photodegradable polar-functionalized polyethylenes. Natl. Sci. Rev. 2023, 10, nwad039. [Google Scholar] [CrossRef]
  53. Sommazzi, A.; Garbassi, F. Olefin-carbon monoxide copolymers. Prog. Polym. Sci. 1997, 22, 1547–1605. [Google Scholar] [CrossRef]
  54. Belov, G.P. Cationic Pd-II, Ni-II, and Ru-II complexes in the synthesis of alternating copolymers of CO with vinyl monomers. Russ. Chem. Bull. 2002, 51, 1605–1615. [Google Scholar] [CrossRef]
  55. Bianchini, C.; Meli, A.; Oberhauser, W. Catalyst design and mechanistic aspects of the alternating copolymerisation of ethene and carbon monoxide by diphosphine-modified palladium catalysis. Dalton. Trans. 2003, 13, 2627–2635. [Google Scholar] [CrossRef]
  56. García Suárez, E.J.; Godard, C.; Ruiz, A.; Claver, C. Alternating and non-alternating Pd-catalysed co-and terpolymerisation of carbon monoxide and alkenes. Eur. J. Inorg. Chem. 2007, 2007, 2582–2593. [Google Scholar] [CrossRef]
  57. Cavinato, G.; Toniolo, L. Carbonylation of ethene catalysed by Pd(II)-phosphine complexes. Molecules 2014, 19, 15116–15161. [Google Scholar] [CrossRef] [PubMed]
  58. Chen, S.Y.; Pan, R.C.; Chen, M.; Liu, Y.; Chen, C.; Lu, X.B. Synthesis of nonalternating polyketones using cationic diphosphazane monoxide-palladium complexes. J. Am. Chem. Soc. 2021, 143, 10743–10750. [Google Scholar] [CrossRef]
  59. Luo, R.; Newsham, D.K.; Sen, A. Palladium-catalyzed nonalternating copolymerization of ethene and carbon monoxide: Scope and mechanism. Organometallics 2009, 28, 6994–7000. [Google Scholar] [CrossRef]
  60. Tang, S.; Seidel, F.W.; Nozaki, K. High density polyethylenes bearing isolated in-chain carbonyls. Angew. Chem. Int. Ed. Engl. 2021, 60, 26506–26510. [Google Scholar] [CrossRef]
  61. Chen, S.-Y.; Song, Y.-H.; Jiao, S.; Zou, C.; Li, S.-H.; Chen, C.; Lu, X.-B.; Liu, Y. Carbonyl functionalized polyethylene materials via Ni- and Pd-diphosphazane monoxide catalyzed nonalternating copolymerization. J. Catal. 2023, 417, 334–340. [Google Scholar] [CrossRef]
  62. Yonezaki, G.; Seidel, F.W.; Takahashi, K.; Nozaki, K. Nickel-catalyzed selective incorporation of isolated in-chain carbonyls into ethylene/carbon monoxide copolymer using metal carbonyls as a carbonyl source. Bull. Chem. Soc. Jpn. 2023, 96, 545–549. [Google Scholar] [CrossRef]
  63. Li, S.H.; Pan, R.C.; Ren, B.H.; Yang, J.W.; Kang, X.; Liu, Y. Cationic palladium catalyzed nonalternating copolymerization of ethylene with carbon monoxide: Microstructure analysis and computational study. Chin. J. Chem. 2022, 41, 417–423. [Google Scholar] [CrossRef]
  64. Voccia, M.; Odenwald, L.; Baur, M.; Lin, F.; Falivene, L.; Mecking, S.; Caporaso, L. Mechanistic insights into Ni(ii)-catalyzed nonalternating ethylene-carbon monoxide copolymerization. J. Am. Chem. Soc. 2022, 144, 15111–15117. [Google Scholar] [CrossRef]
  65. Soomro, S.S.; Cozzula, D.; Leitner, W.; Vogt, H.; Müller, T.E. The microstructure and melt properties of CO–ethylene copolymers with remarkably low CO content. Polym. Chem. 2014, 5, 3831–3837. [Google Scholar] [CrossRef]
  66. Ortmann, P.; Wimmer, F.P.; Mecking, S. Long-spaced polyketones from ADMET copolymerizations as ideal models for ethylene/CO copolymers. ACS Macro. Lett. 2015, 4, 704–707. [Google Scholar] [CrossRef]
  67. Hearley, A.K.; Nowack, R.A.J.; Rieger, B. New single-site palladium catalysts for the nonalternating copolymerization of ethylene and carbon monoxide. Organometallics 2005, 24, 2755–2763. [Google Scholar] [CrossRef]
  68. Bettucci, L.; Bianchini, C.; Claver, C.; Suarez, E.J.; Ruiz, A.; Meli, A.; Oberhauser, W. Ligand effects in the non-alternating CO-ethylene copolymerization by palladium(II) catalysis. Dalton. Trans. 2007, 47, 5590–5602. [Google Scholar] [CrossRef] [PubMed]
  69. Newsham, D.K.; Borkar, S.; Sen, A.; Conner, D.M.; Goodall, B.L. Inhibitory role of carbon monoxide in palladium(II)-catalyzed nonalternating ethene/carbon monoxide copolymerizations and the synthesis of polyethene-poly(ethene-alt-carbon monoxide). Organometallics 2007, 26, 3636–3638. [Google Scholar] [CrossRef]
  70. Haras, A.; Michalak, A.; Rieger, B.; Ziegler, T. Theoretical analysis of factors controlling the nonalternating CO/C2H4 copolymerization. J. Am. Chem. Soc. 2005, 127, 8765–8774. [Google Scholar] [CrossRef]
  71. Haras, A.; Michalak, A.; Rieger, B.; Ziegler, T. Comparative study on catalytic systems for the alternating and nonalternating CO/ethene copolymerization. Organometallics 2006, 25, 946–953. [Google Scholar] [CrossRef]
  72. Conley, M.P.; Jordan, R.F. cis/trans isomerization of phosphinesulfonate palladium(II) complexes. Angew. Chem. Int. Ed. Engl. 2011, 50, 3744–3746. [Google Scholar] [CrossRef] [PubMed]
  73. Nakamura, A.; Anselment, T.M.; Claverie, J.; Goodall, B.; Jordan, R.F.; Mecking, S.; Rieger, B.; Sen, A.; van Leeuwen, P.W.; Nozaki, K. Ortho-phosphinobenzenesulfonate: A superb ligand for palladium-catalyzed coordination-insertion copolymerization of polar vinyl monomers. Acc. Chem. Res. 2013, 46, 1438–1449. [Google Scholar] [CrossRef] [PubMed]
  74. Nakamura, A.; Munakata, K.; Ito, S.; Kochi, T.; Chung, L.W.; Morokuma, K.; Nozaki, K. Pd-catalyzed copolymerization of methyl acrylate with carbon monoxide: Structures, properties and mechanistic aspects toward ligand design. J. Am. Chem. Soc. 2011, 133, 6761–6779. [Google Scholar] [CrossRef] [PubMed]
  75. Noda, S.; Nakamura, A.; Kochi, T.; Chung, L.W.; Morokuma, K.; Nozaki, K. Mechanistic studies on the formation of linear polyethylene chain catalyzed by palladium phosphine-sulfonate complexes: Experiment and theoretical studies. J. Am. Chem. Soc. 2009, 131, 14088–14100. [Google Scholar] [CrossRef] [PubMed]
  76. Zuidema, E.; Bo, C.; van Leeuwen, P.W. Ester versus polyketone formation in the palladium-diphosphine catalyzed carbonylation of ethene. J. Am. Chem. Soc. 2007, 129, 3989–4000. [Google Scholar] [CrossRef] [PubMed]
  77. Bianchini, C.; Meli, A.; Oberhauser, W.; Segarra, A.M.; Passaglia, E.; Lamac, M.; Stepnicka, P. Palladium(II) complexes with phosphanylferrocenecarboxylate ligands and their use as catalyst precursors for semialternating CO-ethylene copolymerization. Eur. J. Inorg. Chem. 2008, 2008, 441–452. [Google Scholar] [CrossRef]
  78. Zhu, L.H.; Gaire, S.; Ziegler, C.J.; Jia, L. Nickel catalysts for non-alternating CO-ethylene copolymerization. ChemCatChem 2022, 14, e202200974. [Google Scholar] [CrossRef]
  79. Drent, E.; van Dijk, R.; van Ginkel, R.; van Oort, B.; Pugh, R.I. Palladium catalysed copolymerisation of ethene with alkylacrylates: Polar comonomer built into the linear polymer chain. Chem. Commun. 2002, 7, 744–745. [Google Scholar] [CrossRef]
  80. Chen, C.; Anselment, T.M.J.; Fröhlich, R.; Rieger, B.; Kehr, G.; Erker, G. o-Diarylphosphinoferrocene sulfonate palladium systems for nonalternating ethene–carbon monoxide copolymerization. Organometallics 2011, 30, 5248–5257. [Google Scholar] [CrossRef]
  81. Dodge, H.M.; Natinsky, B.S.; Jolly, B.J.; Zhang, H.C.; Mu, Y.; Chapp, S.M.; Tran, T.V.; Diaconescu, P.L.; Do, L.H.; Wang, D.W.; et al. Polyketones from carbon dioxide and ethylene by integrating electrochemical and organometallic catalysis. ACS Catal. 2023, 13, 4053–4059. [Google Scholar] [CrossRef]
  82. Chen, M.; Chen, C. A versatile ligand platform for palladium- and nickel-catalyzed ethylene copolymerization with polar monomers. Angew. Chem. Int. Ed. Engl. 2018, 57, 3094–3098. [Google Scholar] [CrossRef]
  83. Xu, M.; Yu, F.; Li, P.; Xu, G.; Zhang, S.; Wang, F. Enhancing chain initiation efficiency in the cationic allyl-nickel catalyzed (co)polymerization of ethylene and methyl acrylate. Inorg. Chem. 2020, 59, 4475–4482. [Google Scholar] [CrossRef] [PubMed]
  84. Zou, C.; Liao, D.H.; Pang, W.M.; Chen, M.; Tan, C. Versatile PNPO ligands for palladium and nickel catalyzed ethylene polymerization and copolymerization with polar monomers. J. Catal. 2021, 393, 281–289. [Google Scholar] [CrossRef]
  85. Klabunde, U.; Tulip, T.H.; Roe, D.C.; Ittel, S.D. Reaction of nickel polymerization catalysts with carbon monoxide. J. Organomet. Chem. 1987, 334, 141–156. [Google Scholar] [CrossRef]
  86. Desjardins, S.Y.; Cavell, K.J.; Hoare, J.L.; Skelton, B.W.; Sobolev, A.N.; White, A.H.; Keim, W. Single component N-O chelated arylnickel(II) complexes as ethene polymerisation and CO/ethene copolymerisation catalysts. Examples of ligand induced changes to the reaction pathway. J. Organomet. Chem. 1997, 544, 163–174. [Google Scholar] [CrossRef]
  87. Klaui, W.; Bongards, J.; Reiß, G.J. Novel nickel(ii) complexes for the catalytic copolymerization of ethylene and carbon monoxide: Polyketone synthesis in supercritical carbon dioxide. Angew. Chem. Int. Ed. Engl. 2000, 39, 3894–3896. [Google Scholar] [CrossRef] [PubMed]
  88. Jia, X.; Zhang, M.; Pan, F.; Babahan, I.; Ding, K.; Jia, L.; Crandall, L.A.; Ziegler, C.J. Zwitterionic nickel(ii) catalyst for CO–ethylene alternating copolymerization. Organometallics 2015, 34, 4798–4801. [Google Scholar] [CrossRef]
  89. Chen, S.Y.; Ren, B.H.; Li, S.H.; Song, Y.H.; Jiao, S.; Zou, C.; Chen, C.; Lu, X.B.; Liu, Y. Cationic P,O-coordinated nickel(ii) catalysts for carbonylative polymerization of ethylene: Unexpected productivity via subtle electronic variation. Angew. Chem. Int. Ed. Engl. 2022, 61, e202204126. [Google Scholar] [CrossRef]
  90. Xin, B.S.; Sato, N.; Tanna, A.; Oishi, Y.; Konishi, Y.; Shimizu, F. Nickel catalyzed copolymerization of ethylene and alkyl acrylates. J. Am. Chem. Soc. 2017, 139, 3611–3614. [Google Scholar] [CrossRef]
  91. Zhang, Y.; Mu, H.; Pan, L.; Wang, X.; Li, Y. Robust bulky [P,O] neutral nickel catalysts for copolymerization of ethylene with polar vinyl monomers. ACS Catal. 2018, 8, 5963–5976. [Google Scholar] [CrossRef]
Figure 1. Three microstructures of CO/ethylene copolymer using different Ni/Pd catalysts. (A): references [47,48]; (B): reference [50]; (C): reference [51].
Figure 1. Three microstructures of CO/ethylene copolymer using different Ni/Pd catalysts. (A): references [47,48]; (B): reference [50]; (C): reference [51].
Ijms 25 01348 g001
Scheme 1. Presented mechanisms in the formation of alternating PKs and PKs without a perfect alternating structure catalyzed by Pd complexes.
Scheme 1. Presented mechanisms in the formation of alternating PKs and PKs without a perfect alternating structure catalyzed by Pd complexes.
Ijms 25 01348 sch001
Scheme 2. Non-alternating pathway in CO/ethylene non-alternating copolymerization.
Scheme 2. Non-alternating pathway in CO/ethylene non-alternating copolymerization.
Ijms 25 01348 sch002
Figure 2. Plausible orbital interactions in phosphine–sulfonate Pd catalyst systems.
Figure 2. Plausible orbital interactions in phosphine–sulfonate Pd catalyst systems.
Ijms 25 01348 g002
Figure 3. Plausible orbital interactions to stabilize the trans π-complex.
Figure 3. Plausible orbital interactions to stabilize the trans π-complex.
Ijms 25 01348 g003
Figure 4. [P,O] palladium catalyst systems generated in situ from Pd(OAc)2 and phosphine–sulfonate ligands.
Figure 4. [P,O] palladium catalyst systems generated in situ from Pd(OAc)2 and phosphine–sulfonate ligands.
Ijms 25 01348 g004
Figure 5. Phosphine–sulfonate palladium catalysts.
Figure 5. Phosphine–sulfonate palladium catalysts.
Ijms 25 01348 g005
Figure 6. Phosphine–sulfonate palladium catalysts with modification of backbones.
Figure 6. Phosphine–sulfonate palladium catalysts with modification of backbones.
Ijms 25 01348 g006
Figure 7. Ferrocene-derived phosphine–sulfonate palladium catalysts.
Figure 7. Ferrocene-derived phosphine–sulfonate palladium catalysts.
Ijms 25 01348 g007
Figure 8. Phosphine–sulfonate palladium catalysts with various phosphine substituents (R).
Figure 8. Phosphine–sulfonate palladium catalysts with various phosphine substituents (R).
Ijms 25 01348 g008
Figure 9. PNPO Ni/Pd catalysts for CO/ethylene non-alternating copolymerization.
Figure 9. PNPO Ni/Pd catalysts for CO/ethylene non-alternating copolymerization.
Ijms 25 01348 g009
Figure 10. Modifiable sites on the PNPO ligands.
Figure 10. Modifiable sites on the PNPO ligands.
Ijms 25 01348 g010
Figure 11. Phosphinophenolate Ni catalysts for CO/ethylene non-alternating copolymerization.
Figure 11. Phosphinophenolate Ni catalysts for CO/ethylene non-alternating copolymerization.
Ijms 25 01348 g011
Figure 12. Competition and rate-determining step in Ni catalyzed synthesis of PE with in-chain ketones.
Figure 12. Competition and rate-determining step in Ni catalyzed synthesis of PE with in-chain ketones.
Ijms 25 01348 g012
Table 1. Microstructure and Tm of three kinds of PKs.
Table 1. Microstructure and Tm of three kinds of PKs.
MicrostructurePKsRange of mCO Content (%)Tm (°C)
Ijms 25 01348 i001Alternating PKsm = 150~260
Non-perfect alternating PKsm > 110~50<260
PE with in-chain ketonesm > 10~10<140
Table 2. Structural fragments and NMR shifts of PKs without a perfect alternating structure.
Table 2. Structural fragments and NMR shifts of PKs without a perfect alternating structure.
Structure UnitFragmentsNMR Shifts
n1H13C
Alternating unit (AU)Ijms 25 01348 i0021/212.9 a/207.0 b
2~2.4 c35.7 c
Non-alternating unit (N2)Ijms 25 01348 i0033/209.8 b
4~2.2 c42.1 c
5~1.5 c23.0 c
Non-alternating unit (N3)Ijms 25 01348 i0046/210.1 b
7~2.2 c42.5 c
8~1.5 c23.7 c
9~1.3 c28.8 c
Non-alternating unit (N4)Ijms 25 01348 i00510/210.3 b
11~2.2 c42.7 c
12~1.5 c24.0 c
13~1.3 c29.2 c
Isolated carbonyl unit (I)Ijms 25 01348 i00614/210.3 b
15~2.4 b42.7 b
16~1.6 b23.8 b
a Alternating PKs in mixture of hexafluoroisopropanol (HFIP, 90%) and C6D6 (10%), reference [48]. b PE with in-chain ketones from 13CO/ethylene copolymerization, in C2D2Cl4, 383K, reference [51]. c Non-perfect alternating PKs in HFIP, reference [58].
Table 3. CO/ethylene non-alternating copolymerization results using phosphine–sulfonate palladium catalysts.
Table 3. CO/ethylene non-alternating copolymerization results using phosphine–sulfonate palladium catalysts.
Cat.Add.Temp.
(°C)
Press.
(Bar)
CO/Thylene FeedAct. aExtra
C2 Insertion b
(Mol %)
Tm cRef
Pd1/110502/31237.3234[50]
Pd2/110502/310311.9231[50]
Pd3/110502/310818.3229[50]
Pd4/110401/3<11.4220~230[67]
Pd5/1101101/101218.3220~230[67]
Pd5/110401/3692.4220~230[67]
Pd6/110401/3478.2220~230[67]
Pd7/110401/32.5<1220~230[67]
Pd8/110401/31064.1220~230[67]
Pd9/110401/31022.1220~230[67]
Pd9/200701/6754.9220~230[67]
Pd10B(C6F5)3 e1101402/52923.1220~230[67]
Pd11B(C6F5)3 e1101402/5672.2220~230[67]
Pd11B(C6F5)3 e1101051/208926.4220~230[67]
Pd12/120561/204127.8193[68]
Pd12BQ e120561/36072.9/d[68]
Pd12/120561/33373.4235[68]
Pd13/120561/31705.6/d[68]
Pd14/110561/30.40.3/d[68]
Pd15/110561/30.40.3/d[68]
Pd16/90551/101812199[80]
Pd16/110651/122125171[80]
Pd17/110502/3751242[80]
Pd18/110502/3743233[80]
Pd18B(C6F5)3 e110502/31591242[80]
Pd18/110551/101827167[80]
a Activity: in units of g PK mmol·Pd−1·h−1. b Total mol % ethylene of copolymer involved in non-alternation, calculated from 13C NMR spectra. c Determined by DSC. d Not determined. e B(C6F5)3 is used as the activator, BQ: benzoquinone as the oxidant.
Table 4. CO/ethylene non-alternating copolymerizations using PNPO Pd/Ni catalysts.
Table 4. CO/ethylene non-alternating copolymerizations using PNPO Pd/Ni catalysts.
Cat.CO/Ethylene FeedAct. aCarbonyl Content (Extra C2 Insertion) b (Mol %)Mn  cPDI cTm  dRef
Pd271/915.8494031.46245[58]
Pd281/919.7494431.37245[58]
Pd291/928.1497701.37241[58]
Pd301/938.3478351.28232[58]
Pd311/940.9479041.16231[58]
Pd321/951.3469481.23227[58]
Pd331/931.4485691.39243[58]
Pd341/935.6486311.33242[58]
Pd351/936.9476541.33238[58]
Pd361/926.2485421.32246[58]
Pd371/928.4485921.42246[58]
Pd381/970.643 (17.7)9151.09186[58]
Pd391/90/e/e/e/e[58]
Pd401/925.5497261.20244[58]
Pd381/1124.050 (1.1)13401.08248[58]
Pd381/3105.047 (4.7)12001.08226[58]
Pd381/580.145 (11.9)11101.14214[58]
Pd381/2054.739 (24.2)10701.14165/147[58]
Pd411/970.148 (2.1)1421.61240[63]
Pd421/969.3471341.76237[63]
Pd431/960.7471261.61237[63]
Pd321/932.546 (4.9)1181.34229[63]
Pd441/9124.147 (3.3)1991.43235[63]
Pd451/925.044 (12.2)671.35196[63]
Pd461/923.344 (11.6)611.27194[63]
Pd471/922.646 (6.7)671.21209[63]
Pd481/959.444 (12.1)1491.45198[63]
Pd491/967.845 (10.1)1511.47206[63]
Pd501/924.146 (6.5)641.37209[63]
Ni10.2%661.18.942.64124[61]
Ni20.2%870.58.822.44122[61]
Ni30.2%910.84.402.95121[61]
Ni40.2%421.13.082.32109[61]
Ni50.2%0/e/e/e/e[61]
Ni60.2%2630.217.82.31124[61]
Ni70.2%3010.55.482.80118[61]
Ni80.2%0/e/e/e/e[61]
Ni61.0%390.224.02.25126[61]
Pd290.2%240.93.332.75125[61]
Pd300.2%330.84.082.33126[61]
Pd310.2%360.97.912.59126[61]
Pd320.2%440.69.562.70126[61]
Pd380.2%680.840.42.62126[61]
Pd380.1%700.593.31.80133[61]
a Activity: in units of g PK (mmol·[M])−1·h−1. b Total mol % ethylene of copolymer involved in non-alternation, calculated from 13C{1H} NMR spectra; carbonyl content calculated from 1H NMR spectroscopy. c Mn in units of kg·mol−1, determined by GPC. d Determined by DSC. e Not determined.
Table 5. CO/ethylene non-alternating copolymerization results using phosphinophenolate Ni catalysts.
Table 5. CO/ethylene non-alternating copolymerization results using phosphinophenolate Ni catalysts.
Cat.CO/Ethylene Feed a (%)Act. bCarbonyl Content c (Mol %)IC/NA/AC dMn ePDI eTm fRef
Ni90.2213.10.3 (0.3)79/21/01401.6134[51]
Ni90.698.80.8 (1.1)32/36/321471.8135[51]
Ni100.656.61.0 (1.4)21/29/49601.7133[51]
Ni110.612.93.7 (4.4)28/34/38841.7134[51]
Ni120.622.11.6 (3.0)7/34/59451.6133[51]
Ni130.641.21.1 (2.5)18/33/49901.7135[51]
Ni140.60.6500/0/100/g/g/g[51]
a 13CO was used to enhance sensitivity. b Activity: in units of g PK·(mmol·Ni)−1·h−1. c Carbonyl content determined by IR spectroscopy; in brackets: carbonyl content calculated from 13C NMR spectroscopy by integration of the 13C labeled C=O signals. d IC: isolated carbonyl; NA: non-alternating carbonyl; AC: alternating carbonyl; see Table 2 for specification. e Mn in units of kg·mol−1, determined by GPC in 1,2-dichlorobenzene at 160 °C. f Determined by DSC. g Not determined.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xiao, X.; Zheng, H.; Gao, H.; Cheng, Z.; Feng, C.; Yang, J.; Gao, H. Recent Advances in Synthesis of Non-Alternating Polyketone Generated by Copolymerization of Carbon Monoxide and Ethylene. Int. J. Mol. Sci. 2024, 25, 1348. https://doi.org/10.3390/ijms25021348

AMA Style

Xiao X, Zheng H, Gao H, Cheng Z, Feng C, Yang J, Gao H. Recent Advances in Synthesis of Non-Alternating Polyketone Generated by Copolymerization of Carbon Monoxide and Ethylene. International Journal of Molecular Sciences. 2024; 25(2):1348. https://doi.org/10.3390/ijms25021348

Chicago/Turabian Style

Xiao, Xieyi, Handou Zheng, Heng Gao, Zhaocong Cheng, Chunyu Feng, Jiahao Yang, and Haiyang Gao. 2024. "Recent Advances in Synthesis of Non-Alternating Polyketone Generated by Copolymerization of Carbon Monoxide and Ethylene" International Journal of Molecular Sciences 25, no. 2: 1348. https://doi.org/10.3390/ijms25021348

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop