Fact-checked by Grok 2 weeks ago

Coordination polymerization

Coordination polymerization is a form of in which monomers, typically olefins or dienes, coordinate to a on a before undergoing migratory insertion into the growing , enabling precise control over polymer microstructure such as stereoregularity, linearity, and molecular weight distribution. This process contrasts with free-radical by producing highly ordered polymers without branching under mild conditions, often at low pressures and temperatures. The discovery of coordination polymerization is credited to , who in 1953 demonstrated that a mixture of triethylaluminum and catalyzes the low-pressure of to linear (HDPE), revolutionizing industrial production from high-pressure processes. In 1954, extended this to propylene, achieving stereoregular isotactic using similar catalysts, which earned them the in 1963 for their foundational work on stereospecific . Early catalysts were heterogeneous Ziegler-Natta systems, consisting of a halide (e.g., TiCl₄) reduced by an organoaluminum compound (e.g., AlEt₃) and often supported on for enhanced activity and selectivity. Subsequent advancements include homogeneous single-site catalysts, such as metallocenes (e.g., zirconocene dichloride with methylaluminoxane cocatalyst), developed in the by Hansjörg Sinn, Walter Kaminsky, and others, which provide even greater uniformity in polymer properties and enable the synthesis of syndiotactic or atactic variants. The mechanism generally involves coordination of the to the metal center, followed by insertion into the metal-carbon bond of the chain, propagation through repeated insertions, and termination via β-hydride elimination or agents like hydrogen. These systems are industrially dominant, producing over 100 million metric tons annually of polyolefins including HDPE, (LLDPE), , and polydienes for applications in packaging, automotive parts, and elastomers.

Introduction

Definition and Scope

Coordination polymerization is a process that employs catalysts, where monomers coordinate to the metal center prior to their insertion into the growing , enabling the formation of stereoregular polymers with controlled microstructures. This mechanism allows for precise control over the 's , as the coordination step orients the in a specific relative to the . Unlike , which proceeds via free radical intermediates leading to less controlled chain propagation, or anionic and cationic methods that rely on ionic initiators for propagation, coordination polymerization offers site-specific control through metal-monomer interactions, resulting in higher and reduced side reactions. Ziegler-Natta catalysts serve as classic examples of systems enabling this controlled addition. The scope of coordination polymerization primarily encompasses the production of polyolefins from olefins such as and , yielding materials like (HDPE), (LLDPE), and isotactic , which are widely used in , pipes, and automotive parts due to their mechanical strength and thermal properties. It also extends to dienes like for synthetic rubbers and, to a lesser extent, certain polar monomers, though challenges with monomer coordination persist for highly polar . Key characteristics include the ability to achieve high molecular weight polymers with tunable chain lengths, exceptional for isotactic or syndiotactic structures, and narrow polydispersity indices, particularly with single-site catalysts that minimize site heterogeneity. These features arise from the catalytic site's uniformity, allowing for living-like behaviors in optimized systems.

Historical Background

The discovery of coordination polymerization began in 1953 when observed the unexpected low-pressure polymerization of into using a catalyst system composed of (TiCl₄) and triethylaluminum (AlEt₃). In 1954, extended this approach by applying similar transition metal-based catalysts to , achieving the synthesis of stereoregular with controlled , which enabled the production of crystalline polymers with enhanced mechanical properties. These breakthroughs marked the foundation of coordination polymerization, shifting from high-pressure free radical methods to controlled catalytic processes at ambient conditions. For their pioneering contributions to the chemistry and technology of high polymers, and Natta were jointly awarded the in 1963. During the 1970s and 1980s, the field advanced toward industrial scalability through the development of supported Ziegler-Natta catalysts, particularly third-generation systems using (MgCl₂) as a support for species, which improved activity, , and catalyst recovery in large-scale reactors. In the mid-1980s, Kaminsky and colleagues introduced metallocene catalysts, such as zirconocene complexes activated by methylaluminoxane (MAO), providing single-site active centers for precise control over microstructure and molecular weight distribution. These innovations facilitated the commercial production of tailored polyolefins, with the first (HDPE) plant operational in 1955 by Hoechst, rapidly expanding to revolutionize the by enabling cost-effective, high-volume of materials like HDPE and isotactic . In the post-2000 era, coordination polymerization evolved with the rise of late catalysts, such as - and palladium-based systems, which tolerate polar monomers like acrylates and enable copolymerization to produce functionalized polyolefins with improved properties for applications in adhesives and packaging. These catalysts, building on earlier diimine designs, address limitations of early systems by reducing sensitivity to impurities and enhancing incorporation. Into the 2020s, advancements have included catalyst systems enabling efficient copolymerization of olefins with polar and bio-based monomers, such as , to promote sustainable production of functionalized polymers.

Fundamental Principles

Coordination Mechanism

In coordination polymerization, the mechanism begins with the coordination of a molecule to a vacant site on the center, typically through π-bonding of the unsaturated moiety. This step forms a transient π-complex, positioning the for subsequent reaction. The metal center, often featuring an alkyl or growing bound to it, must possess sufficient coordinative unsaturation—commonly achieved in d-block elements like or —for this coordination to occur effectively. Following coordination, the undergoes migratory insertion into the metal-carbon σ-bond of the initiating or propagating , transferring the to the 's former position while regenerating a vacant site for further propagation. This insertion involves minimal nuclear displacement and is driven by electronic reorganization at the metal center, as originally proposed in the Cossee-Arlman model. The phase consists of repeated coordination-insertion cycles, enabling rapid growth. A simplified representation of the propagation step is: \ce{M-R + CH2=CHR' -> M-CH2-CHR'-R} Here, M represents the center, R the growing , and CH₂=CHR' the olefin . Each insertion adds one unit, with the regiochemistry (1,2- vs. 2,1-) and determined by the orientation of the coordinated relative to the metal-alkyl . In heterogeneous systems, such as Ziegler-Natta catalysts, site epimerization can occur between insertions, allowing the to reset for subsequent addition. Stereocontrol in this arises from two primary modes: chain-end control, where the of the terminal unit on the growing chain directs the facial selectivity of the incoming , or enantiomorphic site control, in which the asymmetric coordination environment of the metal site imposes on the insertion, favoring isotactic or syndiotactic placement. The Cossee-Arlman particularly underpins enantiomorphic site control in heterogeneous , where the fixed geometry of the catalytic site influences approach and minimizes epimerization errors. Chain termination typically proceeds via β-hydride elimination, in which a hydrogen from the β-carbon of the polymer chain migrates to the metal, yielding a metal species and an α-olefin-terminated polymer with a vinylidene end group. Alternatively, to or to another metal alkyl can occur, producing polymers with saturated or unsaturated end groups depending on the pathway; these processes are reversible under certain conditions but limit molecular weight when dominant. β- elimination is particularly prevalent in systems with accessible β-hydrogens and low concentrations. The ligand environment profoundly impacts the overall mechanism by dictating the metal's , which governs vacant availability and monomer orientation, and by modulating insertion barriers through steric hindrance and electronic effects. For example, bulky or electron-donating can stabilize the for insertion, lowering the energy barrier and enhancing propagation rates, while also influencing by restricting monomer access.

Catalyst Components and Activation

Coordination polymerization catalysts typically consist of a compound as the primary component, which provides the for coordination and insertion. Common compounds include -based halides such as TiCl₄ for heterogeneous systems and metallocene complexes like Cp₂ZrCl₂ for homogeneous ones. These compounds are chosen for their ability to form low-valent metal centers that facilitate olefin binding, with (IV) chlorides being particularly effective due to their high activity in producing linear polyolefins. Cocatalysts, often organoaluminum compounds like triethylaluminum (AlEt₃, also known as ), are essential for activating the precursor. These cocatalysts perform multiple roles, including the reduction of the metal center to a lower and the provision of alkyl groups for initiating polymerization. Additionally, they act as to remove impurities such as water or oxygen, which would otherwise deactivate by the active sites. In heterogeneous systems, supports like MgCl₂ are employed to disperse the compound, enhancing surface area and catalytic stability. The support not only prevents aggregation of active sites but also influences the electronic environment of the metal, promoting higher activity and selectivity. additives, such as ethyl benzoate, are incorporated during catalyst preparation to modify the , improving by selectively non-stereospecific sites. The activation process begins with the interaction between the compound and cocatalyst, primarily through of the metal center to generate an active M-alkyl . For example, in a classic Ziegler-Natta system, reacts with AlEt₃ to form an ethylated : \text{TiCl}_4 + \text{AlEt}_3 \rightarrow \text{TiCl}_3\text{-Et} + \text{AlEt}_2\text{Cl} This simplified equation illustrates the transfer of an to , creating a vacant coordination site while the aluminum byproduct helps scavenge poisons. The process reduces Ti(IV) to Ti(III), enabling the formation of multiple active sites per catalyst particle. Catalyst variations, such as single-site metallocenes versus multi-site heterogeneous Ziegler-Natta systems, significantly impact uniformity. Single-site s, activated by methylaluminoxane (MAO) instead of simple alkylaluminums, yield more uniform molecular weight distributions due to fewer distinct types, whereas multi-site heterogeneous s produce broader distributions but higher overall productivity.

Polymerization of Olefins

Heterogeneous Ziegler-Natta Polymerization

Heterogeneous Ziegler-Natta polymerization employs supported catalysts, primarily MgCl₂-based systems with active centers, to produce polyolefins such as and on an industrial scale. The catalyst preparation typically involves the synthesis of a MgCl₂ support with high surface area, often achieved through reactive or ball-milling methods, followed by impregnation or deposition of TiCl₄ or TiCl₃ as the component. Internal electron donors, such as alkyl (e.g., ), are incorporated during support formation to enhance by coordinating to Ti sites and blocking non-specific adsorption, while external donors like silanes (e.g., dicyclopentyldimethoxysilane) are added later with the cocatalyst to further regulate isotacticity. These multi-component systems create heterogeneous active centers distributed across the support's porous structure, enabling high catalyst productivity but leading to site heterogeneity. Polymerization occurs under slurry or gas-phase conditions, with typical temperatures ranging from 50 to 100°C and pressures of 1 to 50 , depending on the and process. In slurry processes, such as those using or as , the reaction proceeds in a loop reactor at around 70°C and 30-40 for , while gas-phase fluidized-bed reactors operate at similar temperatures but lower pressures (20-30 ) for ethylene- copolymers. Trialkylaluminum cocatalysts (e.g., triethylaluminum) alkylate the Ti centers, and hydrogen is introduced as a agent to control molecular weight, typically at partial pressures of 0.1-0.5 to achieve desired polymer viscosities without excessive branching. These conditions promote rapid insertion via a coordination-insertion at multiple active sites, resulting in high yields exceeding 10 kg per gram catalyst. The resulting polymers exhibit broad molecular weight distributions, with polydispersity indices (PDI) often greater than 5 due to the multiplicity of active sites with varying and rates. For isotactic , this heterogeneity yields high crystallinity (typically 40-60%), contributing to superior mechanical strength and stiffness compared to atactic forms. Particle is replicated from , producing spherical products with (100-500 μm), which facilitates like . Industrially, heterogeneous Ziegler-Natta systems account for the majority (approximately 65%) of global , around 90 million tons as of 2024. The Spheripol process, a slurry-loop licensed by , exemplifies this, combining bulk with gas-phase copolymerization for impact-modified grades, achieving high throughput and low residue catalysts. These catalysts offer cost-effectiveness through low metal loadings and scalability, with activities up to 50 kg PP/g cat/h, though they provide less precise control than single-site alternatives, often requiring donor optimization for >95% isotacticity.

Homogeneous Ziegler-Natta Polymerization

Homogeneous Ziegler-Natta polymerization employs soluble compounds, primarily or halides, in combination with aluminum alkyl cocatalysts to facilitate the coordination polymerization of olefins in solution. Typical catalyst systems include soluble compounds such as TiCl₄ paired with AlEt₃, or vanadium-based variants like VCl₄ with Al(i-Bu)₃, which enable the of monomers through alkyl and to lower oxidation states. These systems operate without solid supports, allowing all components to dissolve in solvents, which contrasts with their heterogeneous counterparts by providing greater homogeneity in distribution. The process typically occurs in solution at temperatures ranging from 100 to 150°C, where monomers like and are polymerized under moderate pressure. This setup facilitates easier product-catalyst separation compared to heterogeneous systems, as the soluble nature avoids filtration challenges, though catalyst activities are generally lower, often by orders of , due to limited site stability. For instance, VCl₄/Al(i-Bu)₃ systems have been pivotal in copolymerizing and , yielding elastomeric products with enhanced properties. The resulting polymers exhibit superior comonomer incorporation, particularly for producing (LLDPE) with uniform α-olefin distribution, and narrower polydispersity indices (PDI) typically in the range of 2-4, reflecting more uniform chain growth than the broader distributions (PDI > 5) from heterogeneous catalysts. Historically, these homogeneous systems gained prominence in the for synthesizing ethylene-propylene rubbers (), where vanadium catalysts like enabled the production of amorphous copolymers with high molecular weights suitable for elastomeric applications. Although largely supplanted by metallocene catalysts in the and 1990s for their higher activity and precision, homogeneous Ziegler-Natta systems persist in niche uses for specialty elastomers, such as ethylene-propylene-diene terpolymers, where their ability to incorporate dienes without stereoregularity constraints is advantageous. Key challenges include the need for catalyst residue removal, as the soluble components can contaminate the , necessitating post-polymerization purification steps like or solvent extraction, unlike high-activity heterogeneous systems where residues are minimal. Additionally, thermal stability remains an issue, with active prone to deactivation through to inactive low-valent states or at elevated temperatures, limiting operational windows and overall productivity.

Polymerization of Dienes and Specialty Monomers

Butadiene Polymerization

Coordination polymerization of , a conjugated , extends the principles of olefin insertion by allowing multiple regioselective additions (1,2 or 1,4) while enabling precise stereocontrol through design. This process typically employs salts, such as those of or , combined with organoaluminum cocatalysts and acids, to produce stereoregular polybutadienes with tailored microstructures. Cobalt-based catalysts, exemplified by CoCl₂ or Co(acac)₂ with AlEt₂Cl and water or halides, predominantly yield high cis-1,4-polybutadiene (up to 98% cis-1,4 content) by coordinating the diene in an s-cis conformation, favoring syn-anti-syn addition during monomer insertion. salts, such as Ni(octanoate)₂ or Ni(naphthenate)₂, activated by AlEt₃ and BF₃·OEt₂, are particularly effective for syndiotactic 1,2-polybutadiene (up to 95% syndiotactic 1,2 units), where the catalyst promotes η³-allyl intermediates leading to alternating tacticities. Microstructure is further tuned by ligand modifications—bulky phosphines enhance cis-1,4 selectivity in systems, while aromatic ligands shift catalysts toward 1,2-syndiotactic products—and by , with lower temperatures (below 20°C) favoring cis over trans configurations. The occurs in , typically in aliphatic or aromatic hydrocarbons at 20–50°C, using low catalyst loadings (e.g., 0.1–1 mmol metal per 100 g ) to achieve high activities (up to 10⁶ g /mol metal·h). Stereoregularity arises from the diene's bidentate coordination to the metal center in the s-cis form, which directs the migratory insertion and chain growth while minimizing side reactions like cyclization. Cis-1,4-polybutadiene, resembling in its low temperature (T_g ≈ -100°C) and high , is widely used in treads for superior and wet grip. In contrast, syndiotactic 1,2-polybutadiene features a high (200–220°C) and crystallinity, enabling applications in engineering plastics, films, and impact modifiers. These coordination methods produce synthetic rubbers with controlled molecular weights (10⁴–10⁶ g/mol) and narrow polydispersities, avoiding the formation and impurities associated with .

Acrylate and Other Specialty Monomer Polymerization

Coordination polymerization of acrylates and other specialty vinyl monomers presents significant challenges due to the polar ester groups, which strongly coordinate to early transition metal catalysts like traditional Ziegler-Natta systems, leading to catalyst poisoning and deactivation. Late transition metal catalysts, particularly Pd(II) complexes with α-diimine ligands developed by Brookhart and coworkers in the 1990s, overcome this by tolerating polar functionalities through weaker interactions with the metal center, enabling effective insertion of monomers like methyl acrylate. These cationic Pd(II) systems facilitate copolymerization of ethylene with methyl acrylate in solution at room temperature, producing random copolymers with up to 18 mol% acrylate incorporation and low branching due to reduced chain-walking. Neutral catalysts, often featuring P,O-coordinating ligands, further advance this field by enabling living characteristics and high polar incorporation without the need for activators like MAO. For instance, bisphosphine monodentate Ni(II) complexes achieve copolymerization of and with contents exceeding 20 mol%, yielding polymers with narrow molecular weight distributions (PDI ≈ 1.2-1.5) indicative of controlled, living-like processes at ambient conditions. These catalysts produce alternating copolymers or homopolymers of acrylates with minimal β-hydride elimination, resulting in linear chains suitable for applications requiring precise microstructures. Polyacrylates synthesized via these methods serve as key materials in adhesives and coatings, leveraging their tunable and low branching for enhanced properties. Styrene-acrylate copolymers, accessible through α-diimine catalysis in the Brookhart system, exhibit improved thermal stability and are used in impact-resistant plastics. In the , developments with Fe(II) and Co(II) complexes bearing tri- and tetradentate nitrogen ligands, activated by MAO, enabled coordination polymerization of acrylates to produce higher homopolymers (Mn up to 10,000 g/mol) with reduced branching compared to methods, though primarily atactic. These systems highlight progress toward stereoregular polyacrylates, with Fe catalysts showing superior activity over Co analogs in solution processes at moderate temperatures.

Coordination Polymerization of Other Substrates

Acetylene Polymerization

Coordination polymerization of employs complexes, notably and species, to form highly conjugated s through controlled insertion or metathesis of the . (I) catalysts, such as the dimer [Rh(norbornadiene)Cl]2 ([Rh(nbd)Cl]2) activated by triethylamine, enable the synthesis of cis-rich films via a living process that yields high molecular weight polymers (up to 106 Da) with narrow polydispersity. This approach was first demonstrated in the early , producing free-standing films of from gaseous under mild conditions. -based catalysts, including alkylidene complexes like W(CPh)(CO)5, facilitate through a metallacyclobutane mechanism, often leading to cyclic or linear with high activity (e.g., 620,000 g/mol/h turnover). The mechanism for rhodium-catalyzed polymerization involves coordination of the to the center, followed by cis-insertion into a metal-carbon bond, propagating the chain while maintaining stereoregularity and avoiding termination for living character. This contrasts with systems, where metathesis via [2+2] cycloaddition forms transient metallacycles, enabling ring-expansion or linear growth, particularly for substituted acetylenes. Polymerizations are typically conducted in solution (e.g., THF or ) at low temperatures (0 to -20°C) to suppress side reactions like cyclotrimerization to , ensuring high yields (>90%) of defect-free conjugated chains. For substituted acetylenes like , catalysts produce stereoregular cis-transoid polymers with molecular weights exceeding 500,000 Da, enhancing solubility and processability. The resulting polyacetylenes exhibit extended π-conjugation, conferring unique electronic properties; undoped cis-polyacetylene is semiconducting, but doping with oxidants like iodine or AsF5 introduces charge carriers, achieving metallic conductivity up to 105 S/cm. This doping-enabled conductivity laid the foundation for conducting polymers, earning the 2000 for related discoveries by Heeger, MacDiarmid, and Shirakawa, though and routes provide alternatives for precise control. Substituted polyacetylenes, such as poly(), offer improved environmental stability and are used in , including field-effect transistors and sensors, due to their tunable bandgap and helical conformations. These materials highlight coordination polymerization's role in accessing functional conjugated systems beyond traditional Ziegler-Natta methods.

Polar Monomer Polymerization

Coordination polymerization of polar monomers, such as (), epoxides, and lactones, enables the synthesis of functional polymers with incorporated heteroatoms, distinguishing it from traditional olefin polymerizations by involving migratory insertion or ring-opening mechanisms that accommodate polar functionalities. These processes typically employ late catalysts to mitigate the strong coordination of oxygen-containing groups that can early metal systems. Key examples include the production of polyketones from CO and olefins, polyethers from epoxides, and polyesters from lactones, offering materials with enhanced properties like barrier performance and biodegradability. A prominent application is the alternating copolymerization of with , catalyzed by complexes bearing diphosphine ligands, which yield high-molecular-weight polyketones. These catalysts, pioneered by Drent and coworkers at in the , facilitate perfectly alternating microstructures through selective insertion of and , leading to semicrystalline thermoplastics commercialized as Carilon for applications such as automotive parts due to their mechanical strength and chemical resistance. The proceeds via migratory insertion of the olefin into a Pd-alkyl , followed by CO insertion, with diphosphine ligands providing hemilabile coordination to achieve high and suppress homopolymerization. For epoxides, chromium-salen complexes serve as effective catalysts for to produce polyethers or, in combination with or CO2, polycarbonates. These tetradentate salen ligands stabilize the Cr(III) center, enabling nucleophilic attack on the coordinated , often with bis()iminium (PPNCl) as a cocatalyst to promote alternating copolymerization. A notable advancement is the CO2/ coupling to form polycarbonates, initially demonstrated with cyclohexene in the late 1990s and refined in the 2020s for , yielding fully alternating, high-molecular-weight polymers (>100,000 g/) with over 99% linkages for sustainable plastics in initiatives. Recent developments emphasize cobalt-salen variants for improved activity and stereocontrol, addressing scalability for industrial CO2 utilization. Lactone polymerization follows a similar coordination-insertion pathway, with aluminum or rare-earth metal alkoxides initiating ring-opening to form polyesters like . These catalysts ensure living polymerization conditions, producing narrow polydispersity materials for biomedical uses, though activity is modulated by design to prevent side reactions. Challenges in these systems include low catalytic activity with early transition metals, such as or , due to excessive oxophilicity that traps polar monomers and deactivates the catalyst. To overcome this, recent efforts focus on biomimetic designs inspired by metalloenzymes, incorporating earth-abundant metals like iron or with tailored ligands to enhance and efficiency in CO2/ processes.

References

  1. [1]
    Coordination Polymerization - an overview | ScienceDirect Topics
    Ziegler–Natta catalysts are formed by complexes of alkyls of metals from groups I–III with halides and other derivatives from the transition metals of groups IV ...
  2. [2]
  3. [3]
    The Influence of Ziegler-Natta and Metallocene Catalysts on ...
    The original equipment used by Karl Ziegler for discovering his catalyst and co-catalyst systems for the polymerization of ethylene to high-density ...
  4. [4]
    Discovery of Syndiotactic Polystyrene: Its Synthesis and Impact
    ### Summary of Coordination Polymerization from https://pubs.acs.org/doi/10.1021/acs.macromol.0c00350
  5. [5]
    Introduction to Polymerization and Depolymerization
    Apr 1, 2025 · Coordination polymerization is a method of polymerization that involves the use of metal catalysts, particularly transition metal complexes, to ...
  6. [6]
    Coordination Polymerization - an overview | ScienceDirect Topics
    Coordination polymerization is another variant of addition polymerization which is neither free radical nor simply ionic.
  7. [7]
    Applications in homogeneous coordination-insertion polymerization ...
    Jul 1, 2019 · Coordination insertion polymerization of polar and non-polar monomers. •. Ring opening polymerization of cyclic esters and related monomers.
  8. [8]
    [PDF] NOVEL PRECISION POLYOLEFINS FROM LIVING ... - DRUM
    These polymerization methods include anionic, cationic, radical, ring opening metathesis, and coordination polymerization. Page 21. 2. Living polymerization, ...
  9. [9]
    Karl Ziegler and Giulio Natta | Science History Institute
    In the early 1950s Ziegler and Natta discovered a new process for synthesizing polymers that made possible a lot of common plastics, including high-density ...
  10. [10]
    Ziegler-Natta catalysis: 50 years after the Nobel Prize | MRS Bulletin
    Mar 13, 2013 · This introductory article will provide a short historical review of the discovery of catalyzed olefin polymerization by Ziegler, Natta, and others
  11. [11]
    The Nobel Prize in Chemistry 1963 - NobelPrize.org
    The Nobel Prize in Chemistry 1963 was awarded jointly to Karl Ziegler and Giulio Natta "for their discoveries in the field of the chemistry and technology ...Missing: coordination | Show results with:coordination
  12. [12]
    The Influence of Ziegler-Natta and Metallocene Catalysts on ... - MDPI
    3rd generation catalysts supported on MgCl2 commercialized by many companies. 1977–1980, Kaminsky and Sinn discover high activity metallocene single-site ...
  13. [13]
    Metallocene Catalysts Initiate New Era In Polymer Synthesis
    The sandwich structures have been known for decades, but were not considered practical as catalysts. Then, in the mid-1980s, German professors Walter Kaminsky ...
  14. [14]
    Manufacturing - Techniques de l'Ingénieur
    Apr 10, 2011 · The first production units date back to the mid-1950s. In 1955, Hoechst started up the first unit based on Ziegler Natta catalysis, followed ...
  15. [15]
    Recent Advances in the Copolymerization of Ethylene with Polar ...
    The less-expensive and earth-abundant nickel catalyst is highly promising in the copolymerization of ethylene with polar monomers and has thus attracted ...
  16. [16]
    Potentially Practical Catalytic Systems for Olefin-Polar Monomer ...
    Nov 20, 2023 · For decades, many transition metal catalysts have been investigated for use in olefin-polar monomer copolymerization. However, most ...
  17. [17]
    Recent advances in thermally robust, late transition metal‐catalyzed ...
    Aug 27, 2018 · This review focuses on the development of Ni-, Pd-, Fe- and Co-based olefin polymerization catalysts that display enhanced thermal stability ...
  18. [18]
    Ziegler-Natta Catalysis I: α-olefin Polymerization Mechanism
    A mechanism for the polymerization of olefins is proposed. An essentially octahedrally coordinated ion of a transition element with empty or nearly empty ...
  19. [19]
    The evolution of stereoselective olefin polymerization modeling
    Jan 12, 2024 · We review the story of stereoselective 1-olefin polymerization with a focus on the production of isotactic polypropylene.
  20. [20]
    Catalytic-Site-Mediated Chain-End Control in the Polymerization of ...
    A dependance of stereocontrol on the steric bulk of the ligand, on the initial monomer concentration and on the symmetry of the catalytic site support that the ...
  21. [21]
    Variability of Chain Transfer to Monomer Step in Olefin Polymerization
    Computational studies of a variety of polymerization catalyst models have revealed an unexpected fluidity in chain termination mechanisms.
  22. [22]
    Theoretical Studies of the Factors Controlling Insertion Barriers for ...
    The use of ligands having donating and bridging atoms that are capable of donating electron density to the cationic metal center decreases the insertion ...
  23. [23]
    Structural Characterization of Electron Donors in Ziegler–Natta ...
    Ziegler–Natta catalysis is a very important industrial process for the production of polyolefins. However, the catalysts are not well-understood at the ...
  24. [24]
    Active Site Formation in MgCl2−Supported Ziegler−Natta Catalysts ...
    The mechanism proposed allows us to rationalize the domination of Ti3+ over Ti2+ under reduction of Ti4+ surface species by AlR3, the absence of an electron ...
  25. [25]
    Metal Complexes for Catalytic Polymerizations | Inorganic Chemistry
    Dec 20, 2019 · (3) Catalysts for the ring-opening polymerization of cyclic monomers may also prove useful in lowering activation barriers to chemical recycling ...
  26. [26]
    Insight into the Synthesis Process of an Industrial Ziegler–Natta ...
    Dec 21, 2018 · We fully characterize a single-step catalyst preparation process, which comprises a reactive precipitation of a MgCl 2 -supported Ziegler–Natta catalyst.Missing: HDPE | Show results with:HDPE
  27. [27]
    Internal donors on Ziegler Natta catalysts for isotactic polypropylene
    Oct 1, 2021 · This 4th Gen catalysts with certain breadth on MWD with polydispersity index (PDI): 4–7 have good performance on temperature profile and also on ...<|control11|><|separator|>
  28. [28]
    Fundamental aspects of heterogeneous Ziegler–Natta olefin ...
    In this review, we have highlighted the chronological development of a heterogeneous Z–N catalyst and its polymerization (ethylene, propylene and 1-butene) ...
  29. [29]
    Data Driven Modeling of Ziegler–Natta Polypropylene Catalysts
    Mar 24, 2025 · In the present paper, we illustrate how a smaller but more homogeneous ID subset containing diesters only can be utilized for “clear-box” QSPR modeling.
  30. [30]
    Influence of the reaction conditions on the Ziegler-Natta catalyzed ...
    Feb 2, 2024 · This work conducts an investigation into how the polymerization conditions affect the kinetics of ethylene homopolymerization and the properties of the final ...
  31. [31]
    a Direct Insight into Ziegler–Natta Catalysis | The Journal of Physical ...
    Jun 25, 2020 · As a matter of fact, it has been reported that the accessible Ti3+ in Ziegler–Natta catalysts reacts with O2 via an electron-transfer mechanism ...
  32. [32]
    Polyolefin Catalyst Market Statistics & Share | Industry Overview - 2026
    Over 90% of global polypropylene production uses ZN catalyst, whereas in polyethylene production Ziegler Natta is used more than 50% in all cases. Phillips ...<|control11|><|separator|>
  33. [33]
  34. [34]
    38. Homogeneous Ziegler–Natta Catalysts and Synthesis of ...
    Most Ziegler–Natta catalysts have maximum activity at some optimal temperature; the activity decreases at both lower and higher temperature because of catalyst ...Missing: thermal stability
  35. [35]
    [PDF] Giulio Natta - Nobel Lecture
    In fact, the only known polymer having 3,4-enchainment, obtained in the presence of the same catalysts yielding syndiotactic 1,2-polybutadiene, is amorphous. 1, ...
  36. [36]
    Recent advances in the polymerization of butadiene over the last ...
    This paper provides an exhaustive overview of the latest developments in the field of the stereospecific polymerization of 1,3-butadiene.Missing: seminal | Show results with:seminal
  37. [37]
    Polymerization of 1,3-butadiene catalyzed by Co(II) and Ni(II ...
    Apr 22, 2025 · Polymerization of 1,3-butadiene has long been studied as a process by which the economically favorable monomer is converted into synthetic ...
  38. [38]
    Synthesis of Ultra-High-Molecular-Weight Aromatic Polyacetylenes
    The cis-transoidal PPA synthesized by the Rh catalyst was amorphous powder with yellow color. The amorphous PPA was found to change into crystalline polymer ...
  39. [39]
    [PDF] A spectroscopic study of polyacetylene prepared by using Rh(I ...
    Recently, we reported for the first time that univalent rhodium complexes are able to promote acetylene polymerization, and thin free-standing films of poly-.
  40. [40]
    Polymerization of Phenylacetylene Initiated by Tungsten Carbene ...
    ABSTRACT: Polymerization of phenylacetylene initiated by tungsten carbene complex 4 was investigated in various conditions. The rate of polymerization in ...
  41. [41]
    Cyclic Polyacetylene - PMC - PubMed Central - NIH
    The tungsten catalyst demonstrates an acetylene polymerization activity of 620,000 g/mol/h, and the c-PA produced has low defect density (<1%, 13C NMR) and ...
  42. [42]
    Recent Advances in the Synthesis of Substituted Polyacetylenes
    The fact that the controlled polymerization of phenylacetylenes using rhodium(I) catalysts occurs via a cis-insertion mechanism suggests that polymerization ...Missing: seminal | Show results with:seminal
  43. [43]
    Mechanism of tungsten-catalyzed cyclic polyacetylenes synthesis
    Mar 26, 2023 · A tetraanionic pincer-ligand supported tungsten (VI) alkylidene polymerizes acetylenes with the highest activity among alkyne polymerization ...
  44. [44]
    A spectroscopic study of polyacetylene prepared by using Rh(I ...
    The polymerization of acetylene, using [Rh(1,5-Cod)Cl]2, where 1,5-Cod is cis,cis-cycloocta-1,5-diene or [Rh(NBD)Cl]2, where NBD is bicyclo〈2,2,1〉hepta-2 ...Missing: catalyzed | Show results with:catalyzed
  45. [45]
    Living Polymerization of Phenylacetylene by Novel Rhodium ...
    Living Polymerization of Phenylacetylene by Novel Rhodium Catalysts. Quantitative Initiation and Introduction of Functional Groups at the Initiating Chain End.
  46. [46]
    The Nobel Prize in Chemistry 2000 - Popular information
    One of Heeger's students measured the conductivity of the iodine-doped trans-polyacetylene and – eureka! The conductivity had increased ten million times!
  47. [47]
    [PDF] Hideki Shirakawa - Nobel Lecture
    At that time, to my regret, I did not recognize that this carbocation could be a charge carrier and thus polyacetylene could be the first conducting polymer. To ...
  48. [48]
    Rhodium-Catalyzed Polymerization of Phenylacetylene and its ...
    May 29, 2013 · The present report overviews relevant aspects of rhodium-catalyzed polymerization of phenylacetylenes and substituted acetylene monomers.
  49. [49]
    Copolymerization of CO2 and Epoxides Catalyzed by Metal Salen ...
    These complexes exhibited enhanced catalytic activity over the first generation heterogeneous zinc catalysts, providing copolymers with very high CO2 content ...Optimization of CO2... · Investigation of CO2... · Attempts to Utilize Main-Group...
  50. [50]
    Efficient palladium catalysts for the copolymerization of carbon ...
    Efficient palladium catalysts for the copolymerization of carbon monoxide with olefins to produce perfectly alternating polyketones ... Drent and E. Klei ...
  51. [51]
    Standing The Heat - C&EN - American Chemical Society
    Aug 24, 2009 · Carilon was an aliphatic polyketone Shell developed when it was looking for industrial uses of carbon monoxide. Carilon had promising ...
  52. [52]
    (Salen)CrIIIX Catalysts for the Copolymerization of Carbon Dioxide ...
    When a strongly donating cocatalyst is present, the chromium−nucleophile bond is weakened suitably to allow concerted epoxide ring-opening to occur. Indeed, ...
  53. [53]
    A New Catalyst for the Living Polymerization of Lactones to Polyesters
    A New Catalyst for the Living Polymerization of Lactones to Polyesters ... Coordination polymerization of heterocyclic and heterounsaturated monomers.<|control11|><|separator|>
  54. [54]
    A review on catalytic copolymerization of carbon dioxide and epoxides
    Schiff base (salen) catalysts have emerged as highly active catalytic systems for CO2/epoxide copolymerization, with chromium- and cobalt-centered Salen ...Missing: seminal | Show results with:seminal
  55. [55]
    Environmentally benign metal catalyst for the ring-opening ...
    Dec 7, 2021 · This work reviews the development of sustainable metal catalysts for the production of CO 2 copolymer, centered by Al, Mg, Ti, Fe, generally acknowledged as ...Missing: seminal | Show results with:seminal<|control11|><|separator|>