Fact-checked by Grok 2 weeks ago

Catalytic cycle

A catalytic cycle is a sequence of steps in which a participates to accelerate a chemical without being consumed, regenerating itself at the end of the cycle to enable multiple turnovers of into product. This process is fundamental to , where the lowers the of the reaction by providing an alternative pathway, often involving binding of reactants to an , through intermediates, and release of products. In molecular catalysis, the cycle typically includes adsorption of the onto the (such as a metal center), stepwise reactions that form the product, and desorption, ensuring the returns to its initial state for reuse. Catalytic cycles are classified into , where the catalyst and reactants are in the same phase (e.g., solution-based organometallic processes); , involving solid catalysts and gaseous or liquid reactants; and biocatalysis, mediated by enzymes in biological systems. Key metrics for evaluating cycle efficiency include the turnover number (TON), which quantifies the total number of substrate molecules converted per catalyst molecule before deactivation, and the turnover frequency (TOF), defined as the number of turnovers per unit time, reflecting the instantaneous rate under specific conditions. The rate-determining step, or turnover-limiting step, within the cycle governs overall efficiency, as it possesses the highest activation barrier; optimizing this step through catalyst design can significantly enhance performance. These cycles are crucial in applications, such as the of pharmaceuticals, polymers, and fuels, where they enable efficient, selective of complex molecules from abundant feedstocks. Representation of catalytic cycles often employs graphical schemes to illustrate of steps, intermediates, and profiles, aiding in mechanistic studies via computational modeling and experimental . Understanding and engineering catalytic cycles continues to drive advancements in sustainable , reducing demands and waste in chemical processes.

Fundamentals

Definition

A catalyst accelerates the rate of a by providing an alternative reaction pathway that possesses a lower compared to the uncatalyzed process, without being consumed in the overall transformation. This phenomenon, known as , enables efficient conversion of substrates into products under milder conditions than would otherwise be required. In , a catalytic cycle refers to a sequence of elementary reactions that form a closed loop, wherein is temporarily transformed or consumed in one step but fully regenerated in a subsequent step, thereby facilitating multiple turnovers of substrate to product without net loss of . The cycle typically involves coordination of substrates to the metal center, bond-breaking and bond-forming events, and release of products, ensuring returns to its initial state to initiate another iteration. Precatalysts, which are inactive precursors, may first undergo to generate the true catalytic species that enters the cycle. The concept of the catalytic cycle emerged in the mid-20th century within the burgeoning field of , building on foundational work in . Early exemplars include Reppe's pioneering investigations into chemistry during the 1940s at , where nickel-based catalysts enabled processes such as the cyclotrimerization of to and cyclotetramerization to , demonstrating regenerative catalytic loops under high-pressure conditions. Conceptually, a catalytic cycle is often represented schematically as a circular or , illustrating 's entry into the loop via interaction with substrates, progression through intermediates leading to product formation, and eventual regeneration of the catalyst for , emphasizing the cyclical that underpins catalytic efficiency.

Key Principles

A catalytic cycle operates as a closed sequence of elementary reactions in which the catalyst is regenerated at the end of each iteration, enabling it to participate repeatedly in the of substrates into products without . This regeneration ensures the cycle's and distinguishes catalysis from stoichiometric processes. Catalytic efficiency is quantified primarily through two key metrics: the (TON), defined as the total number of turnovers (cycles) a catalyst achieves until deactivation, which reflects its overall productivity and stability; and the turnover frequency (TOF), the instantaneous rate of turnovers per per unit time (typically in s⁻¹), indicating the catalyst's speed under specified conditions such as and . From a thermodynamic perspective, catalytic cycles facilitate reactions by providing an alternative pathway that lowers the (E_a) barrier, thereby increasing the reaction rate without altering the overall thermodynamics of the process. The change (ΔG) for the net reaction remains unchanged, as catalysts accelerate both forward and reverse reactions equally, preserving the equilibrium constant derived from thermodynamic parameters like (ΔH) and (ΔS). For the cycle to drive net progress, the overall ΔG must be negative (favorable), ensuring spontaneity while the catalyst merely modulates the to make the process viable at lower energies. Kinetic analysis of catalytic cycles focuses on the interplay of steps, where the rate-determining step—the slowest in the sequence—governs the overall , often dictating the cycle's efficiency. To model this, the steady-state is commonly applied, assuming that concentrations of transient intermediates remain constant over time because their rates of formation and consumption are balanced. This simplification allows derivation of effective rate laws, such as for a simple cycle involving catalyst binding to : \text{rate} = k \, [\text{cat}] \, [\text{substrate}] where k is the effective rate constant aggregating the cycle's kinetics. Such approximations are essential for predicting behavior in complex mechanisms involving intermediates.

Components

Precatalysts

Precatalysts are stable compounds that serve as precursors to the active catalytic in , generating the catalytically active form during the reaction. This design addresses the inherent instability of many active catalysts toward air or moisture, allowing for safer handling and storage without compromising reactivity. Precatalysts are particularly prevalent in homogeneous catalytic cycles, where they enable the use of sensitive organometallic under practical conditions. Common types of precatalysts include complexes with stabilizing s, such as palladium(II) species in cross-coupling reactions that are reduced to the active palladium(0) form. For instance, (η³-allyl)Pd(NHC)(Cl) complexes, where NHC denotes an N-heterocyclic , act as effective precatalysts for Suzuki-Miyaura couplings due to their robustness. Earlier examples feature phosphine-based s, as seen in , RhCl()₃, which employs to stabilize the center for . Activation of precatalysts typically involves processes like ligand dissociation, , or to unmask the active species. In , the initial step is the dissociation of one ligand to form the coordinatively unsaturated RhCl(PPh₃)₂, which then undergoes of dihydrogen. For Pd(II) precatalysts, activation often proceeds via reduction to Pd(0), facilitated by alcohols or other reductants in the reaction medium, with solvent polarity and additives playing key roles in controlling the rate and speciation. These mechanisms ensure efficient conversion to the active catalyst while minimizing side reactions. The primary advantages of precatalysts include enhanced stability for improved handling and prolonged shelf life, as well as tunable selectivity through ligand choice. Historically, Wilkinson's catalyst marked a pivotal development in 1965, demonstrating how a stable Rh(I) complex could be activated for selective hydrogenation under mild conditions, paving the way for modern precatalyst designs.

Intermediates and Regeneration

In catalytic cycles, particularly those involving organometallic compounds, reactive intermediates play a crucial role as transient species that facilitate bond breaking and forming processes. Sigma (σ) complexes, formed through σ-bond donation from ligands such as η²-H₂ or alkyl groups to the metal center, often serve as key intermediates in oxidative addition steps, where they enable the cleavage of substrate bonds by increasing the metal's oxidation state and coordination number. Pi (π) complexes, involving π-bond donation from unsaturated ligands like alkenes (η²-C₂H₄) accompanied by back-bonding into the ligand's π* orbital, activate substrates for migratory insertion or nucleophilic attack, thereby promoting C-C or C-H bond formation while stabilizing the complex during turnover. These intermediates typically exist at steady-state concentrations, ensuring efficient cycle progression without accumulation. The regeneration step concludes the catalytic cycle by returning the catalyst to its initial active form through a final elementary reaction, most commonly reductive elimination or ligand exchange. In reductive elimination, two cis-ligands on the metal undergo coupling to form a new bond in the product, simultaneously decreasing the metal's oxidation state, coordination number, and electron count by two, as seen in the reverse of oxidative addition. Ligand exchange, often dissociative, replaces spent ligands with incoming substrates or auxiliaries, restoring the coordination sphere and enabling reuse, particularly in cycles requiring flexible ligand environments. This step is essential for achieving high turnover numbers, as it closes the loop and prevents irreversible binding. Spectroscopic methods are vital for detecting and characterizing these elusive intermediates. Nuclear magnetic resonance (NMR) identifies species like metal hydrides (chemical shifts 0–60 ppm) or π-bound alkenes through characteristic signals and relaxation times (T₁ measurements), often under conditions to capture dynamic behavior. Infrared (IR) spectroscopy detects π-backbonding effects via shifts in vibrational frequencies, such as CO stretches (1820–2150 cm⁻¹) or M–H modes (1500–2200 cm⁻¹), providing evidence of intermediate stability during reaction monitoring. Complementing these, (DFT) computational modeling predicts intermediate geometries and energies with typical accuracies of ±0.2 Å for bond lengths and ±5 kcal/mol for relative stabilities, aiding in the design of robust cycles by simulating unobserved species. A major challenge in maintaining catalytic efficiency arises from side reactions that lead to catalyst decomposition, notably β-hydride elimination. This process converts metal-alkyl intermediates into alkenes and metal-hydrides when a β-hydrogen is available and the M–C–C–H allows , often resulting in irreversible deactivation and reduced lifetimes. Strategies to mitigate this include using sterically hindered ligands to enforce noncoplanar geometries or selecting substrates lacking β-hydrogens, such as aryl groups, to preserve the 's integrity throughout multiple turnovers.

Types

Homogeneous Cycles

Homogeneous catalytic cycles occur in a single , typically a , where the catalyst and reactants are molecularly dispersed without phase boundaries, allowing for intimate interactions that facilitate precise control over reaction pathways. These cycles generally involve soluble organometallic complexes that undergo a sequence of elementary steps—such as , insertion, elimination, and —to convert substrates into products while regenerating the active catalyst . Precatalysts, such as Pd(OAc)<sub>2</sub>, are commonly employed in these systems and are activated to generate the true catalytic . A representative example is the palladium-catalyzed , which enables the formation of carbon-carbon bonds between s and s. The cycle begins with the of an aryl halide to a Pd(0) , forming an aryl-Pd(II)-halide complex. This is followed by coordination of the alkene to the metal center and subsequent migratory insertion, yielding a σ-alkyl-Pd(II) intermediate. Beta- elimination from this intermediate produces the trans-stilbene-like product and a Pd(II)- , which undergoes to release the product and regenerate the Pd(0) catalyst. In optimized homogeneous conditions, this cycle achieves turnover numbers (TONs) exceeding 2000 and turnover frequencies (TOFs) up to 150 h<sup>-1</sup>, demonstrating efficient for C-C . Another prominent example is mediated by ruthenium-based catalysts, which redistributes substituents through a series of . The mechanism, established by Chauvin, initiates with a [2+2] cycloaddition between the metal (alkylidene) and an incoming , forming a metallacyclobutane intermediate. This intermediate then undergoes a reverse [2+2] cycloaddition, releasing a new product and regenerating the metal catalyst. The cycle's efficiency is highlighted by TOFs reaching up to 1000 h<sup>-1</sup> in ring-closing metathesis applications, with TONs often in the range of 1000–4000 for various substrates. Homogeneous cycles offer advantages such as high selectivity due to the ability to fine-tune environments around the metal center, enabling stereochemical control and tolerance not easily achievable in other systems. However, a key limitation is the challenge in recovery and , as the soluble nature of the often leads to losses during product separation, necessitating advanced techniques like biphasic solvents or strategies to mitigate economic and environmental impacts.

Heterogeneous Cycles

Heterogeneous catalytic cycles involve immobilized catalysts, such as metal particles supported on oxides or zeolites, where reactants adsorb onto the catalyst surface, undergo transformation through surface-bound intermediates, and desorb as products, thereby regenerating the active sites. This process contrasts with solution-phase by relying on solid-gas or solid-liquid interfaces, enabling facile catalyst recovery but often at the expense of lower turnover frequencies (TOFs) compared to homogeneous systems. The cycle's efficiency depends on surface coverage and adsorption energetics, with active sites typically comprising undercoordinated metal atoms or defects that facilitate bond breaking and formation. A seminal example is the Haber-Bosch process for ammonia synthesis, utilizing iron-based catalysts promoted with potassium and alumina. In this cycle, dinitrogen (N₂) adsorbs dissociatively on the Fe(111) surface, forming surface nitride species; dihydrogen (H₂) then dissociates and adds stepwise to produce NH₃, which desorbs, freeing the sites for regeneration. This mechanism proceeds via the Langmuir-Hinshelwood pathway, where both reactants adsorb prior to reaction, with the rate-determining step being N₂ dissociation, achieving industrial TOFs on the order of 10⁻³ s⁻¹ per site under high-pressure conditions (200–300 bar, 400–500°C). Surface-bound intermediates, such as adsorbed NHₓ species, play a critical role in stabilizing the cycle. Another representative case is Ziegler-Natta polymerization of olefins, employing TiCl₄ supported on MgCl₂ with aluminum alkyl cocatalysts. The cycle initiates with of Ti sites to form Ti-alkyl bonds, followed by coordination and migratory insertion into the growing chain, culminating in or that renews the . This Cossee-Arlman operates at heterogeneous interfaces, yielding high-molecular-weight polyolefins with TONs reaching 10⁶–10⁷ per Ti atom, though site heterogeneity leads to broader distributions. Surface science underpins these cycles through mechanisms like Langmuir-Hinshelwood, which models adsorption isotherms and bimolecular surface reactions assuming uniform sites and rapid diffusion. Characterization techniques, including (XPS) for elemental composition and oxidation states, and (TEM) for morphology and particle size, reveal active site dynamics and deactivation pathways such as . While heterogeneous cycles generally exhibit lower TOFs (often <10² s⁻¹) due to mass transport limitations, their overall TONs can be high due to catalyst durability, and their primary advantage lies in straightforward separation via , enabling reuse in large-scale processes.

Biocatalytic Cycles

Biocatalytic cycles are mediated by in aqueous biological environments, where the protein scaffold provides a precise for binding, transformation via transient intermediates, and product release, regenerating the for subsequent cycles. These cycles often involve multiple steps coordinated by residues or cofactors, achieving remarkable selectivity and efficiency under mild conditions (ambient temperature and neutral ). A representative example is the zinc-dependent enzyme, which catalyzes the reversible hydration of CO₂ to and protons, essential for and regulation. The begins with of a Zn-bound molecule to form a , followed by CO₂ coordination and nucleophilic attack to generate a intermediate, which is then displaced by a new molecule to regenerate the . This enables extraordinarily high TOFs up to 10⁶ s⁻¹, with TONs limited primarily by enzyme lifetime rather than deactivation, often exceeding 10⁸ per . Biocatalytic cycles excel in regio- and due to the chiral pocket, facilitating complex biosyntheses, but challenges include sensitivity to environmental changes and difficulties in large-scale , addressed through or .

Sacrificial Catalysts

Sacrificial , sometimes referred to as stoichiometric catalysts, are substances that accelerate chemical by providing an pathway with lower but are irreversibly consumed during the process, without regeneration for further cycles. Unlike true catalysts, which operate in substoichiometric amounts and remain unchanged, sacrificial are employed in equimolar quantities relative to the substrates and become incorporated into the products or form byproducts. This consumption distinguishes them from regenerative catalytic cycles, as they lack the turnover essential for sustained activity. A representative example occurs in Friedel-Crafts reactions, such as the synthesis of for the dyestuffs industry, where aluminum chloride acts as a sacrificial by coordinating with the to form a reactive acylium ion, but ultimately complexes with the aromatic product, requiring and rendering it non-reusable. In the Suzuki-Miyaura cross-coupling, organoboranes like boronic acids function as stoichiometric agents, transferring the organic group to the center and integrating into the final biaryl product without regeneration, in contrast to the catalytic species. Similarly, in hydrosilylation reactions, serve as stoichiometric reductants, supplying the silyl moiety to the unsaturated substrate (e.g., alkenes or carbonyls) while enabling the catalyst to cycle, but the itself is fully consumed. Historically, the concept of sacrificial agents predates modern , with early applications sometimes mislabeling stoichiometric reagents as catalysts, such as degradable enzymes in certain bioprocesses before techniques were developed; the clear distinction from true was emphasized amid the rise of homogeneous organometallic systems, highlighting regeneration and turnover. This delineation gained prominence in frameworks, particularly Principle 9 of the 1998 guidelines, which prioritizes catalytic over stoichiometric processes to reduce waste. The primary implications of sacrificial reagents include elevated costs due to the need for equimolar addition and increased waste generation from byproducts, making them less efficient than true catalytic cycles for large-scale applications. However, they remain valuable in scenarios demanding exceptional selectivity during complex , where the precise incorporation of the reagent enhances product purity despite non-reuse.

Stoichiometric Processes

Stoichiometric processes in refer to in which the are consumed in exact 1:1 ratios relative to the product formed, without any recycling or regeneration of components, leading to no amplification of the reaction beyond the initial . This contrasts sharply with catalytic cycles, where a small amount of facilitates multiple turnovers, enabling efficient conversion of substrates without proportional consumption. In relation to catalysis, certain stoichiometric reactions mimic key intermediates observed in catalytic mechanisms but fail to achieve cycle closure, resulting in permanent consumption of the reactive species. For instance, the Grignard addition employs organomagnesium to form carbon-carbon bonds by nucleophilic attack on carbonyl compounds, producing alcohols after , yet requires fresh reagent for each equivalent of product without regeneration. This differs from catalytic C-C bond formations, such as those using transition metals, which recycle the catalyst to achieve higher efficiency. Representative examples of stoichiometric processes include oxidations with potassium permanganate (KMnO₄), where the permanganate ion acts as a one-electron or multi-electron oxidant to convert alkenes or alcohols to carboxylic acids or ketones, fully reducing to Mn²⁺ without recovery. A notable historical evolution occurred in the production of acetaldehyde from ethylene: early methods used stoichiometric palladium(II) chloride, which was consumed and generated waste, but the Wacker process transformed this into a catalytic system by incorporating copper(II) chloride and oxygen for palladium regeneration, drastically reducing metal usage. Stoichiometric processes often incur higher economic and environmental costs due to increased waste generation compared to catalytic cycles, as measured by like and the E-factor. calculates the percentage of reactant atoms incorporated into the desired product; for example, a stoichiometric oxidation might achieve only 50-60% due to byproduct formation, whereas catalytic variants approach 90-100%. The E-factor, which quantifies total waste per unit of product (kg waste/kg product), can exceed 100 for bulk stoichiometric processes like oxidations, versus under 5 for efficient catalytic ones, highlighting the advantages of shifting to cycles.

References

  1. [1]
    “Turning Over” Definitions in Catalytic Cycles | ACS Catalysis
    Nov 8, 2012 · A convenient way to express catalytic activity is by means of a turnover number equal to the number of reactant molecules converted per minute per catalytic ...Introduction · Typical Misuses of the TOF · “Turning Over” Definitions · Conclusion
  2. [2]
    [PDF] Understanding Catalysis - RSC Publishing
    Jun 20, 2014 · The large majority of chemical compounds underwent at least one catalytic step during synthesis. While it is common knowledge that catalysts ...
  3. [3]
    Molecular Catalysis | PNNL
    Catalysis is the process of speeding up a chemical reaction by adding a compound—a catalyst—that is the same at the start and end of the reaction.
  4. [4]
    [PDF] Organometallic Catalysis
    A catalyst may be defined by its Turnover Number (TN). Each time the complete catalyst cycle occurs, we consider one catalytic turnover to have been completed.
  5. [5]
    An Information-Rich Graphical Representation of Catalytic Cycles
    Oct 24, 2019 · In a catalytic cycle at steady state, the rate of each step of the cycle is the same. Reactions with small rate constants accumulate enough of ...
  6. [6]
    Activation Energies of Plasmonic Catalysts | Nano Letters
    In simple, textbook terms, a catalyst is defined by its ability to reduce the activation barrier for a chemical reaction.
  7. [7]
    [PDF] Key Topic Collection for Catalysis Teaching - Dechema
    The general definition specifies the catalytic cycle as a closed cycle of a multistep reaction that involves a cat- alyst as key component of the chemical ...
  8. [8]
    Catalytic Reactions of Acetylene: A Feedstock for the Chemical ...
    In Reppe's work, (29, 115) attempts to react acetylene with alcohols in the gas phase at temperatures between 150 and 350 °C over solid catalysts such as soda ...
  9. [9]
    Catalyst Fundamentals - DieselNet
    In this thermodynamic model, the catalyst provides a path with lower activation energy ... Thus, the reaction enthalpy (ΔH), reaction free energy (ΔG) and ...
  10. [10]
    Catalysis and the Use of Energy by Cells - NCBI - NIH
    The Concentration of Reactants Influences ΔG​​ As we have just described, a reaction A ⇌ B will go in the direction A → B when the associated free-energy change, ...
  11. [11]
    4.12: Steady-State Approximation - Chemistry LibreTexts
    Feb 12, 2023 · The steady-state approximation is a method used to derive a rate law. The method is based on the assumption that one intermediate in the reaction mechanism is ...The Steady-State Approximation · Definition: Intermediates · Example 4 . 12 . 1Missing: cycles | Show results with:cycles
  12. [12]
    Some Considerations on the Fundamentals of Chemical Kinetics
    Jul 20, 2017 · The elementary reaction sequence A ⇄ I → Products is the simplest mechanism for which the steady-state and quasi-equilibrium kinetic approximations can be ...A Simple Mechanism With... · Quasi-Equilibrium... · Scheme 1
  13. [13]
    Understanding Precatalyst Activation and Speciation in Manganese ...
    Apr 3, 2023 · In most reactions catalyzed by organometallic complexes, the bench-stable reagent added to a reaction is a precatalyst that requires activation ...
  14. [14]
    [PDF] THE ORGANOMETALLIC CHEMISTRY OF THE TRANSITION ...
    In this chapter we review some fundamental ideas of coordination chemistry, which also apply to organometallic complexes. 1.1 WERNER COMPLEXES. Complexes in ...
  15. [15]
    Homogeneous Catalysis - an overview | ScienceDirect Topics
    Homogeneous catalysis is defined as a catalytic process where the catalyst is in the same phase as the reactants, typically in a solution, and is characterized ...
  16. [16]
    Conventional reactors for homogeneous catalytic processes
    One of the key aspects of homogeneous catalysis is the catalytic cycle, which involves a series of steps that enable the transformation of reactants into ...
  17. [17]
    Mechanism, Reactivity, and Selectivity in Palladium-Catalyzed ...
    Jan 11, 2014 · The Heck reaction is an excellent method to introduce an aromatic group to one end of a double bond using palladium-catalysis ...Isomerization Of The Initial... · Scheme 3 · Free Energy Profiles And The...
  18. [18]
    Palladium Catalyst Recycling for Heck‐Cassar‐Sonogashira Cross ...
    Apr 27, 2021 · To our knowledge, a turnover number (TON) up to 2375, a turnover frequency (TOF) of 158 h−1, and a process mass intensity (PMI) around 7 that ...Missing: typical | Show results with:typical
  19. [19]
    Olefin metathesis: what have we learned about homogeneous ... - NIH
    ... olefin coordination prior to [2 + 2] cycloaddition of the olefin. The [2 + 2] cycloaddition yields a metallacyclobutane intermediate with a trigonal ...
  20. [20]
    Heterogeneous catalysts for olefin metathesis - ScienceDirect.com
    On the other hand, a significant limitation of the Grubbs catalyst was its poor thermal stability and low activity in the metathesis of substituted olefins.
  21. [21]
    Combining the Benefits of Homogeneous and Heterogeneous ... - NIH
    The greatest advantage of heterogeneous catalysis is the ease of separation, while the disadvantages are often limited activity and selectivity.
  22. [22]
    Chapter 13.8: Catalysis - Chemistry LibreTexts
    Jun 5, 2019 · Heterogeneous catalysts are easier to recover. · Collision frequency is greater for homogeneous catalysts. · Homogeneous catalysts are often more ...
  23. [23]
    Molecular approaches to heterogeneous catalysis - ScienceDirect
    Dec 1, 2021 · An alternative way to add molecular definition to catalytic sites on solid surfaces is via the adsorption of appropriate discrete molecules.
  24. [24]
    Insight into why the Langmuir–Hinshelwood mechanism is generally ...
    Mar 15, 2002 · In heterogeneous catalysis, the two main reaction mechanisms which have been proposed are the Langmuir–Hinshelwood and the Eley–Rideal.
  25. [25]
    [PDF] Reaction Mechanism and Kinetics for Ammonia Synthesis on the Fe ...
    May 13, 2019 · To provide guidelines to accelerate the Haber-Bosch (HB) process for synthesis of ammonia from hydrogen and nitrogen, we used Quantum Mechanics ...
  26. [26]
    Decoding technical multi-promoted ammonia synthesis catalysts
    Aug 21, 2025 · Ammonia is industrially produced by the Haber-Bosch process over a fused, multi-promoted iron-based catalyst. Current knowledge about the ...
  27. [27]
    Development and Recent Progress on Ammonia Synthesis ...
    Dec 10, 2020 · The focus of this review will be on the catalytic activity of iron, cobalt, nickel, and nitride catalyst systems plus the latest progress on ...
  28. [28]
    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
  29. [29]
    [PDF] Ziegler–Natta Polymerization and the Remaining Challenges
    catalytic polymerization of ethylene using titanium tetrachloride and ... [1] P Cossee, Ziegler–Natta Catalysis I. Mechanism of Polymerization of α-olefins.
  30. [30]
    In Situ Characterization of Catalysis and Electrocatalysis Using APXPS
    Jan 14, 2021 · Ambient pressure X-ray photoelectron spectroscopy (APXPS) has become a useful characterization tool for studies in heterogeneous catalysis and electrocatalysis.
  31. [31]
    Chemical Electron Microscopy (CEM) for Heterogeneous Catalysis ...
    Transmission electron microscopy (TEM) is an effective characterization tool that enables us to directly observe and reveal the microstructures of materials [1] ...
  32. [32]
    Catalysts for a green industry | Feature | RSC Education
    Jun 30, 2009 · A sacrifice: worst case catalyst. A sacrificial, or stoichiometric, catalyst is used once and discarded. The amount of waste produced is not ...Missing: definition | Show results with:definition
  33. [33]
    Catalytic versus stoichiometric reagents as a key concept for Green ...
    Abstract. 25 years ago catalysis was highlighted as a main principle of Green Chemistry. Selective catalysis enables an efficient utilisation of resources ...
  34. [34]
    Catalytic cycle - Sacrificial catalysts - chemeurope.com
    As the name implies, the sacrificial catalyst is not regenerated and irreversibly consumed. This sacrificial compound is also known as a stoichiometric catalyst ...
  35. [35]
    Suzuki Coupling - Organic Chemistry Portal
    Suzuki Coupling is a palladium-catalyzed cross coupling between organoboronic acid and halides, with the boronic acid activated with base.
  36. [36]
    Iron-Catalyzed H/D Exchange of Primary Silanes, Secondary ...
    Feb 18, 2022 · A synthetic study into the catalytic hydrogen/deuterium (H/D) exchange of 1° silanes, 2° silanes, and 3° siloxanes is presented, facilitated by iron-β- ...
  37. [37]
    (PDF) The History of Catalysis – From the Beginning to Nobel Prizes
    Aug 9, 2025 · It started long before the creation of the term "catalysis" in 1835 by Berzelius (1779-1848) to describe the phenomena of accelerating reaction ...
  38. [38]
    What is Catalysis? - American Chemical Society
    Catalysis – Selective catalytic reagents are superior to stoichiometric ones; they can minimize or eliminate waste, depending on the reaction, which is ...
  39. [39]
    Catalytic versus stoichiometric reagents as a key concept for Green ...
    Catalysts allow us to decrease the activation energy of chemical transformations, enhance the reaction rate, facilitate high selectivity and are crucial to ...
  40. [40]
    Stoichiometric vs Catalytic Reagents
    May 3, 2024 · Catalytic reagents are favored when efficiency, selectivity, and waste reduction are crucial, while stoichiometric reagents are preferred for their simplicity ...
  41. [41]
    Empowering alcohols as carbonyl surrogates for Grignard-type ...
    Nov 26, 2020 · The Grignard reaction is a fundamental tool for constructing C-C bonds. Although it is widely used in synthetic chemistry, it is normally ...
  42. [42]
    Best Synthetic Methods: Carbon-Carbon Bond Formation
    May 28, 2007 · The alkoxides derived from the addition of Grignard reagents can be converted to the ketones, but yields are better with the alkoxides derived ...
  43. [43]
    Oxidation of Organic Molecules by KMnO4 - Chemistry LibreTexts
    Jan 22, 2023 · Treatment of an alkylbenzene with potassium permanganate results in oxidation to give the benzoic acid. 1-header16.png. Note. The position ...The half-reaction and... · General Reactivity with... · Reactions with Specific...
  44. [44]
    [PDF] Progress in the synthesis of aldehydes from Pd-catalyzed Wacker ...
    May 9, 2023 · This report sank into oblivion until the stoichiometric to catalytic evolution of the reaction by researchers from Wacker Chemie AG, leading to ...
  45. [45]
    12 Principles of Green Chemistry - American Chemical Society
    Atom economy, which was developed by Barry Trost1, asks the question “what atoms of the reactants are incorporated into the final desired product(s) and what ...
  46. [46]
    Engineering a more sustainable world through catalysis and green ...
    Mar 2, 2016 · In practice, the E factor is more than 3 owing to the use of an excess of reagents and the less than 100% yield. This should be compared with ...Missing: comparison | Show results with:comparison