Fact-checked by Grok 2 weeks ago

Catalysis

Catalysis is the acceleration of a by a substance known as a catalyst, which provides an alternative pathway with a lower while remaining unchanged at the end of the . The catalyst does not alter the overall of the , such as the standard Gibbs change, but significantly enhances the by facilitating bond breaking and formation. The concept of catalysis was first systematically described in 1835 by Swedish chemist , who coined the term to explain phenomena where substances influenced reactions without being consumed, attributing it to a "catalytic force." Early observations of catalytic effects date back further, including the decomposition of by noted by French chemist Louis Jacques Thénard in 1818, but Berzelius's work unified these into a coherent framework. Subsequent developments, such as Wilhelm Ostwald's thermodynamic interpretations in the late , elevated catalysis to a cornerstone of , earning him the 1909 . Catalysis is broadly classified into several types based on the nature of the catalyst and reaction conditions. occurs when the catalyst is in the same (typically or gas) as the reactants, allowing for intimate molecular interactions but complicating catalyst recovery; examples include acid- or base-catalyzed esterification reactions. In contrast, involves a catalyst in a different , often a interacting with gaseous or reactants, which is prevalent in due to ease of separation; in catalytic converters exemplifies this by facilitating purification. Biocatalysis, or , employs biological catalysts like proteins to achieve high specificity and efficiency under mild conditions, as seen in metabolic pathways where enzymes such as accelerate . Additionally, autocatalysis arises when a reaction product itself acts as the catalyst, leading to sigmoidal rate profiles, such as in the oxidation of by . Catalysis plays a pivotal role in modern , underpinning over 90% of chemical processes and enabling the of essential commodities like fertilizers, fuels, and pharmaceuticals with reduced energy input and waste. For instance, the Haber-Bosch process for synthesis relies on iron-based heterogeneous catalysts to fix atmospheric , supporting global by producing fertilizers that sustain approximately half of the world's population. In petroleum refining, catalytic cracking using zeolites converts heavy hydrocarbons into , optimizing fuel yields. Beyond , catalysis is crucial in environmental applications, such as three-way catalysts in vehicles that convert harmful pollutants like , , and hydrocarbons into less toxic substances, mitigating . Emerging areas, including electrocatalysis for and for , highlight catalysis's potential in transitions. Overall, advancements in catalyst design continue to drive efficiency, selectivity, and sustainability across chemical, biological, and environmental domains.

Fundamentals

Definition and Principles

Catalysis is the process by which a substance, known as a , increases the rate of a without undergoing any permanent change itself or altering the overall standard Gibbs energy change of the reaction. The participates in the reaction by forming temporary intermediates but is regenerated at the end, allowing it to facilitate multiple reaction cycles. This acceleration occurs because the provides an alternative reaction pathway that circumvents the highest energy barrier of the uncatalyzed process, thereby enabling the reaction to proceed more rapidly under the same conditions. A key principle of catalysis is the reduction of the (E_a), which is the minimum energy required for reactants to reach the . In the uncatalyzed reaction, molecules must overcome a high E_a to form products, resulting in a slower rate; the catalyzed pathway lowers this E_a through intermediate steps, as illustrated in an energy diagram where the catalyzed curve shows a shallower peak compared to the uncatalyzed one. This lower E_a increases the proportion of reactant molecules with sufficient energy to react, without shifting the equilibrium position. Catalysts are also highly selective, favoring specific reactions or substrates while remaining inert to others, which enhances their utility in targeted chemical transformations. The efficiency of a catalyst is quantified by metrics such as the turnover number (TON), defined as the maximum number of moles of product formed per mole of catalyst under specified conditions before deactivation. The turnover frequency (TOF) extends this by measuring the TON per unit time, providing a rate of catalytic activity in moles of product per mole of catalyst per second. These parameters highlight the catalyst's productivity and are crucial for evaluating performance. The rate enhancement can be expressed generally as \text{rate}_{\text{catalyzed}} = k \cdot [\text{reactants}], where the rate constant k is amplified due to the reduced E_a. According to the , k = A e^{-E_a / RT}, a decrease in E_a exponentially increases k, with A as the , R the , and T the temperature in .

Basic Examples

One of the simplest demonstrations of catalysis is the decomposition of hydrogen peroxide (H₂O₂) into water and oxygen gas, represented by the reaction $2 \mathrm{H_2O_2} \rightarrow 2 \mathrm{H_2O} + \mathrm{O_2}. Without a catalyst, this reaction proceeds slowly at room temperature, producing oxygen bubbles at a minimal rate over hours or days. Adding manganese dioxide (MnO₂) as a heterogeneous catalyst dramatically accelerates the process, causing vigorous bubbling and rapid evolution of oxygen gas within seconds, as the solid catalyst provides a surface for peroxide molecules to adsorb and react without being consumed. Similarly, in a homogeneous catalysis example, potassium iodide (KI) dissolved in solution introduces iodide ions (I⁻) that speed up the decomposition, producing a visible foam of oxygen bubbles almost immediately upon mixing, with the iodide acting as an intermediate in the reaction pathway. Quantitative comparisons show that the catalyzed rate can be over 2,000 times faster than the uncatalyzed reaction under similar conditions, highlighting how catalysts lower the activation energy barrier to enhance reaction speed. In the atmosphere, catalysis plays a critical role in through a cycle involving radicals from chlorofluorocarbons (CFCs). The process begins with a (•) reacting with () to form (•) and oxygen (O₂): \mathrm{Cl \cdot + O_3 \rightarrow ClO \cdot + O_2}. The • then reacts with an oxygen (O) to regenerate • and produce another O₂ molecule: \mathrm{ClO \cdot + O \rightarrow Cl \cdot + O_2}. This net reaction destroys two molecules per cycle while recycling the , allowing a single • to deplete thousands of O₃ molecules before being removed. Proposed in seminal work by Molina and Rowland, this catalytic mechanism explains the enhanced loss in the due to anthropogenic CFCs. A practical application of catalysis appears in automotive catalytic converters, which reduce harmful emissions from internal combustion engines. These devices use platinum (Pt) and rhodium (Rh) supported on a ceramic honeycomb to facilitate oxidation of carbon monoxide (CO) to carbon dioxide (CO₂): \mathrm{CO + \frac{1}{2} O_2 \rightarrow CO_2}, and reduction of nitrogen oxides (NOₓ) to nitrogen gas (N₂) and oxygen: $2 \mathrm{NO} \rightarrow \mathrm{N_2 + O_2} (or similar for NO₂)./Kinetics/07:_Case_Studies-_Kinetics/7.01:_Catalytic_Converters) The metals provide active sites for these redox reactions at exhaust temperatures around 300–800°C, converting up to 98% of CO and NOₓ into less toxic products in modern three-way converters. This everyday example underscores catalysis's role in environmental protection by enabling efficient pollutant transformation without altering the overall engine chemistry./Kinetics/07:_Case_Studies-_Kinetics/7.01:_Catalytic_Converters)

Measurement and Units

The assessment of catalytic performance relies on standardized quantitative metrics that capture activity, productivity, and efficiency, enabling comparisons across catalysts and reaction conditions. Key among these are the turnover frequency (TOF) and (TON), which focus on the intrinsic activity per catalytic site. TOF quantifies the rate at which a catalyst converts substrate molecules, defined as the number of moles of reactant converted per mole of active sites per unit time, with common units of s⁻¹ or h⁻¹. According to IUPAC, this corresponds to molecules reacting per active site in unit time, a definition borrowed from but widely applied in chemical catalysis. TON, a dimensionless measure, represents the total number of moles of substrate converted per mole of catalyst before deactivation, serving as an indicator of catalyst stability and lifetime under specified conditions. Another important metric is space-time yield, which evaluates overall process productivity as the mass of product generated per unit reactor volume per unit time, typically in units of g/(L·h); this is particularly relevant for scaling up catalytic processes in industrial settings. Efficiency metrics complement activity measures by addressing product distribution. Selectivity is defined as the ratio of moles of desired product to the total moles of all products formed, often expressed as a , reflecting the catalyst's ability to favor specific reaction pathways over side reactions. , on the other hand, is the ratio of moles of desired product produced to moles of reactant converted, also typically reported as a , providing a direct gauge of conversion effectiveness toward the target product. These metrics are interrelated; for instance, yield can be expressed as the product of and selectivity, emphasizing the need for balanced optimization in catalyst design. The International Union of Pure and Applied Chemistry (IUPAC) provides foundational standardization for catalytic activity, defining it as the catalyzed rate of reaction in moles of reactant converted per unit time, with the coherent SI unit being the katal (kat ≡ 1 mol/s). In heterogeneous catalysis, where catalyst mass varies, specific catalytic activity is normalized to mol/(s·kg catalyst) to account for loading and enable comparability across systems. This unit facilitates precise reporting of performance independent of scale, as outlined in IUPAC guidelines for catalyst characterization. Experimental techniques for measuring these metrics are tailored to the reaction setup and catalyst type. Batch reactors are commonly used for and TOF determination, where a fixed amount of catalyst is added to reactants, and product evolution is tracked over time—often via sampling and analysis—to compute initial rates and total turnovers. Continuous flow reactors, by contrast, assess space velocity metrics, such as weight hourly space velocity (WHSV in g feed/(g catalyst·h)), by maintaining steady-state feed rates through a catalyst bed, allowing evaluation of long-term productivity and space-time under operational conditions. To probe catalyst state and mechanisms, spectroscopic techniques like () spectroscopy and (NMR) are employed; detects surface species and adsorption modes , while NMR reveals molecular-level changes in homogeneous or supported catalysts during operation. A representative calculation illustrates TOF application: for a reaction yielding 100 mol of product from 1 mol of catalyst over 1 h, assuming all catalyst sites are active, \TOF = \frac{100 \, \mol}{1 \, \mol \cdot 1 \, \h} = 100 \, \h^{-1}. This value highlights the catalyst's per-site efficiency, with higher TOF indicating greater activity, though actual computation requires verification of active site concentration via techniques like or .

Mechanisms and Energetics

Reaction Mechanisms

Catalytic reaction mechanisms typically proceed through a series of discrete steps that enable the catalyst to lower the for the of reactants into products while remaining unchanged at the end of the cycle. The fundamental sequence includes: (1) adsorption of one or more reactants onto active sites of the catalyst or coordination to the catalyst's reactive center; (2) a surface or intramolecular within the coordinated ; (3) desorption of the products from the catalyst sites; and (4) regeneration of the original catalyst form, allowing the cycle to repeat. These steps ensure efficient turnover, with the catalyst interacting reversibly with substrates to facilitate bond breaking and formation. In many catalytic processes, the Langmuir-Hinshelwood mechanism governs bimolecular reactions where both reactants must adsorb onto adjacent sites of the catalyst before undergoing transformation. This pathway assumes independent adsorption of each reactant following the Langmuir isotherm, leading to a surface-bound that reacts in the rate-determining step. The is proportional to the product of the fractional coverages of the two species, expressed as \text{rate} = k \theta_A \theta_B, where k is the rate constant and \theta_A, \theta_B are the coverages of reactants A and B, respectively. This mechanism highlights the importance of surface saturation and competitive adsorption in controlling overall . An alternative pathway, the Eley-Rideal mechanism, occurs when one reactant is adsorbed on the catalyst while the second approaches directly from the surrounding phase without prior adsorption. Here, the adsorbed species reacts with the incoming molecule at the interface, bypassing dual-site occupancy. The rate law simplifies to \text{rate} = k P_A \theta_B, where P_A is the (or concentration) of the non-adsorbed reactant A and \theta_B is the coverage of adsorbed B. This mechanism is favored in systems where co-adsorption leads to inhibition or when limitations prevent full surface coverage. Homogeneous catalytic mechanisms often operate via closed catalytic cycles comprising iterative coordination events, such as ligand exchange, which allow substrates to bind and products to depart while preserving the catalyst's integrity. In transition metal complexes, a typical cycle might involve oxidative addition of a reactant to the metal center, followed by migratory insertion or ligand coupling, and concluding with reductive elimination to release the product; ligand exchange steps, like substitution of a labile ancillary ligand for the substrate, ensure continuous regeneration. These cycles enable high selectivity and turnover in solution-phase reactions, with the metal's electronic and steric properties dictating the pathway. Stereoselectivity emerges in catalytic mechanisms through the catalyst's ability to impose facial discrimination on prochiral substrates, directing approach to one enantiotopic face during key bond-forming steps. In asymmetric transformations, chiral ligands or frameworks create a non-symmetric environment that stabilizes one over its , often via differential non-covalent interactions like hydrogen bonding or steric shielding. For example, in rhodium-catalyzed hydrogenations, diphosphine ligands enforce facial selectivity by positioning the substrate such that delivery occurs preferentially from the less hindered face, yielding enantiomerically enriched products with . This principle underpins the efficiency of chiral catalysts in generating molecular .

Energy Considerations

In catalysis, the energy profile of a reaction pathway reveals key distinctions between catalyzed and uncatalyzed processes. The uncatalyzed reaction typically exhibits a single, high barrier separating reactants from products. In contrast, the catalyzed pathway involves a series of steps with reaction intermediates, where each step has a lower , particularly for the rate-determining step, thus enabling the to proceed at a faster rate under milder conditions. Despite these kinetic advantages, the overall change (ΔG) from reactants to products remains identical in both cases, as the catalyst does not alter the thermodynamic favorability or position of the . Catalysts achieve this acceleration by reducing the (Ea) through preferential stabilization of the relative to reactants and intermediates. This reduction, expressed as ΔEa = Ea_uncat - Ea_cat, lowers the barrier that molecules must overcome to reach the reactive , often by several kcal/mol. Seminal work by emphasized that effective catalysts, such as enzymes, possess structures complementary to the , binding it more tightly than the ground-state substrates and thereby decreasing the required for its formation. Thermodynamically, catalysis impacts only the reaction kinetics and not the , preserving the overall ΔG and ensuring that the catalyst emerges unchanged. In reversible reactions, the catalyst enhances both forward and reverse rates proportionally, allowing the system to approach more rapidly without shifting the position dictated by . Catalysts facilitate this by modulating bond energies, such as weakening reactant bonds through adsorption or coordination to promote cleavage, or stabilizing high-energy intermediates via electrostatic or hydrogen-bonding interactions that lower their relative to uncatalyzed paths. The fundamental barrier in these processes is captured by the of activation from , developed by Henry Eyring: \Delta G^\ddagger = \Delta H^\ddagger - T \Delta S^\ddagger Here, catalysts primarily reduce ΔG‡ by lowering the enthalpic term ΔH‡ through binding, though favorable entropic changes (ΔS‡) can contribute in cases involving desolvation or conformational flexibility. This equation underscores how even modest reductions in ΔG‡ can yield exponential increases in reaction rates, as per the Eyring formulation of the rate constant.

Kinetics

In catalytic reactions, the rate law expresses the reaction velocity as a function of reactant concentrations, catalyst amount, and other factors, often derived from the underlying mechanism where the rate-determining step governs the overall kinetics. For homogeneous catalysis, where reactants and catalysts are in the same phase, rate laws typically involve the concentrations of substrates and catalyst; when the catalyst concentration is much lower than that of the substrates, pseudo-first-order approximations simplify the kinetics, treating the reaction as first-order in substrate with an effective rate constant proportional to catalyst concentration. In heterogeneous catalysis, involving distinct phases such as gas-solid systems, empirical power-law rate expressions are frequently applied, such as \text{rate} = k P^n, where P is the partial pressure of the reactant, k is the rate constant, and n is the reaction order reflecting pressure dependence, useful for initial modeling before detailed mechanistic insights. A key example of saturation kinetics in catalysis, originally developed for enzymes but generalizable to binding-limited processes in homogeneous and heterogeneous systems, is the Michaelis-Menten equation: v = \frac{V_{\max} [S]}{K_m + [S]} Here, v is the , V_{\max} is the maximum achievable rate when the catalyst is fully , [S] is the concentration, and K_m is the Michaelis constant representing the concentration at half V_{\max}, analogous to a for the catalyst- complex. This form emerges from mechanisms where binding precedes the rate-determining catalytic step and applies broadly when active sites become saturated at high levels, as observed in surface adsorption for . Inhibition modifies these rate laws by reducing effective catalyst activity; competitive inhibition occurs when an inhibitor binds reversibly to the , competing with the and increasing the apparent K_m while leaving V_{\max} unchanged, as higher substrate concentrations can overcome the inhibition. involves binding at a site distinct from the , decreasing V_{\max} by lowering the concentration of active catalyst without altering K_m, since binding affinity remains unaffected. These effects are derived from steady-state analysis of modified mechanisms and are prominent in biocatalysis but extend to poison adsorption in heterogeneous systems. Catalytic mechanisms are modeled using the steady-state approximation, which assumes intermediates like catalyst-substrate complexes maintain constant concentrations, allowing derivation of simplified rate laws from coupled differential equations for multi-step cycles. For complex catalytic cycles with multiple intermediates, numerical simulations solve the full system of ordinary differential equations to predict time-dependent concentrations and rates, enabling validation against experimental data without analytical simplifications. The temperature dependence of the catalytic rate constant k_{\text{cat}}, defined as the turnover frequency, follows the adapted : k_{\text{cat}} = A e^{-E_{a,\text{cat}}/RT} where A is the pre-exponential factor, E_{a,\text{cat}} is the lowered activation energy due to catalysis, R is the gas constant, and T is temperature; this results in k_{\text{cat}} being orders of magnitude larger than the uncatalyzed rate constant at typical operating temperatures, highlighting catalysis's kinetic enhancement./Kinetics/06%3A_Modeling_Reaction_Kinetics/6.02%3A_Temperature_Dependence_of_Reaction_Rates/6.2.03%3A_The_Arrhenius_Law/6.2.3.01%3A_Arrhenius_Equation)

Classification of Catalysis

Heterogeneous Catalysis

Heterogeneous catalysis occurs when the catalyst and reactants exist in different phases, most often involving a solid catalyst interacting with gaseous or liquid reactants at the of the phases. This setup leverages surface phenomena where reactions primarily take place on the catalyst's surface, enabling the of reaction rates without the catalyst being consumed. The inherent to heterogeneous systems allows for straightforward recovery and reuse of the catalyst through simple or , reducing operational costs and minimizing waste in large-scale processes. This characteristic makes heterogeneous catalysis dominant in industrial applications, accounting for over 80% of catalytic processes in chemical . A key feature of heterogeneous catalysts is the presence of active sites on the , which are specific locations where reactants adsorb and react. These sites can be coordinatively unsaturated atoms on metal surfaces, such as or , or defect sites on metal oxides like or zirconia, facilitating bond breaking and formation. To maximize the number of active sites and prevent aggregation, catalysts are often supported on high-surface-area materials, such as alumina (Al₂O₃), which disperses the active phase into fine particles, increasing accessibility and stability. For instance, in Ziegler-Natta , titanium-based active sites supported on enable the stereospecific of olefins like into , a process central to plastics production since its development in the . Observed reaction rates in heterogeneous catalysis can be influenced by diffusion limitations, where mass transfer of reactants to and products from the active sites becomes rate-controlling. External diffusion involves the transport of species from the bulk fluid phase to the external surface of the catalyst particle, often mitigated by increasing flow rates or reducing particle size. Internal diffusion, or pore diffusion, occurs within the porous structure of the catalyst, leading to concentration gradients that lower the effective rate, particularly in larger particles or highly exothermic reactions; this is quantified using effectiveness factors that compare observed and intrinsic kinetics. These limitations highlight the importance of catalyst design, such as optimizing pore size and particle morphology, to ensure surface reactions dominate. Catalyst deactivation is a common challenge in heterogeneous systems, arising from mechanisms like , where high temperatures cause active metal particles to , reducing surface area, or , the deposition of carbonaceous residues that block sites and pores. is thermally driven and often irreversible without specialized treatments, while results from side reactions in hydrocarbon processing. Regeneration methods, such as controlled burning off of coke deposits in an oxygen-containing atmosphere, can restore activity, though repeated cycles may lead to permanent loss; for example, catalysts are routinely regenerated this way to maintain performance. A seminal example of is the Haber-Bosch process for synthesis, where and gases react over an iron-based catalyst promoted with and alumina: \mathrm{N_2 + 3 H_2 \rightleftharpoons 2 NH_3} The iron catalyst provides active sites for dissociation, the rate-limiting step, operating at high pressures (150-300 atm) and temperatures (400-500°C) to achieve industrial yields exceeding 10-20% per pass. This process, developed in the early , produces over 150 million tons of annually, underscoring the scale of in production.

Homogeneous Catalysis

Homogeneous catalysis involves chemical where the catalyst is in the same phase as the reactants, typically a liquid solution, allowing for intimate molecular interactions that facilitate pathways. This uniformity enables precise control over dynamics, often leading to high reactivity and selectivity compared to heterogeneous systems. A key advantage of homogeneous catalysis is the ability to tune selectivity through modifications to the catalyst's ligands, which can alter electronic and steric properties to favor specific products or pathways. For instance, in complexes, ligand variations can direct or in bond-forming reactions. However, a major challenge is the separation of the catalyst from products and byproducts due to the shared phase, often requiring energy-intensive methods like or the design of biphasic systems where the catalyst partitions into a distinct layer post-reaction. Biphasic approaches, such as aqueous-organic or fluorous systems, mitigate this by enabling catalyst while maintaining homogeneous conditions during catalysis. In metal complex catalysis, , chlorotris(triphenylphosphine)rhodium(I) (RhCl(PPh₃)₃), exemplifies homogeneous of alkenes under mild conditions. This (I) complex activates dihydrogen and adds it across the C=C bond, converting substrates like RCH=CHR' to RCH₂CH₂R' with high efficiency and selectivity for terminal alkenes. The reaction proceeds at and , showcasing the precision of soluble organometallic catalysts. Acid and base catalysis in homogeneous media often relies on proton transfer mechanisms, where Brønsted or bases accelerate reactions by stabilizing charged intermediates. A classic example is the acid-catalyzed of esters, such as to acetic acid and , involving of the carbonyl oxygen to enhance electrophilicity, followed by addition and elimination steps. This process highlights how homogeneous acids, like in , lower activation barriers for without phase boundaries impeding diffusion. Organometallic cycles in frequently involve coordination steps like , migratory insertion, and , which enable multi-step transformations at a single metal center. occurs when a low-valent metal binds and cleaves a like H₂, increasing the metal's and to form a dihydride species. Migratory insertion then follows, where an alkyl or alkenyl shifts to the , forming an alkyl , often with π-acceptor ligands stabilizing the . completes the cycle by coupling the ligands to release the product, regenerating the low-valent catalyst and restoring its . These steps, common in cross-coupling and , allow for turnover numbers exceeding 10⁴ in optimized systems. Solvent polarity plays a crucial role in homogeneous catalysis by influencing reaction rates and selectivity through stabilization of transition states or charged species. Polar protic solvents, such as or alcohols, can enhance rates of ionic mechanisms like by solvating ions, while nonpolar solvents favor apolar substrates and may improve selectivity in organometallic insertions by reducing competing coordination. For example, in rhodium-catalyzed , switching from to more polar can shift regioselectivity by altering solvation and metal-substrate interactions.

Biocatalysis

Biocatalysis refers to the acceleration of chemical reactions using biological catalysts, primarily enzymes and whole microbial cells, which integrate the specificity of biological systems with fundamental chemical principles. Enzymes, typically proteins but occasionally molecules known as ribozymes, possess dedicated active sites where substrates bind and undergo transformation. These active sites are regions of precise three-dimensional structure that facilitate catalysis through proximity, orientation, and stabilization of transition states. The interaction between enzyme and substrate is classically described by the lock-and-key model, proposed by Emil Fischer in 1894, wherein the substrate's shape precisely matches the rigid active site, akin to a key fitting a lock, ensuring high specificity. This model was later refined by the induced fit hypothesis introduced by Daniel Koshland in 1958, which posits that the enzyme undergoes a conformational change upon substrate binding to achieve optimal alignment, enhancing catalytic efficiency and accommodating minor substrate variations. Many enzymes require coenzymes, non-protein organic molecules that act as transient carriers of chemical groups; for instance, nicotinamide adenine dinucleotide (NAD+) and flavin adenine dinucleotide (FAD) serve as electron acceptors and donors in redox reactions. A prominent example is alcohol dehydrogenase, which utilizes NAD+ to catalyze the oxidation of alcohols to aldehydes or ketones, playing a key role in ethanol metabolism. In industrial applications, biocatalysts are often engineered for enhanced performance, with immobilization techniques—such as in gels or attachment to solid supports—improving stability against denaturation, enabling reuse, and simplifying product separation. , a laboratory mimicry of involving iterative and screening of variants, optimizes properties like activity, , and specificity for commercial processes. Whole-cell biocatalysts, employing intact microorganisms, offer advantages in multi-enzyme cascades, as cellular compartments protect enzymes and provide endogenous cofactors. Key benefits of biocatalysis include exceptional enantioselectivity, allowing production of single stereoisomers vital for pharmaceuticals, and operation under mild aqueous conditions at ambient temperatures and neutral , which minimizes energy use and avoids harsh reagents.01283-9) A representative example is the use of lipases, serine hydrolases, to catalyze the esterification or of fatty acids with alcohols, such as in from vegetable oils or waste fats. Immobilized lipases, like those from Candida antarctica, achieve yields exceeding 90% under mild conditions (40–60°C, ), outperforming chemical catalysts by reducing byproducts and enabling feedstock flexibility. This approach exemplifies how biocatalysis enhances in renewable fuel synthesis.

Specialized Forms

Electrocatalysis

Electrocatalysis involves the acceleration of electrochemical reactions at surfaces by applying an electrical potential, where the acts as the catalyst to lower the barriers. In this process, the surface facilitates the transfer of electrons between the and reactants, enabling efficient energy conversion in systems like . A key challenge is the required to drive reactions such as the reaction (OER), where the catalyst reduces the extra voltage needed beyond the to achieve practical current densities, thereby improving overall efficiency. Common electrocatalytic materials are selected based on their ability to optimize reaction kinetics for specific half-reactions. For the (HER), represented as $2H^+ + 2e^- \rightarrow H_2, (Pt) serves as a catalyst due to its near-zero and high intrinsic activity in acidic media. Non-precious alternatives, such as nickel-iron (Ni-Fe) oxides, have emerged for the OER ($2H_2O \rightarrow O_2 + 4H^+ + 4e^-), exhibiting low overpotentials (around 300 mV at 10 mA/cm²) and enhanced stability through iron doping that modifies the electronic structure and active sites on the surface. These materials leverage heterogeneous surface catalysis principles, where adsorption energies of intermediates dictate performance. The , \eta = a + b \log(j), quantifies the relationship between (\eta) and (j), with the Tafel slope (b) providing insights into the and rate-determining step; for instance, slopes near 120 mV/dec indicate a Volmer-Heyrovsky pathway limited by recombination. Intrinsic activity is often assessed via the (i_0), which measures the rate at and highlights efficiency independent of mass loading—for Pt in HER, i_0 values reach approximately 1 mA/cm², underscoring its superior performance. Electrocatalysts find critical applications in fuel cells, where they enable oxygen reduction at the , and electrolyzers, facilitating through , with operational stability under potential cycling being essential to withstand voltage fluctuations over thousands of hours. For example, Pt-based catalysts in fuel cells maintain activity with minimal degradation (<10% loss) during accelerated stress tests simulating load changes. Ni-Fe oxides in alkaline electrolyzers demonstrate durability, retaining over 90% efficiency after 1000 cycles, addressing scalability for storage.

Photocatalysis

Photocatalysis involves the acceleration of chemical reactions through the absorption of by a catalyst, typically leading to the generation of reactive species that drive processes. In semiconductor-based , illumination with photons of energy greater than or equal to the material's (E_g) excites electrons from the valence band to the conduction band, creating electron-hole pairs (e⁻/h⁺). These charge carriers can migrate to the catalyst surface, where electrons reduce acceptors and holes oxidize donors, enabling reactions such as or pollutant degradation; the energy determines the required wavelength, with wider gaps like 3.0–3.2 eV for TiO₂ limiting activity to () . Semiconductor photocatalysts, particularly (TiO₂), have been pivotal in demonstrating photocatalytic potential. In the seminal Honda-Fujishima effect, UV irradiation of a TiO₂ in an splits into and oxygen via the reaction $2 \mathrm{H_2O} \rightarrow 2 \mathrm{H_2} + \mathrm{O_2}, with anodic oxidation producing O₂ and cathodic reduction yielding H₂, achieving stoichiometric gas evolution without external bias after initial setup. TiO₂'s phase, with its suitable band edge positions (conduction band ≈ -0.5 V vs. NHE, valence band ≈ +2.7 V), aligns well for potentials, though its wide restricts solar efficiency. Molecular photocatalysts, such as tris(2,2'-bipyridine) ([Ru(bpy)₃]²⁺), operate via metal-to-ligand charge transfer (MLCT) excitation under , forming a long-lived triplet ([Ru(bpy)₃]²⁺*). This excited complex acts as a strong reductant (E ≈ -1.33 V vs. ) or oxidant depending on , often in sacrificial systems where it donates electrons to acceptors like methyl viologen, regenerating via a donor such as EDTA. Such complexes enable homogeneous , bridging to applications in and . The core mechanisms in photocatalysis revolve around efficient charge separation to minimize recombination losses, where e⁻/h⁺ pairs annihilate rapidly (on timescales) if not separated, reducing to below 10% in many systems. Effective separation occurs via surface trapping, heterojunctions, or , leading to formation: holes generate hydroxyl radicals (•OH) for oxidation, while electrons produce (O₂⁻•) for . Recombination, either bulk or surface-mediated, remains a primary bottleneck, often addressed by doping or co-catalysts like to facilitate charge transfer. Applications of photocatalysis include environmental remediation through pollutant degradation, where TiO₂ under UV light mineralizes organic dyes and pesticides to CO₂ and H₂O, achieving up to 95% removal of compounds like in aqueous suspensions via •OH attack on chromophores. For CO₂ reduction, photocatalysts convert CO₂ to value-added fuels like CO or CH₄ using H₂O as the , with quantum yields reaching 0.1–1% under visible light in optimized systems. Z-scheme configurations, inspired by , couple two semiconductors (e.g., TiO₂ with BiVO₄) where conduction band electrons from one recombine with valence band holes of the other, preserving high potentials for simultaneous CO₂ and water oxidation.

Organocatalysis

Organocatalysis refers to the acceleration of chemical reactions by small organic molecules that operate without metal centers, often mimicking the activation strategies of enzymes but using synthetically accessible, metal-free catalysts. These catalysts typically function through non-covalent interactions such as hydrogen bonding or covalent mechanisms like nucleophilic or electrophilic , enabling precise control over reaction pathways. In asymmetric organocatalysis, chiral organic molecules induce , producing enantioenriched products essential for pharmaceuticals and fine chemicals. Key principles of organocatalysis involve hydrogen bonding to activate electrophiles by stabilizing transition states or coordinating substrates, as well as nucleophilic activation where the catalyst forms transient covalent bonds with reactants to enhance reactivity. For instance, electrophilic activation via hydrogen bonding lowers the energy barrier for nucleophilic attack, while nucleophilic catalysts like amines add to carbonyls to generate activated intermediates. Chiral organocatalysts, often derived from or simple heterocycles, enforce asymmetry by creating diastereomeric transition states, leading to high enantioselectivities in reactions such as C-C bond formations. Prominent types of organocatalysts include derivatives for s and (DMAP) for acyl transfers. acts as a bifunctional catalyst, forming an with ketones to activate them as nucleophiles in the aldol addition to aldehydes, achieving enantioselectivities up to 99% in the direct asymmetric between acetone and various aldehydes. This was demonstrated in seminal work showing 's efficiency in promoting intermolecular aldolizations under mild conditions. Similarly, DMAP functions as a nucleophilic catalyst in acylation reactions, forming an acylpyridinium intermediate that facilitates rapid or formation from alcohols or amines and anhydrides, with turnover numbers exceeding 10^4 in some cases. Introduced as a superior alternative to , DMAP enables efficient group transfers in both achiral and chiral contexts when modified with stereogenic centers. Organocatalysis offers significant advantages in , including low toxicity due to the absence of , straightforward synthesis from abundant feedstocks, and elimination of metal residues in products, which simplifies purification and reduces environmental impact. These catalysts are often air- and moisture-stable, operable in aqueous or benign solvents, and recyclable in many protocols, aligning with principles of by minimizing waste and energy use. For example, in industrial scouting, organocatalytic processes have demonstrated E-factors below 10 for asymmetric syntheses, far superior to metal-catalyzed analogs. A landmark example is the use of MacMillan's imidazolidinone catalyst in the Diels-Alder reaction, where the chiral condenses with an α,β-unsaturated to form an intermediate that accelerates the with dienes, yielding cycloadducts with up to 99% and endo selectivity. This im/ activation strategy expanded organocatalysis to pericyclic reactions, enabling asymmetric synthesis of complex polycycles. The field experienced a surge in asymmetric organocatalysis after 2000, spurred by independent reports from List on proline-mediated aldol reactions and MacMillan on iminium-catalyzed Diels-Alder processes, which demonstrated broad applicability and high stereocontrol without metals. This renaissance led to over 20,000 publications by 2020, integrating organocatalysis into total syntheses and industrial processes, and culminated in the 2021 for List and MacMillan. The post-2000 developments emphasized multifunctional catalysts and hybrid activations, transforming organocatalysis into a cornerstone of sustainable synthetic chemistry.

Applications and Significance

Industrial Production

Industrial production of chemicals heavily relies on catalytic processes to synthesize bulk and fine chemicals efficiently on a large scale. The Haber-Bosch process exemplifies this for , utilizing an iron-based catalyst at pressures of 150-200 atm and temperatures around 400°C to convert and hydrogen into , yielding approximately 180 million metric tons annually as of 2023 to support global needs. Similarly, the produces by oxidizing to over a vanadium pentoxide catalyst, enabling the manufacture of this essential industrial chemical used in fertilizers, batteries, and detergents. In fine chemicals, particularly pharmaceuticals, employs complexes with DuPHOS ligands to achieve high enantioselectivity, facilitating the synthesis of chiral intermediates with minimal byproducts and supporting scalable production of single-enantiomer drugs. in industrial catalysis varies by type: often uses fixed-bed reactors where catalyst pellets remain stationary, allowing continuous flow of reactants through the bed for gas-phase reactions like ammonia synthesis. In contrast, typically employs continuous stirred-tank reactors, ensuring uniform mixing of soluble catalysts and reactants for liquid-phase processes such as . Catalysts underpin over 90% of chemical processes, driving economic value through higher yields, reduced , and lower , which collectively lower costs and enhance competitiveness in the global . Sustainability efforts in industrial catalysis increasingly incorporate enzymes for processing bio-based feedstocks, such as , to produce platform chemicals like , reducing reliance on and minimizing environmental footprints through milder conditions and renewable resources.

Environmental Impact

Catalysis plays a pivotal role in mitigating environmental pollution and promoting sustainable processes by enabling efficient chemical transformations that reduce emissions and waste. In control, (SCR) using vanadium pentoxide (V₂O₅)-based catalysts has become a standard technology in power plants to remove (), which contribute to and formation. The process involves the reaction of with over V₂O₅ supported on (TiO₂), typically promoted with tungsten oxide (WO₃), operating at temperatures around 300–400°C to achieve over 90% conversion. The key reaction is: $4 \mathrm{NH_3} + 4 \mathrm{NO} + \mathrm{O_2} \rightarrow 4 \mathrm{N_2} + 6 \mathrm{H_2O} This method has been widely adopted since the 1970s, significantly lowering emissions from stationary sources like coal-fired boilers. In , with (TiO₂) offers an effective approach for degrading organic pollutants, such as dyes from textile effluents, which can harm aquatic ecosystems by reducing oxygen levels and causing toxicity. Under light, TiO₂ generates that mineralize dyes like into harmless CO₂ and H₂O, with degradation efficiencies exceeding 95% in many lab-scale studies. Modifications, such as doping with or metals, extend TiO₂'s activity to visible light, enhancing its practicality for large-scale remediation without secondary . For management, catalysis facilitates CO₂ capture and conversion into valuable products, addressing by utilizing this abundant waste gas. Copper-zinc oxide (Cu/ZnO) catalysts, often supported on alumina, enable the of CO₂ to at moderate pressures (50–100 bar) and temperatures (200–300°C), with selectivities up to 80% for . This process not only sequesters CO₂ from industrial flue gases but also produces a clean fuel or chemical feedstock, reducing reliance on fossil-derived . Catalysis also advances green chemistry principles by enabling atom-efficient reactions that minimize waste. Olefin metathesis, for instance, exemplifies high atom economy—often approaching 100%—as it rearranges carbon-carbon double bonds in alkenes with minimal byproducts, primarily ethylene, using ruthenium or molybdenum catalysts. This reaction supports sustainable synthesis of pharmaceuticals and polymers, aligning with the goal of waste prevention in chemical manufacturing. However, environmental challenges persist, including catalyst leaching, where active metal species dissolve into reaction media and potentially enter ecosystems, causing bioaccumulation and toxicity in soil and water. Strategies like immobilization on stable supports are essential to mitigate these risks and ensure long-term ecological safety.

Biological and Food Processing

In biological systems, catalysis plays a pivotal role in , enabling efficient energy production and processes essential for . Cytochrome P450 enzymes, a superfamily of heme-containing monooxygenases, catalyze the oxidation of xenobiotics and endogenous compounds, facilitating in the liver and other tissues. These enzymes introduce oxygen atoms to substrates, converting lipophilic toxins into water-soluble metabolites for excretion, thus preventing cellular damage from drugs, pollutants, and dietary components. In parallel, —a foundational —relies on a series of enzymes to convert glucose into pyruvate, generating ATP and NADH under conditions. Key catalysts include , which phosphorylates glucose in the initial step, and phosphofructokinase-1, which regulates flux through irreversible of fructose-6-phosphate, ensuring rapid energy mobilization in cells like erythrocytes and muscle fibers. In food processing, enzymatic catalysis enhances flavor, texture, and preservation while minimizing energy use compared to thermal methods. During yeast fermentation, invertase (β-fructofuranosidase) secreted by Saccharomyces cerevisiae hydrolyzes sucrose into glucose and fructose in the periplasmic space, fueling alcoholic fermentation for products like bread, beer, and wine. This glycoside hydrolase operates optimally at acidic pH and moderate temperatures, yielding equimolar monosaccharides that yeast rapidly metabolizes to ethanol and carbon dioxide. Similarly, in cheese production, chymosin (also known as rennin), an aspartic protease from calf stomachs or recombinant sources, catalyzes the specific hydrolysis of κ-casein at the Phe105-Met106 bond, destabilizing milk micelles to form curds. This coagulation step is crucial for separating whey from solids, with recombinant chymosin enabling consistent yields and vegetarian-compatible variants without altering cheese quality. Biocatalysis extends to pharmaceutical manufacturing, where enzymes streamline synthesis of complex molecules like statins, cholesterol-lowering drugs. hydrolase, a serine from , selectively cleaves the 2-methylbutyryl ester from lovastatin to produce monacolin J, a key intermediate for simvastatin. This enzymatic deacylation achieves high and efficiency under mild aqueous conditions, reducing waste and enabling scalable production for global supply. Such processes exemplify how biocatalysts, building on fundamentals from the broader field of biocatalysis, replace multi-step chemical routes with greener alternatives. Catalysis also underpins by aiding and absorption in the . Salivary and pancreatic hydrolyzes starches into and dextrins, initiating breakdown in the mouth and for . Proteases, including in the stomach and in the , cleave bonds in proteins to release , supporting muscle repair and . Deficiencies in these enzymes can impair , highlighting their role in maintaining dietary health. Emerging applications leverage engineered microbes for sustainable , bridging biological catalysis with production akin to processes. Metabolic of yeasts and bacteria, such as and , introduces synthetic pathways to convert lignocellulosic sugars into advanced biofuels like and with yields exceeding 90% of theoretical maxima. These genetically modified strains enhance tolerance to inhibitors and optimize cascades, potentially reducing dependence while drawing on principles used in industries.

Historical Development

Early Observations

The earliest known observations of catalytic phenomena trace back to ancient civilizations, particularly in the production of fermented foods and beverages. Around 5000 BCE, ancient Egyptians utilized natural processes for and leavened , where acted as an unrecognized biocatalyst to convert sugars into and . These empirical practices demonstrated acceleration of biochemical reactions without awareness of the underlying mechanisms, marking the inadvertent harnessing of biocatalysis in human society. In the , scientific inquiry began to uncover more deliberate examples of catalytic effects. During the 1770s, English chemist conducted experiments showing that a sprig of placed in a sealed container with "dephlogisticated air" (oxygen) depleted by could restore the air's ability to support or burning after several days. This observation, later linked to , highlighted how living organisms could catalyze the renewal of atmospheric gases essential for and life. The formal conceptualization of catalysis emerged in the early . In 1811, Russian chemist Gottlieb Sigismund Kirchhoff demonstrated that heating in dilute produced a sweet containing glucose, illustrating acid-catalyzed as a non-biological acceleration of . This experiment provided a key inorganic example, showing how small amounts of acid facilitated the breakdown of complex carbohydrates into simpler sugars without being consumed. In 1835, Swedish chemist coined the term "catalysis" from the Greek "katalysis" (meaning dissolution or loosening), defining it as a process where a foreign substance invigorates a slumbering while remaining unchanged. Berzelius applied this to both inorganic and organic reactions, including those involving ferments. Nineteenth-century developments further solidified catalysis as a chemical principle amid philosophical debates. German chemist Justus von Liebig promoted the idea of "organic catalysis," arguing that processes like fermentation in living systems resulted from contact actions between organic substances, akin to inorganic catalysts, rather than requiring a mystical vital force. Liebig's views positioned catalysis as a unifying chemical phenomenon applicable to both lifeless and vital processes, influencing agricultural and physiological chemistry. However, these ideas fueled controversies with vitalism, a doctrine asserting that organic reactions demanded a unique life force inaccessible to purely chemical explanations; debates intensified over fermentation, pitting mechanistic interpretations against vitalistic ones, as seen in Liebig's exchanges with Louis Pasteur. These early disputes underscored the tension between empirical observations and theoretical frameworks in establishing catalysis as a rigorous scientific concept.

Modern Advances

The formalization of catalysis as a scientific discipline advanced significantly in the early with Wilhelm Ostwald's contributions to its principles and industrial applications, earning him the 1909 . Ostwald's work established key laws governing catalytic action, emphasizing how catalysts accelerate reaction rates without being consumed, and extended these concepts to practical processes like the for production. Mid-20th century breakthroughs in catalysis came with the independent discoveries of and in the 1950s, who developed titanium-based Ziegler-Natta catalysts enabling stereospecific of olefins into and isotactic . Their innovations transformed plastics manufacturing by allowing precise control over polymer microstructure, for which they shared the 1963 . The 1960s and 1970s marked the rise of , exemplified by Geoffrey Wilkinson's development of chlorotris(triphenylphosphine)rhodium(I), known as , first reported in 1966 for efficient under mild conditions. This square-planar complex revolutionized selective reductions in , paving the way for soluble metal catalysts that offered mechanistic insights unattainable with heterogeneous systems. Building on this, the late saw asymmetric catalysis flourish, with William S. Knowles, Ryoji Noyori, and K. Barry Sharpless awarded the 2001 for chiral catalysts enabling enantioselective hydrogenations and oxidations, crucial for producing pure enantiomers in pharmaceuticals. A major milestone in biocatalysis came with the 2018 Nobel Prize in Chemistry awarded to Frances H. Arnold, George P. Smith, and Sir Gregory P. Winter for the of enzymes and development of techniques, which enabled the creation of customized biocatalysts with improved efficiency and specificity for industrial applications. In the 2010s and 2020s, biocatalysis advanced further through CRISPR-Cas9-mediated , allowing precise to engineer enzymes with enhanced stability and activity for industrial processes like production. This gene-editing tool facilitated high-throughput variant screening, yielding optimized biocatalysts that outperform traditional chemical methods in specificity. Concurrently, single-atom catalysts (SACs) emerged as a paradigm in , featuring isolated metal atoms on supports to maximize atom efficiency and tunability, with applications in oxygen reduction reactions showing turnover frequencies exceeding 10 times those of counterparts. Computational methods transformed catalyst design from the onward, with (DFT) enabling detailed modeling of active sites on surfaces and in complexes. Early applications of DFT to catalysis, such as adsorbate interactions on metal surfaces, provided quantitative predictions of reaction barriers and selectivity, accelerating the rational design of heterogeneous catalysts.

Advanced Concepts

Inhibitors and Promoters

In catalysis, , often referred to as , are substances that reduce the activity of a catalyst by strongly adsorbing onto active sites, thereby blocking access for reactants. This adsorption can occur through , where the inhibitor forms a bond with the catalyst surface, leading to deactivation. A classic example is of catalysts in reforming processes, where species like irreversibly chemisorb onto sites, preventing activation and necessitating feed pretreatment to remove below 1 . Such irreversible contrasts with temporary inhibitors that adsorb weakly and can desorb under reaction conditions, allowing partial recovery of activity. Promoters, on the other hand, are additives that enhance catalytic performance by modifying the electronic or geometric properties of the active sites. Electronic promoters alter the electronic structure of the catalyst, such as by donating electrons to increase metal and weaken reactant adsorption bonds. For instance, (K₂O) acts as an electronic promoter in iron-based catalysts for , facilitating by enhancing to the iron surface. Geometric promoters, by contrast, influence the arrangement of surface atoms, creating ensembles that favor specific reaction pathways or dispersing active sites to prevent . These effects can be synergistic, as seen in multi-promoted systems where both types optimize turnover frequencies without blocking sites. Inhibitors and promoters can be distinguished by their adsorption nature: temporary inhibitors involve reversible , akin to competitive adsorption in , while permanent poisons rely on irreversible that deactivates sites indefinitely. effects from poisons, such as charge transfer that modifies d-band centers, can exacerbate deactivation, whereas geometric effects involve site blocking that reduces ensemble sizes for multi-atom reactions. In , these additives often lead to deactivation patterns observed in long-term operation, influencing overall . Mitigation strategies for poisoning include alloying the catalyst with metals that resist strong adsorption or provide alternative sites for poison sequestration. For example, in proton exchange membrane fuel cells, carbon monoxide poisoning of platinum anodes—where CO binds strongly to undercoordinated Pt sites, reducing hydrogen oxidation efficiency—is alleviated by alloying with ruthenium, which promotes CO oxidation via bifunctional mechanisms at lower potentials. Promoter optimization through controlled doping, such as varying K₂O loading in iron catalysts, fine-tunes electronic density to maximize activity while minimizing over-promotion that could lead to instability. These approaches, including surface engineering via alloying, enhance poison resistance and extend catalyst lifetimes in industrial applications.

Prebiotic Role

In the context of , catalysis played a pivotal role in facilitating the and of prebiotic biomolecules on , bridging simple geochemical precursors to complex self-replicating systems. surfaces and emerging molecules likely accelerated reactions that were otherwise kinetically unfavorable under aqueous, energy-limited conditions, enabling the accumulation of life's building blocks such as and . This prebiotic catalysis is hypothesized to have occurred in environments like hydrothermal vents and mineral-rich pools, where heterogeneous and homogeneous mechanisms promoted the formation of oligomers without enzymatic intervention. The RNA world hypothesis proposes that RNA molecules functioned dually as genetic carriers and catalysts, with enabling and the emergence of functional polymers. , such as variants, can synthesize copies of themselves and complementary strands from template-directed monomers, mimicking the autocatalytic processes essential for in prebiotic settings. selections have demonstrated ribozyme ligases capable of using prebiotically plausible substrates like 2-aminoimidazole-activated , supporting a transition from non-enzymatic to RNA-catalyzed replication. However, achieving sustained replication required overcoming hydrolysis-prone linkages, with experiments showing that short ribozymes (as few as 50 ) could catalyze under mild aqueous conditions. Mineral surfaces provided heterogeneous catalytic sites for and assembly, concentrating reactants and lowering activation energies. clay, a abundant in prebiotic sediments, catalyzes the regioselective of activated ribonucleotides into oligomers up to 50 units long in aqueous solutions at ambient temperatures, with yields enhanced by interlayer adsorption that protects monomers from degradation. Similarly, (FeS) minerals in alkaline hydrothermal vents facilitate carbon fixation and reactions, such as the of CO₂ to and the conversion of pyruvate to metabolic precursors, simulating protometabolic cycles under conditions with gradients of and . These sulfides, including mackinawite, promote and at rates comparable to modern enzymes, potentially compartmentalizing reactions within porous structures. Peptide formation via condensation was similarly aided by mineral catalysis, particularly on (FeS₂) surfaces, which activate carboxyl groups through adsorption and proton transfer, enabling s in aqueous media despite thermodynamic hurdles. Experiments indicate that surfaces accelerate formation from such as under simulated prebiotic conditions, including temperatures up to 150°C and geochemical gradients, yielding short like dipeptides. Extensions of the Miller-Urey experiment incorporating catalytic minerals, such as iron-rich meteorites, have demonstrated enhanced yields of organics like and nucleobases from CO₂ and N₂ under spark discharge or volcanic conditions, yielding significantly higher amounts (over 200 times more) of organic compounds, including key prebiotic molecules such as and nucleobases, than uncatalyzed runs. Despite these advances, prebiotic catalysis faced significant challenges, including the chemical instability of RNA oligomers under hydrolytic and UV-exposed conditions, which limited chain lengths and fidelity in replication. Non-enzymatic RNA synthesis suffers from high error rates (up to 10⁻¹ per ), hindering the of complex functions, while the transition to protein enzymes required hybrid RNA-peptide systems to stabilize catalysis and expand substrate specificity. Modified , such as those with 2-thio or 8-aza substitutions, may have mitigated degradation, allowing longer, more stable ribozymes to emerge before the dominance of DNA-protein paradigms. Ongoing experiments highlight that while mineral catalysts enable initial oligomerization, achieving enzymatic demanded selective pressures favoring robust, cooperative networks. Recent studies (as of 2025) have further illuminated prebiotic catalysis. For instance, have been shown to catalyze RNA formation from under ambient dry alkaline conditions, suggesting organic molecules could drive early without minerals. Additionally, the facilitates reproducible oligomerization of Ala-Gly dipeptides in prebiotic simulations, highlighting diverse mineral roles in assembly.

References

  1. [1]
    catalyst (C00876) - IUPAC Gold Book
    A substance that increases the rate of a reaction without modifying the overall standard Gibbs energy change in the reaction; the process is called catalysis.
  2. [2]
    The history of biocatalysis - Nobel Prize
    The history of biocatalysis ; 1835, The Swede Jöns Jacob Berzelius describes a catalyst as a substance which can breathe life into slumbering chemical reactions.
  3. [3]
    1: History of Catalysis - Books - The Royal Society of Chemistry
    May 11, 2017 · The name 'catalysis' was coined by Berzelius in 1836.2 He concluded that besides 'affinity', a new force was operative, a 'Catalytic Force'. The ...History of Catalysis... · Petrochemical Industry · Environmental Catalysis
  4. [4]
    History of Catalysis - Smith - Major Reference Works
    Sep 15, 2010 · In 1836, Berzelius coined the word catalysis to generalize a growing body of experimental data. A half century later, Wilhelm Ostwald linked ...<|separator|>
  5. [5]
    1. An Introduction to Types of Catalysis - Chemistry LibreTexts
    Jun 30, 2023 · Catalysis includes heterogeneous (catalyst in different phase), homogeneous (catalyst in same phase), and autocatalysis (catalyzed by a product ...
  6. [6]
    Towards more efficient catalysts
    Feb 16, 2024 · The research is published in Nature Catalysis. The chemical industry relies on catalysts for over 90 percent of its processes and nearly all ...
  7. [7]
    Catalysis in industry
    Catalysts are substances that speed up reactions by providing an alternative pathway for the breaking and making of bonds.
  8. [8]
    Catalysis making the world a better place - PMC - PubMed Central
    Catalysis is involved at some point in the processing of over 80% of all manufactured products. Hence catalysis is immensely important not only because it is an ...
  9. [9]
    Catalysis | PNNL
    Catalysis can be generally defined into three broad categories: homogeneous catalysis, heterogeneous catalysis, and biocatalysis. In homogeneous catalysis ...
  10. [10]
    The Effect of a Catalyst on Rate of Reaction - Chemistry LibreTexts
    Jun 30, 2023 · A catalyst provides an alternative route for the reaction with a lower activation energy. It does not "lower the activation energy of the ...
  11. [11]
    Effect of catalysts - Chemistry 302
    The effect of a catalyst is that it lowers the activation energy for a reaction. Generally, this happens because the catalyst changes the way the reaction ...
  12. [12]
    DOE Explains...Catalysts - Department of Energy
    Catalysts make this process more efficient by lowering the activation energy, which is the energy barrier that must be surmounted for a chemical reaction to ...
  13. [13]
    “Turning Over” Definitions in Catalytic Cycles | ACS Catalysis
    Nov 8, 2012 · “Turnover number (TON) specifies the maximum use that can be made of a catalyst for a special reaction under defined reaction conditions by the ...Introduction · “Turning Over” Definitions · Beyond the Turnover Measures
  14. [14]
    6.2.3.1: Arrhenius Equation - Chemistry LibreTexts
    Feb 13, 2024 · In addition, the Arrhenius equation implies that the rate of an uncatalyzed reaction is more affected by temperature than the rate of a ...
  15. [15]
    Hydrogen peroxide decomposition using different catalysts
    From fresh liver, to powdered manganese, create different catalysts to explore the effervescent world of hydrogen peroxide decomposition.
  16. [16]
    [PDF] Chemical Transformations II: Decomposition with H2O2 with KI
    KI catalyzes the decomposition of hydrogen peroxide to oxygen and water. ... Therefore, the rate of catalyzed is 2439 times the rate of the uncatalyzed.
  17. [17]
    Hydrogen Peroxide Decomposition by Iodide
    30% Hydrogen peroxide in a large round bottom flask decomposes rapidly to hissing steam and oxygen when potassium iodide is added.
  18. [18]
    The evolution of catalytic converters | Feature | RSC Education
    May 31, 2011 · The TWC comprises platinum or palladium to catalyse oxidation reactions, and rhodium for reduction of the NOx, and was made possible due to ...
  19. [19]
    turnover frequency (T06534) - IUPAC Gold Book
    Commonly called the turnover number, , and defined, as in enzyme catalysis, as molecules reacting per active site in unit time.Missing: activity TON TOF
  20. [20]
  21. [21]
    Biocatalysis: Enzymatic Synthesis for Industrial Applications - Wu
    Jun 18, 2020 · Space-time-yield (STY, g L−1 h−1). Catalyst consumption/load (i.e., g enzyme kg−1 product). Looking at the sheer numbers, one can quickly ...
  22. [22]
    Catalyst Selectivity - an overview | ScienceDirect Topics
    Catalyst selectivity is defined as the ability of a catalyst to preferentially facilitate the conversion of a specific reactant into a desired product ...
  23. [23]
    Yield | Encyclopedia MDPI
    Oct 18, 2022 · According to the Elements of Chemical Reaction Engineering manual, yield refers to the amount of a specific product formed per mole of reactant ...
  24. [24]
    Pure and Applied Chemistry - iupac
    Defining the kind-of-quantity "catalytic activity" as a property of the catalyst measured by the catalyzed rate of conversion, the coherent SI unit is mole per ...
  25. [25]
    MANUAL ON CATALYST CHARACTERIZATION - iupac
    In heterogeneous catalysis a fluid (liquid or gas) is brought into contact ... The determination of specific quantities such as catalytic activity from ...
  26. [26]
    Comparative Study of Batch and Continuous Flow Reactors ... - MDPI
    A continuous flow reactor is a steady-state reactor, in which the reagents are fed constantly to the inlet of the reactor and move through the catalyst bed.
  27. [27]
    Application of reactor engineering concepts in continuous flow ...
    May 4, 2021 · In this review we highlight the application of new reactor configurations and designs in the context of either bespoke or commercial flow apparatus.
  28. [28]
    In Situ Infrared Spectroscopy as a Tool for Monitoring Molecular ...
    Apr 1, 2019 · This study showed that infrared measurements in transmission mode are able to detect active catalytic species and can follow deactivation phenomena.
  29. [29]
    Magnetic Resonance Methods in Catalysis
    Sep 12, 2024 · A well-established tool to probe the structure and operation of catalysts. It is efficient for both homogeneous and heterogeneous catalysis.
  30. [30]
    THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS ...
    Article September 1, 1918. THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS, MICA AND PLATINUM. Click to copy article linkArticle link copied! Irving Langmuir.
  31. [31]
    The mechanism of activation at catalytic surfaces - Journals
    Hinshelwood and Pritchard, in considering this problem, believed that adsorption of the reactant might increase as often as it decreased the stability of the ...
  32. [32]
    The catalysis of the parahydrogen conversion by tungsten - Journals
    The parahydrogen conversion has been measured on a tungsten surface, as a function of the surface concentration of adsorbed oxygen, acting as a poison.
  33. [33]
    Mechanism and Stereoselectivity of Asymmetric Hydrogenation
    Rhodium complexes containing chiral phosphine ligands catalyze the hydrogenation of olefinic substrates such as α-aminoacrylic acid derivatives.
  34. [34]
    The Central Role of Enzymes as Biological Catalysts - The Cell - NCBI
    Energy diagrams for catalyzed and uncatalyzed reactions. The reaction illustrated is the simple conversion of a substrate S to a product P. Because the final ...Missing: profile | Show results with:profile
  35. [35]
    6.4: Catalysis - Chemistry LibreTexts
    Dec 27, 2022 · In other words, a catalyst affects the kinetics of a reaction, but not the thermodynamics. Catalysts play a hugely important role in biochemical ...Missing: unchanged | Show results with:unchanged
  36. [36]
    A5. Transition State Stabilization - Chemistry LibreTexts
    May 8, 2019 · Linus Pauling postulated long ago that the only thing that a catalyst must do is bind the transition state more tightly than the substrate.
  37. [37]
    Le Chatelier's Principle Fundamentals - Chemistry LibreTexts
    Jan 29, 2023 · For reversible reactions, the enthalpy value is always given as if the reaction was one-way in the forward direction. The back reaction (the ...
  38. [38]
    Phenomena Affecting Catalytic Reactions at Solid–Liquid Interfaces
    Solvents can affect the rates of catalytic reactions on surfaces by stabilizing or destabilizing surface-bound intermediates and transition states. (32-36) In a ...
  39. [39]
    6.4.1: Eyring equation - Chemistry LibreTexts
    Feb 12, 2023 · The Eyring Equation, developed by Henry Eyring in 1935, is based on transition state theory and is used to describe the relationship between reaction rate and ...
  40. [40]
    [PDF] Chemical Kinetics
    k' (pseudo-first-order reaction). Finding k' immediately allows to find k: k ... • The most impressive examples of homogeneous catalysis occur in nature ...
  41. [41]
    [PDF] Kinetic Measurements in Heterogeneous Catalysis
    Equation (3) can be written as: (5) where is the space time [kg s mol-1]. The space time is proportional to the average nominal residence.<|separator|>
  42. [42]
    Mechanism of Action Assays for Enzymes - NCBI
    May 1, 2012 · Without a mechanism to clear the substrate, a competitive inhibitor will lose potency. This is not the case for a noncompetitive inhibitor.
  43. [43]
    Kinetic modeling of homogeneous catalytic processes - ScienceDirect
    In this paper, a brief review of kinetic modeling in homogeneous catalysis has been presented. Approaches using empirical as well as molecular level rate ...
  44. [44]
    [PDF] Heterogeneous Catalysis in Organic Chemistry Baran Group Meeting
    Mar 1, 2024 · Heterogeneous catalysis involves catalysts in a different phase than reactants/products, and is used in over 80% of chemical production. It can ...
  45. [45]
    Combining the Benefits of Homogeneous and Heterogeneous ... - NIH
    Heterogeneous catalysis is widely used in industrial applications because of the facile separation, which often results in lower operating costs. On the other ...Missing: characteristics | Show results with:characteristics
  46. [46]
    A Practical Guide to Heterogeneous Catalysis in Hydrocarbon ...
    Supports like activated carbon or alumina disperse the metals, increasing active surface area. Oxidation Reactions: Metal oxides or mixed‐metal oxides ...
  47. [47]
    Parallel Catalyst Synthesis Protocol for Accelerating Heterogeneous ...
    Dec 17, 2023 · Heterogeneous Ziegler-Natta catalysts play a crucial role in the industrial production of polyolefins. They function at two distinct scales: the ...
  48. [48]
    [PDF] Metal Oxides in Heterogeneous Oxidation Catalysis: State of the Art ...
    Metal oxides may also be used simply as supports of active phases, for example, silica, alumina, silica-alumina, mesoporous oxides, hierarchical porous oxides, ...
  49. [49]
    [PDF] A Review of Mass Transfer Controlling the Reaction Rate in ...
    Jul 7, 2011 · In this chapter, the influences of the mass transfer, i.e. external and internal diffusion, controlling the rate of heterogeneous catalysis ...
  50. [50]
    [PDF] Reaction Chemistry & Engineering - UU Research Portal
    Mar 26, 2024 · Internal mass transport limitations describe the limitations of diffusion within a catalyst grain, as a gradient is induced from the surface to ...
  51. [51]
    [PDF] Heterogeneous Catalyst Deactivation and Regeneration: A Review
    Feb 26, 2015 · Abstract: Deactivation of heterogeneous catalysts is a ubiquitous problem that causes loss of catalytic rate with time.
  52. [52]
    [PDF] Regeneration of catalysts deactivated by coke deposition: A review
    Fortunately, catalyst deactivation due to coke deposition is usually reversi- ble, and the coke can easily be removed by oxidation with air (O2) [3,8,11].
  53. [53]
    [PDF] Gerhard Ertl - Nobel Lecture
    conclusion that the rate-limiting step in ammonia synthesis over iron catalysts is the chemisorption of nitrogen. The question as to whether the nitrogen ...
  54. [54]
    [PDF] Sustainable Ammonia Synthesis DOE Roundtable Report - OSTI.GOV
    As compared in Figure 5, the Haber-. Bosch process (left) consists of several steps: (1) adsorbing and splitting of dinitrogen while providing dihydrogen and (2) ...
  55. [55]
    18.7 Catalysis – Chemistry Fundamentals
    Catalysts affect the rate of a chemical reaction by altering its mechanism to provide a lower activation energy. Catalysts can be homogenous (in the same ...
  56. [56]
    Ligand Effects in Homogeneous Au Catalysis | Chemical Reviews
    This review will address ligand effects in homogeneous catalysis through December 2007. Obvious differences in reactivity and selectivity arising from changing ...
  57. [57]
    Recycling of Homogeneous Catalysts Basic Principles, Industrial ...
    Apr 25, 2024 · The separation of products and catalysts is always incomplete, which means that not only some of the catalyst can be lost to the product phase ...
  58. [58]
    Separation/recycling methods for homogeneous transition metal ...
    This review aims to provide the reader with an overview of the current literature on the separation/recycling methods of homogeneous transition metal catalysts ...
  59. [59]
    Journal of the Chemical Society A - RSC Publishing
    J. A. Osborn, F. H. Jardine, J. F. Young and G. Wilkinson, J. Chem. Soc. A, 1966, 1711 DOI: 10.1039/J19660001711. To request permission to reproduce material ...
  60. [60]
    Mechanisms of ester hydrolysis in aqueous sulfuric acids
    Different kinetic behaviors for unimolecular and bimolecular ester hydrolysis reactions in strongly acidic microemulsions.
  61. [61]
    Solvent effects in catalysis: rational improvements of catalysts via ...
    Feb 17, 2016 · Solvents interact directly with the catalyst, the substrates and products, and all these interactions can increase or decrease reaction rate and/or selectivity.
  62. [62]
    Enzymes: principles and biotechnological applications - PMC
    This chapter covers the basic principles of enzymology, such as classification, structure, kinetics and inhibition, and also provides an overview of industrial ...
  63. [63]
    Looking Back: A Short History of the Discovery of Enzymes and How ...
    Sep 9, 2020 · ... enzymes worked. Through the famous lock-and-key model, proposed by Emil Fischer in 1894, as well as the Michaelis-Menten model of enzyme ...
  64. [64]
    Yeast Alcohol Dehydrogenase Structure and Catalysis | Biochemistry
    Aug 26, 2014 · The H44R substitution in yeast ADH1 decreased the dissociation constants for NAD+ and NADH by 2–4-fold and decreased turnover numbers by 4–6- ...
  65. [65]
    Industrial applications of immobilized enzymes—A review
    For all enzymes, the possibility to be immobilized and used in a heterogeneous form brings important industrial and environmental advantages, such as simplified ...
  66. [66]
    Role of Biocatalysis in Sustainable Chemistry | Chemical Reviews
    The biocatalytic process is also more cost-effective and eliminated the use of noble metals. Moreover, the process is conducted in standard multipurpose ...
  67. [67]
    Lipase‐catalyzed production of biodiesel: a critical review on ...
    Dec 7, 2023 · Biodiesel is a green and renewable alternative energy source which can be produced through lipase-catalyzed transesterification process.Introduction · Lipases for biodiesel production · Process factors affecting lipase...
  68. [68]
    Biodiesel production through lipase catalyzed transesterification
    Jan 2, 2010 · This paper is meant to review the latest development in the field of lipase catalyzed transesterification of biologically derived oil to produce biodiesel.
  69. [69]
    Introduction to Electrocatalysts | ACS Symposium Series
    Dec 9, 2022 · An electrocatalyst is a surface where chemical energy is electrochemically converted into electrical energy in fuel cells.
  70. [70]
    Surface transformations of electrocatalysts during the oxygen ...
    May 24, 2023 · The sluggish reaction kinetics of the anodic oxygen evolution reaction (OER) represent the largest source of efficiency loss (or overpotential) ...
  71. [71]
    Platinum single-atom and cluster catalysis of the hydrogen evolution ...
    Nov 30, 2016 · Platinum-based catalysts have been considered the most effective electrocatalysts for the hydrogen evolution reaction in water splitting.
  72. [72]
    Electrochemical Activation of Ni–Fe Oxides for the Oxygen Evolution ...
    Jun 18, 2025 · NiFe oxides have been historically identified as active for the OER, though they have been less studied in their more commercially relevant bulk ...Introduction · Materials and Methods · Results and Discussion · Conclusions
  73. [73]
    Tafel Slope Plot as a Tool to Analyze Electrocatalytic Reactions
    The Tafel slope is then obtained from the linear fit of the overpotential vs log J or from the overpotential vs log 1/Rct for LSV (CA/CP) and impedance ...Experimental Section · Supporting Information · Author Information · References
  74. [74]
    [PDF] Tafel Kinetics of Electrocatalytic Reactions: From Experiment to First
    Oct 23, 2014 · The empirical Tafel constants, a and b, can now be derived from eq 4, where the a = 2.303((RT)/ (αF))logi0 and b = 2.303((RT)/(αF)) Both ...
  75. [75]
    Cutting-edge innovations and sustainable catalysts for fuel cells and ...
    Sep 18, 2025 · This review offers an in-depth analysis of catalyst technologies for fuel cells and electrolyzers, emphasizing their pivotal role in driving ...
  76. [76]
    A Guide to Electrocatalyst Stability Using Lab-Scale Alkaline Water ...
    Jan 23, 2024 · We tested the electrolyzer performance between 20 and 80 °C, setting 80 °C as the upper limit to avoid excessive damage and minimize corrosion.
  77. [77]
    Principles of Photocatalysts and Their Different Applications: A Review
    Oct 31, 2023 · Review of material design and reactor engineering on TiO2 photocatalysis for CO2 reduction. J Photochem Photobiol, C. 2015;24:16–42. doi ...
  78. [78]
    Electrochemical Photolysis of Water at a Semiconductor Electrode
    Jul 7, 1972 · Electrochemical Photolysis of Water at a Semiconductor Electrode. AKIRA FUJISHIMA &; KENICHI HONDA. Nature volume 238, pages 37–38 ( ...Missing: original | Show results with:original
  79. [79]
    Visible Light Photoredox Catalysis with Transition Metal Complexes
    Many organic molecules may function as visible light photocatalysts; analogous to metal complexes such as Ru(bpy)32+, organic dyes such as eosin Y, 9,10- ...
  80. [80]
    Kinetics and mechanisms of charge transfer processes in ...
    Charge transfer in photocatalysis involves charge generation, trapping, recombination, and electron/hole transfer. These processes are vital for efficient  ...
  81. [81]
    Photocatalytic degradation of organic pollutants using TiO2-based ...
    Sep 20, 2020 · Results suggested that TiO2 was one of the most effective catalysts for the removal of organic chemicals particularly dyes and phenolic ...
  82. [82]
    Photocatalytic CO2 Reduction Using TiO2-Based Photocatalysts ...
    Inspired by natural photosynthesis, the artificial Z-scheme of photocatalyst is constructed to provide an easy method to enhance efficiency of CO2 reduction.
  83. [83]
    Z‐Scheme Photocatalytic Systems for Carbon Dioxide Reduction
    Although Z-schemes are applied in relevant photocatalytic reactions, such as environmental applications [34, 40], energetic applications [40], transformation of ...
  84. [84]
    Advances in asymmetric organocatalysis over the last 10 years
    Jul 29, 2020 · Here, we discuss how new activation modes are conquering challenging stereoselective transformations and the recent integration of organocatalysis with ...
  85. [85]
    Quantification of Electrophilic Activation by Hydrogen-Bonding ...
    A review. The hydrogen bond is the most important of all directional intermol. interactions. It is operative in detg. mol. conformation, mol. aggregation, and ...
  86. [86]
    Hydrogen-Bonding Activation in Chiral Organocatalysts - IntechOpen
    Sep 28, 2016 · This review briefly summarizes the hydrogen-bonding activation by chiral noncovalent organocatalysts. First, the differences between hydrogen- ...
  87. [87]
    Proline-Catalyzed Direct Asymmetric Aldol Reactions
    Here we report our finding that the amino acid proline is an effective asymmetric catalyst for the direct aldol reaction between unmodified acetone and a ...Author Information · ReferencesMissing: seminal | Show results with:seminal
  88. [88]
    Green Chemistry Meets Asymmetric Organocatalysis: A Critical ...
    May 13, 2021 · This review analyzes the synthesis of organocatalysts, focusing on green chemistry principles and the E factor, and classifies catalysts based ...Abstract · Lewis Base Catalysts · Brønsted Base Catalysts · Brønsted Acid Catalysts
  89. [89]
    Organocatalysis: A Brief Overview on Its Evolution and Applications
    Dec 3, 2018 · These substances, the so-called organocatalysts, present a lot of advantages, like being less toxic, less polluting, and more economically ...
  90. [90]
    [PDF] Scientific Background on the Nobel Prize in Chemistry 2021
    Oct 6, 2021 · Post-2000 developments. Since the papers by List and MacMillan in 2000, impressive developments have followed in the area of organocatalysis ...
  91. [91]
    [PDF] The advent and development of organocatalysis - Macmillan Group
    Sep 18, 2008 · Organocatalysis uses small organic molecules to catalyze organic transformations, and it became a thriving area in the past decade, with a ...
  92. [92]
    A simply smashing ammonia synthesis | C&EN Global Enterprise
    Dec 6, 2021 · Haber-Bosch combines N2 and H2 over an iron catalyst at high pressures and temperatures; it makes about 150 million tons of ammonia per year ...Missing: annual | Show results with:annual
  93. [93]
    Haber-Bosch Process - Stanford University
    Dec 14, 2022 · The reaction runs at temperatures in the range of 400°C-500 °C, pressure in the range of 150-300 bar, and with the presence of an iron-based ...<|separator|>
  94. [94]
    Production Of Sulfuric Acid By "contact Process" + Film
    Sulfur is first burned to produce SO2 in this process. Then, applying oxygen and vanadium pentoxide catalyst, it is converted into sulfur trioxide through an ...
  95. [95]
  96. [96]
    Manufacture of Asymmetric Hydrogenation Catalysts
    Oct 4, 2007 · Single-enantiomer drugs represent an increasingly large share of new chemical entities, leading to approaches in asymmetric synthesis.
  97. [97]
    Catalytic Fixed‐Bed Reactors - Eigenberger - Wiley Online Library
    Jul 15, 2012 · Catalytic fixed-bed reactors are the most important type of reactors for the synthesis of large-scale basic chemicals and intermediates.
  98. [98]
    CSTR - Visual Encyclopedia of Chemical Engineering Equipment
    Apr 1, 2022 · Continuous stirred-tank reactors are most commonly used in industrial processing, primarily in homogeneous liquid-phase flow reactions, where ...
  99. [99]
    All About the Cost Savings of Chemical Process Catalysts
    Sep 2, 2025 · Take a closer look at the ways chemical process catalysts lower costs by reducing energy use and waste, while boosting yields and production ...Boosts In Product Yields · Raw Material Savings · Catalysis And Scaling...
  100. [100]
    Enzyme Catalysis for Sustainable Value Creation Using Renewable ...
    Dec 6, 2024 · Enzyme catalysis plays an increasingly prominent role in industrial innovation and responsible production in various areas, such as green and sustainable ...
  101. [101]
    [PDF] Chapter 2 Selective Catalytic Reduction
    Selective catalytic reduction (SCR) has been applied to stationary source fossil fuel-fired combustion units for emission control since the early 1970s and ...
  102. [102]
    [PDF] Reaction Pathways and Kinetics for Selective Catalytic Reduction of ...
    Selective Catalytic Reduction (SCR) of NOx with NH3 by supported vanadium oxide catalysts is an important technology for reducing acidic NOx emissions from ...
  103. [103]
    Photocatalytic Dye Degradation from Textile Wastewater: A Review
    Photocatalysis stands out as a promising method for degrading diverse toxic and organic pollutants present in wastewater.
  104. [104]
    Principles and mechanisms of photocatalytic dye degradation on ...
    This review starts with (i) a brief overview on dye pollution, dye classification and dye decolourization/degradation strategies; (ii) focuses on the mechanisms ...
  105. [105]
    N–TiO2 Photocatalysts: A Review of Their Characteristics ... - MDPI
    Titanium dioxide is the most used photocatalyst in wastewater treatment; its semiconductor capacity allows the indirect production of reactive oxidative ...
  106. [106]
    Highly effective hydrogenation of CO2 to methanol over Cu/ZnO ...
    Jan 15, 2023 · In this work, improved operating conditions were reported for the sustainable catalytic hydrogenation of CO 2 to methanol using Cu/ZnO/Al 2 O 3 catalyst.
  107. [107]
    Improving the Cu/ZnO-Based Catalysts for Carbon Dioxide ...
    This paper reviews the history of the methanol production in industry, the impact of CO 2 emission on the environment, and analyzes the possibility of the Cu/ ...
  108. [108]
    The Atom Economy—A Search for Synthetic Efficiency - Science
    Efficient synthetic methods required to assemble complex molecular arrays include reactions that are both selective (chemo-, regio-, diastereo-, and enantio-) ...<|separator|>
  109. [109]
    Jojoba oil olefin metathesis: a valuable source for bio-renewable ...
    A high atom economy is expected for this catalytic process given that all products obtained may be used either as sources for bio-fuel (hydrocarbons), or as ...
  110. [110]
    Waste Catalyst Utilization: Extraction of Valuable Metals from Spent ...
    The metals present in the spent catalysts can be leached by water and pollute the environment. As these catalysts are hazardous wastes, landfills or other ...
  111. [111]
    Toward developing a reconcile solution for leaching of catalyst in ...
    Catalyst leaching is a problem commonly encountered in heterogeneous catalysis resulting in contamination of biodiesel product, reduction of catalyst's long- ...
  112. [112]
    Cytochrome P450 Enzymes and Drug Metabolism in Humans
    Nov 26, 2021 · Human cytochrome P450 (CYP) enzymes, as membrane-bound hemoproteins, play important roles in the detoxification of drugs, cellular metabolism, and homeostasis.
  113. [113]
    The Central Role of Cytochrome P450 in Xenobiotic Metabolism—A ...
    This short-review is intended to provide a summary on the major roles of CYPs in Phase I xenobiotic metabolism.
  114. [114]
    Biochemistry, Glycolysis - StatPearls - NCBI Bookshelf - NIH
    These reactions are all catalyzed by their own enzyme, with phosphofructokinase being the most essential for regulation as it controls the speed of glycolysis.
  115. [115]
    Three-dimensional Structure of Saccharomyces Invertase
    Invertase (EC 3.2.1.26; β-fructofuranosidase) catalyzes the hydrolysis of the disaccharide sucrose (table sugar) into glucose and fructose and is a major enzyme ...
  116. [116]
    Production and characterization of intracellular invertase ... - Nature
    Sep 28, 2023 · Invertase (β-fructofuranosidase, E.C. 3.2.1.26) is a carbohydrase that catalyzes the hydrolysis of sucrose to yield glucose and fructose in ...
  117. [117]
    Effects of animal rennet, fermentation-produced chymosin, and ...
    Coagulants play a crucial role in cheese production by catalyzing milk curdling, with traditional animal rennet long serving as the primary choice.
  118. [118]
    Obtaining of Recombinant Camel Chymosin and Testing Its Milk ...
    This recombinant enzyme has high specific activity and versatility and has been used to produce cheese from cow's, goat's, ewes', camel's and mare's milk.
  119. [119]
    Structural insights into the catalytic mechanism of lovastatin hydrolase
    This enzyme mainly catalyzes the transfer of the 2-methylbutyl side chain to monacolin J for the synthesis of lovastatin. In addition to acyltransferase ...
  120. [120]
    Biocatalyzed Synthesis of Statins: A Sustainable Strategy for ... - MDPI
    In this review, we will comment on the pleiotropic effects of statins and will illustrate some biotransformations nowadays implemented for statin synthesis.
  121. [121]
    Physiology, Digestion - StatPearls - NCBI Bookshelf
    Sep 12, 2022 · Pancreatic amylase, like salivary amylase, digests starch into maltose and maltotriose. Pancreatic lipase, secreted by the pancreas with an ...
  122. [122]
    Biofuel production: exploring renewable energy solutions for a ...
    Oct 15, 2024 · Genetically engineered microbes are utilized in fourth-generation biofuels to increase hydrocarbon output while lowering carbon emissions [33].
  123. [123]
    Microbial pathways for advanced biofuel production - Portland Press
    This review surveys the range of natural and engineered microbial systems for advanced biofuels production and summarises some of the techno-economic challenges ...Alcohol Biofuels · Third Generation Microbial... · Engineering Microbial...
  124. [124]
    Yeast: Making Food Great for 5000 Years. But What Exactly Is it?
    Sep 6, 2013 · Yeast was second. In the early days of ancient Egypt, around 3100 B.C., there lived a ruler named Scorpion, who probably did not look like The ...
  125. [125]
    Evolution of Food Fermentation Processes and the Use of Multi ...
    Nov 18, 2021 · The earliest form of fermentation was discovered by analysing the stone mortars from the Natufian burial sites of a semi-sedentary foraging ...
  126. [126]
    Experiments to investigate photosynthesis - OCR Gateway - BBC
    He put a mint plant in a closed container with a burning candle. The candle flame used up the oxygen and went out. After 27 days, Priestley was able to re-light ...
  127. [127]
    Vista de The History of Catalysis. From the Beginning to Nobel Prizes
    In 1811 Sigismund Konstantin Kirchhoff (1764-1833) dis-covered that heating an aqueous solution of starch with min-eral acids changed it into a gum, dextrin ...Missing: Gottfried hydrolysis
  128. [128]
    Justus Liebig and animal chemistry - PubMed
    Liebig's views on catalysis and fermentation brought him into a controversy with Louis Pasteur. Liebig's Animal Chemistry stimulated an interest in clinical ...
  129. [129]
    [PDF] Two centuries of catalysis - Indian Academy of Sciences
    Great scientists who are remembered for other reasons proved to be on the wrong side of the vitalism debate for at least part of the time, or they were on the ...
  130. [130]
    Wilhelm Ostwald – Facts - NobelPrize.org
    In 1894 he revealed what happens: a substance—a catalyst—can affect a chemical reaction's speed, but is not included in its end-products. This understanding ...
  131. [131]
    The Nobel Prize in Chemistry 1909 - NobelPrize.org
    The Nobel Prize in Chemistry 1909 was awarded to Wilhelm Ostwald in recognition of his work on catalysis and for his investigations into the fundamental ...
  132. [132]
    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 ...
  133. [133]
    Karl Ziegler and Giulio Natta | Science History Institute
    In 1963 Ziegler and Natta were awarded the Nobel Prize in Chemistry for their discoveries in the science and technology of polymers.
  134. [134]
    The Nobel Prize in Chemistry 2001 - NobelPrize.org
    The Nobel Prize in Chemistry 2001 was divided, one half jointly to William S. Knowles and Ryoji Noyori for their work on chirally catalysed hydrogenation ...
  135. [135]
    [PDF] Density Functional Theory in Surface Chemistry and Catalysis
    This has been essential in benchmarking computational surface science based on density functional theory (DFT) calculations and in providing experimental ...
  136. [136]
    Density Functional Theory - an overview | ScienceDirect Topics
    Since about 1990, the application of density functional theory (DFT) to chemical problems in general, and transition metal systems in particular, have exploded.
  137. [137]
    Advanced research and exploration of CRISPR technology in the ...
    This review provides a comprehensive overview of CRISPR tools and their applications in directed evolution, highlighting the principles, technological ...
  138. [138]
    Theoretical insights into single-atom catalysts - RSC Publishing
    Sep 1, 2020 · Single-atom catalysts (SACs) with atomically dispersed metals have emerged as a new class of heterogeneous catalysts and have attracted ...
  139. [139]
    Catalyst deactivation - ScienceDirect.com
    ... and irreversible, whereas inhibitors generally weakly and reversibly adsorb on the catalyst surface. Poisons can be classified as “selective” or “non-selective” ...
  140. [140]
    Sulfur poisoning of Pt/Al 2 O 3 catalysts, I. Determination of sulfur ...
    In the case of Pt/Al2O3 catalysts, a spectroscopic method was proposed to determine the amount of sulfur irreversibly bonded to platinum.
  141. [141]
    Decoding technical multi-promoted ammonia synthesis catalysts
    Aug 21, 2025 · CaO also increases the surface area as well as the resistance against gas impurities, and K2O is considered as an electronic promoter that ...
  142. [142]
    [PDF] "Promoters and Poisons" in: Handbook of Heterogeneous Catalysis ...
    Historically, discussions of the influence of one metal on another have centered on electronic (ligand) effects or geometric (ensemble, site-blocking) effects.
  143. [143]
    Indroductory lecture. Promotion in heterogeneous catalysis
    The aim of this introductory paper is to survery current views on the mode of action of promoters in heterogeneous catalysis. Examples of current beliefs on ...
  144. [144]
    An NMR Investigation of CO Tolerance in a Pt/Ru Fuel Cell Catalyst
    CO chemisorption significantly degrades fuel cell performance, and since CO is always present, an effective means for increasing the CO tolerance of these (and ...
  145. [145]
    Advanced Strategies for Mitigating Catalyst Poisoning in Low and ...
    Among them, CO and H2S pose the greatest challenges. Strong CO adsorption onto transition metals severely compromises catalytic processes in many applications.
  146. [146]
    Reactivity landscape of pyruvate under simulated hydrothermal vent ...
    Jul 19, 2013 · We report here that pyruvate reacts readily in the presence of transition metal sulfide minerals under simulated hydrothermal vent fluids at more moderate ...
  147. [147]
    A prebiotically plausible scenario of an RNA–peptide world - Nature
    May 11, 2022 · It predicts that life evolved from increasingly complex self-replicating RNA molecules. The question of how this RNA world then advanced to the ...
  148. [148]
    An RNA polymerase ribozyme that synthesizes its own ancestor
    Jan 27, 2020 · To achieve self-sustained replication, an RNA polymerase ribozyme must be able to synthesize copies of both itself and its complement in ...An Rna Polymerase Ribozyme... · Results · The Polymerase Can...
  149. [149]
    In vitro selection of ribozyme ligases that use prebiotically plausible ...
    Mar 2, 2020 · The RNA world hypothesis proposes a central role for RNA as both a catalytic and an informational polymer during the emergence of life.
  150. [150]
    Montmorillonite catalysis of RNA oligomer formation in aqueous ...
    Clay minerals mediate folding and regioselective interactions of RNA: A large-scale atomistic simulation study.
  151. [151]
    Montmorillonite-catalysed formation of RNA oligomers - NIH
    This montmorillonite serves as an efficient catalyst for the formation of RNA oligomers, if the exchangeable cation is an alkali or an alkaline earth metal ion.
  152. [152]
    Metals likely promoted protometabolism in early ocean alkaline ...
    Jun 19, 2019 · One of the most plausible scenarios of the origin of life assumes the preceding prebiotic autotrophic metabolism in sulfide-rich ...
  153. [153]
    Synthesis of prebiotic organics from CO 2 by catalysis with meteoritic ...
    May 25, 2023 · We experimentally demonstrate that iron-rich meteoritic and volcanic particles activate and catalyse the fixation of CO 2 , yielding the key precursors of life ...
  154. [154]
    Formation of nucleobases in a Miller–Urey reducing atmosphere
    Apr 10, 2017 · The study shows that Miller–Urey experiments produce RNA nucleobases in discharges and laser-driven plasma impact simulations carried out in a simple prototype ...
  155. [155]
    Dynamics and stability in prebiotic information integration: an RNA ...
    Jan 9, 2020 · The unavoidably high error rate of non-enzymatic RNA replication is the central problem of the RNA world scenario of the origin of life, which ...
  156. [156]
    Modified nucleotides may have enhanced early RNA catalysis - PNAS
    Mar 30, 2020 · We propose that noncanonical ribonucleotides, which would have been inevitable under prebiotic conditions, might decrease the RNA length required to have ...