Fact-checked by Grok 2 weeks ago

Heterogeneous catalysis

Heterogeneous catalysis is a chemical process in which the catalyst operates in a distinct from the reactants and products, most commonly involving a solid catalyst that facilitates reactions occurring in the gaseous or liquid . This surface-mediated phenomenon relies on the adsorption of reactant molecules onto the catalyst's active sites, where bond breaking and formation occur, followed by desorption of products, enabling the catalyst's regeneration and reuse. Catalysts in this context typically consist of an active metal or component dispersed on a high-surface-area support material, often enhanced by promoters to improve selectivity and stability. Heterogeneous catalysis dominates industrial chemical production, accounting for over 85% of all manufactured chemical products worldwide through its role in enabling efficient, large-scale transformations. It offers advantages such as straightforward separation of the catalyst from reaction mixtures, facilitating continuous processing and reducing operational costs compared to homogeneous alternatives. Key challenges include catalyst deactivation due to poisoning, , or , which necessitates ongoing research into robust designs, including nanostructured and single-atom catalysts. Prominent industrial applications underscore its economic impact: the Haber-Bosch process synthesizes from and using iron-based catalysts at high pressure and temperature, supporting production and . The employs vanadium pentoxide catalysts to oxidize to , enabling efficient manufacturing essential for fertilizers, dyes, and detergents. In petroleum refining, uses zeolite-based catalysts to convert heavy hydrocarbons into and olefins, contributing to roughly half of output. Emerging uses extend to , such as automotive exhaust converters that abate pollutants via platinum-group metals.

Fundamentals

Definition and Scope

Heterogeneous catalysis refers to a process in which the catalyst exists in a different from the reactants and products, most commonly involving a solid catalyst interacting with gaseous or liquid reactants. This phase separation facilitates the reaction at the interface between the catalyst surface and the reactants, enabling efficient chemical transformations without the catalyst dissolving into the reaction mixture. The key principles of heterogeneous catalysis revolve around surface-mediated reactions, where reactants adsorb onto specific active sites on the catalyst surface, lowering the and thereby enhancing rates while leaving the unchanged. These active sites—often defects, edges, or specific atomic arrangements on the catalyst—serve as localized regions where bonds in the reactants are weakened and reformed into products. Adsorption typically initiates the process, allowing reactants to interact closely with the catalyst surface before proceeding to and desorption steps. The scope of heterogeneous catalysis encompasses gas-solid systems, such as ammonia synthesis, liquid-solid interactions like in solution, and more complex multiphase setups in industrial reactors. It is distinctly differentiated from , where the catalyst shares the same phase (e.g., all in solution) with the reactants, and from biocatalysis, which relies on enzymes or biological systems typically operating in aqueous environments. A basic schematic of the catalyst-reactant interface illustrates reactants approaching and binding to the :
Reactants (gas/liquid) → Adsorption on active sites → Surface reaction → Desorption → Products
          Solid Catalyst Surface
This framework highlights the spatial confinement of the reaction to the catalyst's exterior or porous interior. Heterogeneous catalysis underpins approximately 90% of chemical processes by volume, playing a crucial role in energy production (e.g., fuel reforming), chemical (e.g., ), and environmental control (e.g., catalytic converters for emission reduction). Its widespread adoption stems from the ease of catalyst recovery and reuse, contributing to and in large-scale operations.

Historical Development

The origins of heterogeneous catalysis trace back to the early , when initial observations highlighted the ability of metal surfaces to accelerate chemical reactions without being consumed. In 1817, reported that a hot platinum wire inserted into a mixture of and air induced at the wire's surface, producing heat without a visible , marking one of the first documented catalytic effects in gas-phase reactions. This discovery laid foundational insights into surface-mediated processes. During the 1850s, Henri Sainte-Claire Deville expanded on such phenomena through systematic studies of 's catalytic role in oxidation and decomposition reactions, including the generation of from metal cyanides using platinum sponge, which demonstrated the material's efficacy in promoting reactions at elevated temperatures. A pivotal advancement occurred in the late 19th and early 20th centuries with the development of practical catalytic processes for industrial applications. In 1897, Paul Sabatier and Jean-Baptiste Senderens pioneered the direct hydrogenation of unsaturated organic compounds using finely divided nickel as a catalyst, enabling efficient conversion of carbon monoxide and other gases into valuable products like methane; this work earned Sabatier the 1912 Nobel Prize in Chemistry for his contributions to catalytic hydrogenation theory and practice. Concurrently, Fritz Haber's research from 1909, later industrialized by Carl Bosch between 1910 and 1913, established the Haber-Bosch process for ammonia synthesis, employing iron-based catalysts under high pressure and temperature to fix atmospheric nitrogen—a breakthrough that revolutionized fertilizer production and earned Haber the 1918 Nobel Prize in Chemistry. Irving Langmuir's contributions in the early 20th century further advanced the field through his 1916-1918 development of the Langmuir adsorption isotherm, which mathematically described monolayer adsorption on surfaces, providing a theoretical framework for understanding heterogeneous catalytic mechanisms and earning him the 1932 Nobel Prize in Chemistry. By the mid-20th century, heterogeneous catalysis expanded into specialized materials and environmental applications. In the 1930s and 1940s, pioneering work led to the synthesis and characterization of zeolites—crystalline aluminosilicates with uniform pores—as documented in early studies on and molecular sieving. These materials emerged as key catalysts in the , leveraging their shape-selective properties for refining processes like cracking. The marked a surge in environmental catalysis following the U.S. Clean Air Act of 1970, which mandated sharp reductions in vehicle emissions; this spurred the widespread adoption of three-way catalytic converters using platinum-rhodium and formulations to oxidize and hydrocarbons while reducing nitrogen oxides, debuting commercially in 1975. In the modern era, from the 2000s onward, computational methods and nanoscale innovations have transformed catalyst design. Density functional theory and high-throughput screening enabled the rational prediction of active sites and reaction pathways, accelerating the development of tailored heterogeneous catalysts for energy and sustainability applications. The 2010s introduced single-atom catalysts, where isolated metal atoms anchored on supports maximize atom efficiency and selectivity; the concept gained prominence with Tao Zhang's 2011 report on platinum single atoms anchored on iron oxide (FeOx) for CO oxidation, heralding a shift toward precise, low-loading systems.

Reaction Mechanisms

Adsorption Processes

In heterogeneous catalysis, adsorption serves as the essential initial step, whereby reactant molecules from the gas or liquid phase bind to the surface of a solid , thereby concentrating them at active sites and enabling the formation of reactive intermediates that facilitate subsequent chemical transformations. This process lowers the for reactions by modifying or breaking molecular bonds upon surface interaction, as exemplified in key industrial processes like ammonia synthesis where adsorption on iron surfaces is rate-limiting. Adsorption in catalysis primarily occurs through two distinct mechanisms: physisorption and chemisorption. Physisorption involves weak, non-specific interactions via van der Waals forces, such as London dispersion or dipole-dipole attractions, resulting in low activation energies (often near zero), reversibility, and typical heats of adsorption below 50 kJ/mol; it allows multilayer formation and is prevalent at low temperatures. In contrast, chemisorption entails strong, specific chemical bonds (e.g., covalent or ionic) between the adsorbate and surface atoms, with higher activation energies, heats of adsorption exceeding 80 kJ/mol, and often irreversibility under typical reaction conditions; it is limited to a monolayer and is crucial for activating reactants in catalytic cycles. The extent and efficiency of adsorption are influenced by several key factors, including the catalyst's surface area (which determines the number of available sites), pore structure (affecting and accessibility), temperature (where exothermic adsorption decreases with rising temperature per ), and adsorbate pressure (driving higher coverage at elevated partial pressures). For instance, high surface area materials like zeolites or supported nanoparticles enhance capacity, while optimized pore sizes in mesoporous catalysts improve by facilitating transport to active sites. A foundational model for describing adsorption equilibrium on uniform surfaces with a finite number of identical sites is the Langmuir isotherm, which assumes no adsorbate interactions and monolayer coverage. The fractional surface coverage \theta is given by \theta = \frac{KP}{1 + KP}, where K is the adsorption equilibrium constant and P is the partial pressure of the adsorbate. This equation derives from a site balance at equilibrium: the rate of adsorption, proportional to P(1 - \theta), equals the rate of desorption, proportional to \theta, yielding k_a P (1 - \theta) = k_d \theta, where k_a and k_d are the rate constants; rearranging gives \theta = \frac{(k_a / k_d) P}{1 + (k_a / k_d) P}, with K = k_a / k_d = e^{-\Delta G^\circ / RT}. Adsorption can further be classified as associative (molecule remains intact) or (molecule breaks into fragments upon binding). For example, adsorption on surfaces like or is typically , where H_2 splits into atomic (H_\text{ad}) with a heat of adsorption around 90 kJ/mol, forming a precursor to reactions.

Surface Reaction Steps

In heterogeneous catalysis, the surface reaction steps involve the chemical transformations of adsorbed on the catalyst surface, where bonds are broken and formed at active sites following initial adsorption. These steps occur after reactants have adsorbed onto the surface, enabling interactions that would be energetically unfavorable in the gas phase. The process typically proceeds through elementary s, including of adsorbates to bring species together, recombination of adsorbed atoms or molecules (such as O(ad) + (ad) → ₂(ad)), and reactions where an adsorbed species extracts an atom from another adsorbate. These elementary steps are crucial for determining the overall and selectivity, with often facilitating the migration of intermediates across the catalyst . Two primary mechanisms describe these surface reactions: the Langmuir-Hinshelwood (LH) mechanism, where both reactants adsorb onto the surface before reacting, and the Eley-Rideal (ER) mechanism, where one reactant remains in the gas phase and reacts directly with an adsorbed . In the LH mechanism, the rate-determining step is often the bimolecular reaction between co-adsorbed , leading to the rate law r = k \theta_A \theta_B, where \theta_A and \theta_B are the surface coverages of reactants A and B, derived under the assumption of uniform active sites and steady-state conditions. The ER mechanism, in contrast, involves a gas-phase colliding with an adsorbed , resulting in a rate expression like r = k \theta_A P_B, where P_B is the of the gas-phase reactant; this is common in reactions with weak adsorption of one . Adsorption serves as the prerequisite for these mechanisms by positioning reactants near active sites. Active sites, often consisting of atomic ensembles such as or corner atoms on metal nanoparticles, play a pivotal role in these reactions by providing coordinatively unsaturated positions that lower barriers and influence selectivity. For instance, sites in nanoparticles facilitate selective bond due to their unique electronic structure, promoting desired pathways while suppressing side reactions. The ensemble size and composition can dictate product distribution; smaller ensembles may favor , whereas larger ones enable complete reactions. A representative example is the oxidation of on Pt surfaces, where the LH dominates under typical conditions. The pathway begins with dissociative adsorption of O₂ to form O(ad) atoms, followed by adsorption of at adjacent sites; surface diffusion brings these species together for recombination to CO₂(ad), which then desorbs. The rate is often limited by O₂ dissociation or CO coverage, with edge sites on Pt nanoparticles enhancing activity by stabilizing O(ad) and reducing the recombination barrier to approximately 100-120 kJ/mol. Studies on Pt(111) single crystals confirm this pathway, showing oscillatory behavior due to competing CO poisoning and oxide formation.

Desorption Processes

Desorption represents the final step in the of heterogeneous catalysis, where reaction products detach from surface, thereby regenerating active sites for subsequent reactant adsorption and reaction. This process is essential for maintaining the turnover of , as the release of products prevents site blocking and ensures continuous operation under steady-state conditions. In many catalytic systems, such as ammonia synthesis on iron catalysts, desorption directly influence the overall by determining the availability of surface sites. Desorption processes in heterogeneous catalysis are broadly classified into thermal desorption, which can be either activated or non-activated, and recombinative (or associative) desorption. Thermal desorption involves the thermally induced release of adsorbed , where non-activated desorption occurs without a significant energy barrier if the adsorption was exothermic and barrierless, allowing spontaneous departure at elevated temperatures; an example is the desorption of from surfaces. Activated thermal desorption requires overcoming an activation barrier, often observed in systems like on , where the process demands higher temperatures (900–1200 K) due to the needed to break strong chemisorption bonds. Recombinative desorption, common for dissociated adsorbates, entails the recombination of atomic on the surface prior to release as a , such as 2H(ads) → H₂(g) on (111), exhibiting second-order due to the bimolecular nature of the step. The energetics of desorption are governed by the E_d, which is typically related to the heat of adsorption by , making desorption endothermic and often rate-limiting at low temperatures. The are described by the Polanyi-Wigner equation, which models the desorption rate as coverage-dependent: -\frac{d\theta}{dt} = k \theta^n where \theta is the surface coverage, k = \nu \exp(-E_d / RT) is the rate constant with \nu (typically $10^{13} s⁻¹ for processes), R is the , T is , and n is the desorption order (e.g., n=1 for non-dissociative, n=2 for recombinative). This equation, derived from transition-state theory, highlights how E_d values, often 50–200 kJ/mol in catalytic systems like CO on Fe₃O₄, dictate the temperature threshold for desorption. Desorption rates exhibit strong dependence on surface coverage \theta, with higher coverage often lowering E_d due to adsorbate-adsorbate interactions, leading to shifts in desorption peaks toward lower temperatures in temperature-programmed experiments. Temperature plays a critical role, as increasing T exponentially accelerates the rate per the Arrhenius form in the Polanyi-Wigner equation, enabling control over desorption in catalytic processes; for instance, in hydrogenation reactions, optimal temperatures balance desorption with preceding surface reactions to maximize yield. A primary technique for studying desorption is temperature-programmed desorption (TPD), which involves adsorbing species on the catalyst at low temperature, then linearly ramping the temperature (e.g., 10 K/s) while monitoring desorbed gases via . TPD spectra reveal desorption peaks whose position, shape, and area provide insights into E_d, order n, and coverage; for example, peak temperature T_p approximates E_d \approx 0.25 T_p (in kJ/mol) via Redhead analysis for first-order processes assuming \nu = 10^{13} s⁻¹. This method is widely applied to characterize catalytic surfaces, such as determining binding energies in metal systems.90024-9)

Catalyst Materials and Design

Types of Catalyst Materials

Heterogeneous catalysts are broadly classified into several material types, each offering distinct properties that influence their reactivity, selectivity, and stability in various chemical processes. These include metals, metal oxides, zeolites, and supported systems, with advanced variants such as nanostructured and bimetallic compositions enhancing performance through tailored structures. Metallic catalysts, particularly transition metals, are widely used due to their ability to facilitate adsorption and bond breaking in reactions like and oxidation. Noble metals such as (Pt) and (Pd) excel in processes, where Pt catalyzes the selective reduction of alkenes to alkanes under mild conditions, owing to its high affinity for dissociation. In contrast, base metals like and provide cost-effective alternatives for large-scale applications, such as in for , though they often require higher temperatures to achieve comparable activity. The distinction between noble and base metals arises from differences in electronic structure and resistance to , with noble metals generally exhibiting greater stability in oxidative environments. Metal oxides serve as both active catalysts and supports, leveraging their properties and surface acidity for diverse reactions. For instance, (TiO₂) and aluminum oxide (Al₂O₃) are employed in oxidation and dehydration processes, with TiO₂ promoting photocatalytic reactions through its bandgap that enables electron-hole pair generation under light. Al₂O₃, often used as a , also acts independently in acid-catalyzed due to its amphoteric nature, balancing and sites to stabilize intermediates. These oxides exhibit tunable behavior, where cations like in CeO₂ facilitate and release, enhancing catalytic cycles in automotive exhaust treatment. Zeolites, microporous aluminosilicates, are prized for shape-selective catalysis, confining reactants within their uniform pores to favor specific products. Their framework enables high selectivity in cracking and , as seen in zeolite for in refining, where pore sizes around 0.5-0.6 nm restrict larger molecules. Acidity in zeolites stems from Brønsted and sites: Brønsted sites, formed by bridging OH groups, protonate substrates for carbocation-mediated reactions, while sites from extra-framework Al³⁺ coordinate with donors to activate C-H bonds. This dual acidity/basicity allows zeolites to catalyze both acid-driven transformations and bifunctional processes involving metal loading. Supports play a crucial role in dispersing active phases to maximize surface area and prevent . High-surface-area materials like silica (SiO₂) provide inert, thermally stable platforms for metal deposition, with its groups aiding uniform distribution in processes like over iron/silica. Carbon-based supports, such as or , offer high electrical conductivity and resistance to acidic conditions, ideal for electrocatalysis; for example, carbon supports enhance Pt dispersion in anodes, improving mass transport. These supports not only increase accessibility but also modulate electronic properties through metal-support interactions. Nanostructured catalysts, including nanoparticles and single-atom variants, address limitations of bulk materials by exposing more active sites. Metal nanoparticles, typically 2-10 nm, exhibit size-dependent activity; for , smaller Pd nanoparticles show higher turnover frequencies due to increased edge sites. Single-atom catalysts (SACs), like Pt₁ on CeO₂, achieve atomic efficiency with enhanced stability and activity; in CO oxidation, isolated Pt atoms leverage CeO₂'s oxygen mobility for low-temperature performance, outperforming nanoparticles by activating oxygen. SACs minimize metal use while maximizing utilization, though challenges in persist for scalability. Bimetallic and alloy catalysts exploit synergistic effects to improve performance beyond monometallic counterparts. In Pt-Ru alloys for fuel cells, Ru enhances CO tolerance by facilitating its oxidation at lower potentials via bifunctional mechanism, where Ru-OH adsorbs and reacts with Pt-bound CO, boosting anode efficiency in proton exchange membrane fuel cells. Such synergies arise from electronic modifications—e.g., Ru donates electrons to Pt, weakening CO binding—and geometric effects that create ensemble sites for reactants. These alloys often show prolonged stability under operating conditions compared to pure metals.

Preparation and Synthesis Methods

Heterogeneous catalysts are typically prepared by depositing active metal or metal oxide components onto high-surface-area supports, such as alumina or silica, to achieve optimal and activity. Common methods focus on controlling the structure and to maximize catalytic performance, with techniques ranging from traditional impregnation to advanced colloidal approaches. These routes aim to produce uniform distributions of active sites while ensuring scalability for industrial applications. Impregnation is one of the most widely used techniques for preparing supported metal catalysts, involving the adsorption of metal precursor onto porous supports. In impregnation, excess is used, allowing the precursor to penetrate the support pores before and ; however, this can lead to uneven distribution if not controlled. Dry impregnation, or incipient wetness, employs a volume of equal to the support's pore volume, promoting uniform loading by and minimizing waste. For example, nitrate is often impregnated onto alumina supports for Fischer-Tropsch synthesis, achieving metal loadings up to 20 wt%. The method's simplicity and scalability make it suitable for , though limitations include potential aggregation during , which affects . Precipitation methods, particularly , enable the simultaneous formation of mixed oxide supports and active phases from aqueous solutions of metal salts. By adding a precipitating agent like or under controlled and temperature, hydroxides or carbonates form and are filtered, washed, dried, and calcined to yield homogeneous compositions. This technique is ideal for high-loading catalysts, such as Cu/ZnO/Al₂O₃ used in synthesis, where precise (around 6-7) ensures small particle sizes (e.g., 3 nm CuO crystallites). Advantages include high dispersion and strong metal-support interactions, but challenges arise from waste generation during washing and sensitivity to local variations, impacting batch-to-batch reproducibility. Advanced techniques like sol-gel synthesis produce porous structures with tailored microstructures through and of metal alkoxides, forming a sol that gels into a network. This method allows incorporation of active metals during gelation, followed by drying to form xerogels or aerogels with high surface areas (up to 700 m²/g). For instance, silica-supported catalysts for are prepared by adjusting the water-to-precursor ratio, yielding mesoporous materials with uniform pores. Benefits include low-temperature processing and homogeneity, but the process is complex and prone to cracking during drying, limiting scalability. Colloidal synthesis offers precise control over , often via seed-mediated growth where small particles (e.g., 5 nm ) serve as sites for controlled overgrowth using molecular precursors and stabilizers. This results in anisotropic structures like nanorods (5 nm , 40-100 nm ) deposited onto supports such as alumina-silica for Fischer-Tropsch catalysts, enhancing stability and activity. The approach excels in achieving narrow size distributions below 10 nm, crucial for high surface-to-volume ratios and activity, though removal steps can introduce variability. Limitations include difficulties in scaling beyond lab quantities due to stabilizer handling. Following synthesis, activation steps are essential to transform precursors into active forms. Calcination involves heating the dried material (typically 300-600°C in air) to decompose precursors, remove volatiles, and stabilize the structure, as seen in converting impregnated nitrates to oxides on alumina supports. Reduction, often with hydrogen at 300-500°C, then reduces metal oxides to metallic particles, enhancing dispersion; for example, low-temperature calcination (250-300°C) followed by H₂ reduction yields highly active cobalt particles for syngas conversion. These steps must be optimized to prevent sintering, which enlarges particles beyond optimal sizes (<10 nm). Key control parameters include particle size distribution, targeted below 10 nm to maximize active site exposure and turnover frequency, achieved by adjusting precursor concentration, pH, or stabilizers in synthesis. Promoter addition, such as potassium in iron-based ammonia synthesis catalysts, enhances electron donation and selectivity by modifying surface basicity; for instance, K₂O doping (1-2 wt%) improves N₂ activation on Fe sites. These additives are incorporated during impregnation or coprecipitation to fine-tune activity without altering bulk structure. Challenges in preparation include ensuring reproducibility across batches, as minor variations in drying rates or pH can lead to inconsistent dispersions, and scaling from lab to industrial levels, where advanced methods like face equipment and cost barriers while maintaining uniformity. Addressing these requires standardized protocols and in-line monitoring to bridge the gap between research and production.

Characterization and Analysis

Surface Structure Techniques

Surface structure techniques are essential for elucidating the atomic and molecular arrangements on catalyst surfaces, which directly influence adsorption, reaction pathways, and overall catalytic performance in heterogeneous systems. These methods provide insights into surface composition, active site geometries, and dynamic changes under reaction conditions, enabling the rational design of more efficient catalysts. By probing the interface between the solid catalyst and reactants, such techniques bridge the gap between material synthesis and mechanistic understanding, revealing how surface defects, facets, and coordinative environments dictate selectivity and activity. Spectroscopic approaches, such as , are widely used to determine the oxidation states, elemental composition, and chemical environments of surface atoms in heterogeneous catalysts. XPS operates by irradiating the sample with X-rays and measuring the kinetic energy of emitted photoelectrons, which provides depth-sensitive information up to about 10 nm, making it ideal for analyzing supported metal particles and oxide layers. For instance, variants allow studies under near-reaction conditions, capturing changes in surface oxidation during catalysis. Complementing XPS, identifies adsorbed species and their bonding modes on catalyst surfaces by measuring vibrational spectra in a reflective setup suitable for powders. DRIFTS is particularly valuable for operando studies, where it detects transient intermediates like CO or NO on metal sites, revealing adsorption geometries that correlate with reaction kinetics. Microscopic techniques offer direct visualization of surface features at high resolution. Scanning tunneling microscopy (STM) achieves atomic-scale imaging of catalyst surfaces by scanning a sharp tip to measure tunneling currents, enabling the observation of single atoms, clusters, and reconstruction under ultrahigh vacuum or elevated pressures. In heterogeneous catalysis, STM has been pivotal in studying model systems like Pt nanoparticles on oxides, showing how surface facets evolve during reactions. Transmission electron microscopy (TEM), including high-resolution variants, characterizes particle size distributions, morphologies, and lattice structures by transmitting electrons through thin samples. Aberration-corrected TEM provides sub-angstrom resolution for identifying atomic arrangements in supported catalysts, such as the dispersion of active metals on carbon supports. Diffraction-based methods probe long-range order and local atomic coordination. X-ray diffraction (XRD) determines the crystallite structure, phase purity, and average particle sizes in polycrystalline catalysts through analysis of , often using the for size estimation from peak broadening. In situ XRD tracks structural transformations, like phase changes in metal oxides during redox cycles. (EXAFS) spectroscopy elucidates coordination environments around absorbing atoms by analyzing oscillations in X-ray absorption spectra beyond the edge, providing bond lengths and neighbor counts without requiring long-range order. EXAFS is crucial for single-atom catalysts, where it confirms low-coordination sites in frameworks like or . A notable application is the use of nuclear magnetic resonance (NMR) spectroscopy to identify active sites in zeolite catalysts, where solid-state techniques like magic-angle spinning NMR probe framework aluminum distributions and metal incorporations. For example, 27Al and 29Si NMR distinguish tetrahedral sites responsible for in proton-form zeolites, linking site density to cracking activity. These insights have guided the design of shape-selective catalysts for hydrocarbon processing. Despite their power, surface structure techniques face limitations, including the need for ultrahigh vacuum in many ex situ measurements, which may not reflect operando conditions and lead to surface reconstructions. In situ and operando variants, such as high-pressure or , mitigate this but often compromise resolution or require specialized setups. Such structural details ultimately inform performance metrics, like turnover frequencies tied to specific site types.

Performance Evaluation Methods

Performance evaluation in heterogeneous catalysis involves quantitative assessment of catalyst activity, selectivity, and stability to determine efficacy under reaction conditions. Key metrics focus on intrinsic rates and overall productivity, ensuring comparisons across studies are standardized and reliable. Experimental setups simulate industrial processes, while computational methods predict performance based on fundamental interactions. These approaches enable optimization of catalysts for practical applications. Activity is primarily quantified using turnover frequency (TOF), defined as the number of product molecules formed per active site per unit time, expressed as \text{TOF} = \frac{\text{moles of product}}{\text{moles of active sites} \times \text{time}}. This metric isolates the intrinsic reactivity of catalytic sites, independent of catalyst loading or reactor geometry, and is essential for comparing site-specific performance. Space-time yield complements TOF by measuring productivity per unit reactor volume per unit time, providing a practical indicator of process efficiency in scaled systems. Standardization of active site determination, often via techniques like chemisorption, is critical to avoid overestimation or underreporting of TOF values. Selectivity, the fraction of converted reactant forming the desired product relative to total products, and yield, the amount of desired product relative to initial reactant, are evaluated in fixed-bed reactors where steady-state flow allows precise monitoring of effluent composition via gas chromatography or mass spectrometry. Conversion, standardized as X = \frac{\text{reactant inlet} - \text{reactant outlet}}{\text{reactant inlet}}, quantifies reactant utilization and is routinely reported alongside selectivity to assess overall process viability. Stability is tested by tracking these metrics over extended periods, typically in continuous operation, to identify deactivation onset and long-term performance. Testing setups include batch reactors, which use finite reactant volumes for initial screening of catalyst behavior in closed systems, versus continuous flow reactors like fixed-bed configurations that mimic industrial steady-state conditions for reliable kinetic data. In-situ and operando spectroscopy, such as infrared or X-ray absorption under working conditions, enables real-time monitoring of active species and reaction intermediates to correlate performance with dynamic surface changes. Computationally, density functional theory (DFT) calculates binding energies of adsorbates, guided by the , which posits optimal catalysis at intermediate adsorption strengths to balance activation and desorption. This volcano-shaped relationship between binding energy and activity informs catalyst design without exhaustive experimentation.

Deactivation and Regeneration

Mechanisms of Catalyst Deactivation

Heterogeneous catalysts deactivate over time due to various mechanisms that reduce active site availability or alter surface properties, leading to diminished reaction rates. These processes are broadly classified into chemical, thermal, and mechanical categories, with deactivation manifesting as a gradual or abrupt loss of catalytic activity. Understanding these mechanisms is essential for designing more stable catalysts, as deactivation remains a primary limitation in industrial applications. Poisoning occurs through the strong, often irreversible chemisorption of impurities on active sites, blocking access for reactants. Common poisons include sulfur compounds, which adsorb dissociatively on metal surfaces like , forming stable sulfides that reduce the ensemble size required for reactions such as . For instance, exposure to H₂S can decrease nickel catalyst activity by several orders of magnitude at concentrations as low as 1–10 ppb, with the poisoning effect being site-specific and dependent on the poison's adsorption energy. Coking, or carbon deposition, involves the formation of carbonaceous species that coat active sites or plug pores, particularly in hydrocarbon processing. These deposits arise from side reactions like dehydrogenation or polymerization, yielding structures such as filamentous carbon or polymeric coke. In reactions involving hydrocarbons, coke buildup can encapsulate metal particles, reducing surface area and hindering mass transport; for example, steam reforming catalysts may form carbon filaments above 450°C, exacerbating deactivation in coke-sensitive processes. Sintering refers to the thermal agglomeration of catalyst particles, which decreases metal dispersion and surface area. This process is driven by mechanisms like Ostwald ripening, where smaller particles dissolve atomically and redeposit onto larger ones via surface or vapor-phase diffusion, favored at high temperatures (>500°C). can accelerate sintering by enhancing metal oxide mobility; studies on supported catalysts show up to 70% loss in surface area after 50 hours at 750°C. Particle migration and coalescence also contribute, especially for larger nanoparticles. Phase transformation involves chemical or structural changes in the catalyst under reaction conditions, converting active phases to less active or inactive ones. This can include oxidation of metallic sites to oxides or solid-state transitions, such as the γ- to α-alumina shift at approximately °C, which drastically reduces from over 200 m²/g to about 1 m²/g. Such transformations alter electronic properties and site geometry, often irreversibly, and are influenced by environments or high temperatures. Fouling results from the physical accumulation of extraneous materials on the catalyst surface or within pores, impeding reactant . This includes deposition of like dust or reaction byproducts that block access to active sites; for example, fly ash in combustion-related catalysis can rapidly plug pore networks, leading to increases and flow restrictions. Unlike , fouling is often mechanical and reversible through cleaning, but it severely impacts mass transfer-limited systems. involves mechanical breakdown of catalyst particles due to or handling, generating fines that reduce effective inventory and increase emissions. Deactivation kinetics are commonly modeled to predict catalyst lifetime, with the first-order model describing activity a (relative to initial activity a_0) as a = e^{-kt}, where k is the deactivation rate constant and t is time. More general approaches, like the power-law expression, account for parallel deactivation pathways: -\frac{da}{dt} = k_d a^m, where m reflects the order (e.g., m=1 for poisoning). These models integrate experimental data to quantify rates, often revealing that deactivation accelerates with temperature via Arrhenius dependence. Regeneration strategies can mitigate these effects by reversing certain mechanisms, such as coke removal via oxidation.

Strategies for Regeneration

Regeneration of heterogeneous catalysts aims to restore activity lost due to reversible deactivation mechanisms, such as coke deposition or mild poisoning, through targeted techniques that minimize structural damage to the catalyst. These strategies are broadly classified as in-situ (performed within the reactor) or ex-situ (offline treatment), with the choice depending on the deactivation type and process economics. Oxidative and chemical methods address surface fouling, while physical approaches target sintering, and preventive measures incorporate design elements to extend catalyst life. One common method for removing coke deposits involves oxidative regeneration, where carbonaceous residues are burned off using oxygen or air at temperatures typically between 400°C and 600°C, converting coke to and CO₂. This approach is effective for catalysts in hydrocarbon processing, as it restores surface area without excessive , though careful control of oxygen is required to avoid over-oxidation of active metals. For instance, in (FCC) units, continuous in-situ oxidative regeneration occurs in a regenerator operating at 500–560°C, allowing seamless catalyst circulation between reaction and regeneration zones. Chemical cleaning techniques are employed to eliminate poisons that adsorb strongly on active sites, such as species. A representative example is the use of oxide (ZnO) beds to irreversibly adsorb H₂S, preventing its migration to downstream catalysts in processes like hydrotreating or Fischer-Tropsch synthesis, where sulfur levels must be reduced below 50 ppb. This ex-situ or guard-bed approach effectively regenerates the primary catalyst by sacrificial removal of contaminants, with ZnO regeneration possible via high-temperature reduction if needed. Physical regeneration methods focus on reversing sintering, where metal particles agglomerate and lose dispersion. Oxychlorination is a widely adopted technique for platinum-based catalysts, involving treatment with a chlorine-oxygen mixture at 500–550°C to form volatile PtClₓ species that redisperse upon subsequent reduction, restoring up to 80–90% of original dispersion on supports like alumina. This in-situ or ex-situ process is standard in catalytic reforming, where sintered Pt is redispersed every 6–12 months to maintain activity. Pressure swing regeneration, applicable in some gas-phase processes, exploits pressure changes to desorb weakly bound species or volatiles, such as in sorption-enhanced reactions, enabling faster cycling without thermal input. Preventive strategies integrated during catalyst design or operation help mitigate deactivation from the outset. Adding promoters, such as tin () to /Al₂O₃ or molybdenum () for sulfur resistance, modifies surface ensembles to reduce coke formation or poison adsorption, extending operational life by 20–50%. Continuous operation modes, like fluidized beds in FCC, inherently incorporate regeneration to avoid downtime, with catalyst inventory continuously refreshed to balance activity. Economically, regeneration strategies involve trade-offs between treatment frequency and full catalyst replacement costs, which can be substantial in large-scale operations. For example, regenerating in-situ avoids the significant expense of while recovering 70–90% activity, though repeated cycles may necessitate periodic full refresh to prevent irreversible losses; overall, effective regeneration optimizes lifespan against and penalties.

Applications and Examples

Gas-Phase Industrial Processes

Heterogeneous catalysis plays a pivotal role in gas-phase industrial processes, enabling efficient conversion of reactants under controlled conditions to produce essential chemicals and mitigate emissions. One of the most significant applications is the synthesis of ammonia via the Haber-Bosch process, where nitrogen and hydrogen gases react over iron-based catalysts to form NH₃, supporting global fertilizer production. The Haber-Bosch process operates at high pressures of 200-400 atm and temperatures of 400-500°C, utilizing promoted (Fe₃O₄) as the primary to achieve industrially viable rates despite the reaction's thermodynamic limitations. This endothermic, (N₂ + 3H₂ ⇌ 2NH₃) requires careful optimization of conditions to maximize yield, with the iron often promoted by potassium, aluminum, and calcium oxides to enhance activity and stability. Ruthenium-based alternatives, such as supported on carbon or metal oxides, offer higher activity at milder conditions (e.g., 300-450°C and 4-15 ), though their higher cost limits widespread adoption to niche, low-pressure applications. These catalysts demonstrate superior dissociation but require promoters like cesium to mitigate inhibition by . In production, heterogeneous catalysis facilitates the selective partial of impurities to , preventing in downstream units. Palladium-based catalysts, typically supported on alumina or silica, are employed for this gas-phase reaction (C₂H₂ + H₂ → C₂H₄), operating at 100-200°C and near-atmospheric to achieve high selectivity (>90%) toward while minimizing over- to . The Pd active sites enable dissociative adsorption of and , with promoter metals like silver or tuning the electronic properties to suppress oligomerization and formation. This process is critical in operations, where levels must be reduced below 5 ppm for polymer-grade . Steam methane reforming (SMR) represents a cornerstone for , converting into (CO + H₂) via an endothermic gas-phase reaction over catalysts supported on alumina (Ni/Al₂O₃). The primary reaction, CH₄ + H₂O → CO + 3H₂, occurs at approximately 800°C and 20-30 atm in tubular reactors, with the nickel particles (typically 10-20 nm) catalyzing C-H bond breaking while the alumina support provides thermal stability and dispersion. Side reactions like the water-gas shift (CO + H₂O → CO₂ + H₂) further adjust the H₂/CO ratio, making SMR the dominant route for approximately 76% of global as of 2024, though carbon deposition remains a challenge managed by excess steam. Ethylene oxide (EO) production exemplifies selective in gas-phase heterogeneous catalysis, where reacts with oxygen over silver catalysts supported on low-surface-area α-alumina (Ag/α-Al₂O₃) to yield the (C₂H₄ + ½O₂ → C₂H₄O). This exothermic process runs at 220-280°C and 10-20 atm, achieving selectivities of 80-90% under optimized conditions, with and promoters enhancing oxygen adsorption and suppressing total to CO₂. The α-Al₂O₃ support (with surface area <5 m²/g) minimizes acid-catalyzed side reactions like isomerization, while silver particle sizes (1-10 μm) influence the electrophilic oxygen species responsible for epoxidation. EO is a key intermediate for ethylene glycol and surfactants, underscoring the process's economic scale. Environmental applications of gas-phase heterogeneous catalysis include automotive three-way catalytic converters, introduced in 1975 to comply with U.S. emission standards, which simultaneously convert CO, hydrocarbons (HC), and NOx in exhaust gases using platinum-group metals (Pt, Pd, Rh) supported on ceramic monoliths coated with γ-Al₂O₃. These converters operate at 400-800°C, where Pt and Pd oxidize CO and HC to CO₂ and H₂O (e.g., 2CO + O₂ → 2CO₂), while Rh reduces NOx to N₂ (e.g., 2NO + 2CO → N₂ + 2CO₂), achieving >90% conversion efficiency near stoichiometric air-fuel ratios. The synergy among metals, stabilized by ceria-zirconia oxygen storage components, enables transient operation during engine load changes, significantly reducing urban air pollutants since their mandated adoption.

Liquid-Phase Industrial Processes

Heterogeneous catalysis in liquid-phase processes involves reactions where solid catalysts interact with liquid reactants, often in solid-liquid or immiscible liquid-liquid systems, enabling efficient transformations in industries such as and . These systems leverage the high surface area of solid catalysts to accelerate reactions like hydrotreating and , while addressing challenges such as and . Key examples include for fuel purification, olefin for plastics production, and selective for food applications, where catalyst and reactor configuration play critical roles in overcoming barriers. Hydrodesulfurization (HDS) is a vital liquid-phase for removing impurities from feedstocks, converting organosulfur compounds like dibenzothiophenes into (H₂S) to meet stringent environmental regulations on fuel content, typically below 10 . Industrial HDS employs cobalt-promoted catalysts supported on alumina (CoMo/Al₂O₃), which operate under moderate conditions of 300–400°C and 2–6 in fixed-bed trickle- reactors, where liquid hydrocarbons downward over the catalyst bed while is co-fed. The active sites on the edges of MoS₂ slabs, promoted by , facilitate two primary pathways: direct desulfurization (DDS) via C-S bond cleavage and (HYD) to saturate aromatic rings before desulfurization, with DDS being more selective for refractory species. This is essential for producing and , significantly reducing SOₓ emissions from engines. Polymerization of olefins, particularly to , exemplifies heterogeneous catalysis in liquid or phases using Ziegler-Natta catalysts. Discovered in the 1950s by and , these catalysts consist of (TiCl₄) supported on (MgCl₂), activated by alkylaluminum cocatalysts like triethylaluminum (AlEt₃), enabling stereospecific at ambient to moderate temperatures (50–80°C) and pressures (1–5 MPa) in or gas-liquid systems. The MgCl₂ support mimics the crystalline structure of TiCl₃ used in early generations, providing active Ti³⁺ or Ti²⁺ sites that coordinate with π-electrons for isospecific insertion, yielding high-molecular-weight isotactic with exceeding 98%. This breakthrough revolutionized the industry, enabling the production of millions of tons annually for , automotive, and consumer goods, with catalyst activity reaching 30–60 kg PP/g cat in modern formulations. Hydrogenation of edible oils represents a classic liquid-phase application, converting unsaturated vegetable oils into semi-solid fats for and shortenings using as a heterogeneous catalyst. Developed by Murray Raney in 1926, this sponge-like catalyst, prepared by aluminum from a Ni-Al with , provides high surface area (up to 100 m²/g) for selective of carbon-carbon double bonds at 120–180°C and 0.1–0.5 MPa pressure in batch or continuous stirred-tank reactors. The process improves oil stability, , and texture by partially saturating polyunsaturated fatty acids while minimizing trans-fat formation in optimized conditions, with billions of pounds of oils processed annually worldwide. 's insolubility allows easy separation and reuse, making it industrially dominant despite alternatives like . In solid-liquid heterogeneous catalysis, limitations significantly influence reaction efficiency, particularly in systems where reactants must diffuse from the bulk liquid to active sites on the surface. External , governed by and , can reduce rates in fixed-bed reactors, where larger pellets (1–10 mm) minimize but exacerbate diffusion barriers, often quantified by the ; internal pore , assessed via the Thiele modulus (Φ), further limits effectiveness in microporous supports when Φ > 1, yielding η < 1. reactors mitigate these issues by suspending fine particles (10–100 μm) in the liquid, enhancing gas-liquid-solid contact and coefficients (k_L a up to 0.1–1 s⁻¹), though they require for recovery and are suited for exothermic reactions like HDS. Fixed-bed trickle-bed reactors, conversely, offer continuous operation for high-throughput processes but demand careful design to avoid flooding or channeling. Deactivation in liquid environments, often from deposition or , can be briefly addressed through periodic regeneration. Liquid-liquid heterogeneous catalysis employs phase-transfer mechanisms with insoluble solid-supported catalysts to facilitate reactions across immiscible phases, such as aqueous-organic systems. Supported phase-transfer catalysts (PTCs), typically salts immobilized on polymers, clays, or carbon materials, transfer anionic reactants (e.g., halides or nucleophiles) from aqueous to phases via ion-exchange, enabling efficient biphasic reactions without soluble PTCs. These heterogeneous systems, exemplified by salts on silica or , promote processes by allowing easy catalyst recovery through , reducing waste, and maintaining activity over multiple cycles in transformations like or oxidation. Advantages include enhanced and selectivity in industrial liquid-liquid extractions, with recent advances in chiral supported PTCs for asymmetric .

Emerging and Specialized Applications

Heterogeneous catalysis has expanded into emerging applications that address challenges, such as production and carbon capture, leveraging to enhance and selectivity. In electrocatalysis, supported on carbon (Pt/C) remains a for the (ORR) in fuel cells, where it facilitates the four-electron reduction of O₂ to water under acidic conditions. The activity of Pt-based catalysts follows a volcano plot, correlating optimal performance with the of oxygen intermediates; Pt sits near the peak, balancing adsorption strength to minimize , typically around 0.4 V versus the . Efforts to reduce Pt loading while maintaining have led to core-shell alloys like Pt₃Ni, which shift the d-band center for improved and exceeding 100,000 cycles in accelerated tests. Photocatalysis represents another frontier, with titanium dioxide (TiO₂) enabling water splitting for hydrogen production under ultraviolet irradiation, driven by its wide band gap of approximately 3.0–3.2 eV that generates electron-hole pairs for redox reactions. In the seminal demonstration, TiO₂ anatase electrodes achieved stoichiometric H₂ and O₂ evolution from water, highlighting its potential for solar-driven fuel generation, though quantum yields remain below 10% due to rapid charge recombination. Band gap engineering, such as doping with nitrogen or coupling with narrower-gap semiconductors like CdS, extends absorption into the visible spectrum, boosting H₂ production rates to over 100 μmol·g⁻¹·h⁻¹ in modified TiO₂ systems. These modifications enhance charge separation, as evidenced by prolonged photoluminescence lifetimes, making TiO₂-based photocatalysts viable for scalable photoelectrochemical cells. In biomass conversion, acidic zeolites catalyze the of to glucose, offering a sustainable route to platform chemicals from lignocellulosic feedstocks without homogeneous acids. H-type zeolites like H-β or H-ZSM-5 provide Brønsted acid sites that cleave β-1,4-glycosidic bonds, achieving glucose yields up to 50% at 150–200°C in media, surpassing traditional processes in recyclability. For production, () and iron ()-based catalysts in Fischer-Tropsch synthesis convert derived from into liquid hydrocarbons, with favoring high selectivity to diesel-range paraffins (chain length C₁₀–C₂₀) at 200–250°C and 20–30 . catalysts, often promoted with , excel in handling CO₂-contaminated , yielding oxygenated compounds alongside hydrocarbons, with Anderson-Schulz-Flory chain growth probabilities around 0.8–0.9 for applications. CO₂ utilization via heterogeneous catalysis mitigates greenhouse emissions by converting it to value-added products like through with H₂. Cu/ZnO catalysts, typically stabilized with Al₂O₃, promote the reaction at 200–300°C and 30–50 , with active sites at the Cu-ZnO interface facilitating intermediates for selectivity over 90%. These catalysts achieve space-time yields up to 1 g·mL⁻¹·h⁻¹, drawing from industrial processes but optimized for CO₂ feeds, where ZnO enhances CO₂ adsorption and spillover of . Recent advances include metal-organic framework (MOF)-derived catalysts, pyrolyzed in the to yield high-surface-area metal nanoparticles embedded in porous carbon, improving dispersion and resistance to in reactions like CO₂ reduction. For instance, ZIF-8-derived Co/N-doped carbon exhibits ORR activity comparable to Pt/C, with half-wave potentials around 0.85 V, due to synergistic metal-nitrogen sites. Similarly, enzyme-mimetic nanomaterials, such as nanozymes, replicate or activities in heterogeneous systems, catalyzing oxidative transformations with turnover frequencies exceeding 10⁴ s⁻¹, offering robust alternatives to fragile biocatalysts in industrial settings. These developments underscore heterogeneous catalysis's role in bridging fundamental mechanisms to innovative, sustainable technologies.

References

  1. [1]
  2. [2]
    Heterogeneous catalysis for green chemistry based on nanocrystals
    Heterogeneous catalysis occurs when the reactant molecules are adsorbed to the surface of the catalyst. In surface science, well-defined single-crystal surfaces ...<|control11|><|separator|>
  3. [3]
    Heterogeneous Catalytic Chemistry by Example of Industrial ...
    Oct 9, 2012 · A heterogeneous metal catalyst typically consists of the active metal component, promoters, and a support material.Economic Background · Industrial Catalytic... · Selective Oxidation...Missing: review | Show results with:review
  4. [4]
    Heterogeneous Catalysis - an overview | ScienceDirect Topics
    Heterogeneous catalysis is defined as a process that involves a catalyst in a different phase, typically solid, which promotes reactions occurring in gaseous ...
  5. [5]
    Fluid catalytic cracking: recent developments on the grand old lady ...
    Sep 18, 2015 · We will discuss several trends in FCC catalysis research by focusing on ways to improve the zeolite structure stability, propylene selectivity and the overall ...
  6. [6]
    Electrochemical Ammonia Synthesis: The Energy Efficiency Challenge
    Dec 13, 2024 · The Haber-Bosch synthesis was the first heterogeneous catalytic system employed in the chemical industry and is still in use today.
  7. [7]
    Synthesis of sulfuric acid by the contact process. A student ...
    heterogeneous catalysis, the synthesis of sulfuric acid by the contact process, as we shall carry it out before our students. The oxidation of sulfur ...
  8. [8]
    18.7 Catalysis – Chemistry Fundamentals - UCF Pressbooks
    A heterogeneous catalyst is a catalyst that is present in a different phase (usually a solid) than the reactants. Such catalysts generally function by ...
  9. [9]
    Heterogeneous Catalysis
    ### Summary of Heterogeneous Catalysis Definition and Scope
  10. [10]
    Catalysis in industry
    A heterogeneous catalyst on the other hand is in a different phase to the reactants and products, and is often favoured in industry, being easily separated from ...Heterogeneous Catalysis · Aluminium Oxide, Silicon... · Bifunctional Catalysts<|separator|>
  11. [11]
    Introduction: Bridging the Gaps: Learning from Catalysis across ...
    May 10, 2023 · Traditionally, catalysis has been subdivided into three fields: heterogeneous catalysis, homogeneous catalysis, and biocatalysis, and it can ...
  12. [12]
    Support–Activity Relationship in Heterogeneous Catalysis for ...
    Heterogeneous catalysis, in particular, is responsible of almost 90% of the total volume of chemical production each year [2], making it one of the most ...
  13. [13]
    Heterogeneous Catalysis Before 1934 - American Chemical Society
    ysis came in 1817 when Sir Humphry Davy discovered that the introduction of a hot platinum wire into a mixture of air and coal gas led the platinum to.
  14. [14]
    (PDF) The History of Catalysis – From the Beginning to Nobel Prizes
    Aug 9, 2025 · Although the effects of catalysis are known from very ancient times, the understanding of the phenomena started only in the 18th century and ...
  15. [15]
    Paul Sabatier – Facts - NobelPrize.org
    Around 1897 Paul Sabatier developed a method for causing unsaturated organic substances to absorb hydrogen and form new organic compounds.
  16. [16]
    Haber-Bosch Process - an overview | ScienceDirect Topics
    The Haber–Bosch process is a well-developed ammonia synthesis technology with the first-generation heterogeneous catalytic system for the main industrial ...
  17. [17]
    Irving Langmuir's Contribution to Catalysis - ACS Publications
    Jun 3, 1983 · Irving Langmuir's contributions to the advancement of catalysis research were many and varied, ranging from important theoretical concepts ...
  18. [18]
    [PDF] Introduction 1 - Wiley-VCH
    Feb 15, 2010 · Thus, by the mid - 1930s the literature described the ion exchange, adsorption, molecular sieving and structural properties of zeolite minerals ...
  19. [19]
    Automobile pollution control using catalysis - ScienceDirect.com
    The national effort to control this automobile pollution can be traced to the 1970 Clean Air Act, which required a 90 percent reduction in CO, HC and NOx ...
  20. [20]
    Computer-aided design of novel heterogeneous catalysts—A ...
    In the present study we investigated the adsorption energies of the NO and SOx on various ion-exchanged ZSM-5 by using a combinatorial computational chemistry.
  21. [21]
    A minireview on the synthesis of single atom catalysts
    Mar 24, 2022 · The catalyst with a single atom on its surface is called a single atom catalyst (SAC). The concept of SAC was raised up by Zhang et al. in 2011, ...Missing: history | Show results with:history
  22. [22]
    [PDF] Gerhard Ertl - Nobel Lecture
    Principle of heterogeneous catalysis. Figure 1. Energy diagram illustrating the progress of a chemical reaction with and without a catalyst. Page ...
  23. [23]
    Equilibria and Isosteric Heat of Adsorption of Methane on Activated ...
    Dec 9, 2020 · As a rule of thumb, the heat of adsorption of 80 kJ/mol or more indicates chemisorption, and lower values indicate physisorption. (41,42) The ...Introduction · Results and Discussion · Materials and Methods · References
  24. [24]
    Classical and new insights into the methodology for characterizing ...
    Nov 1, 2023 · This situation means that physisorption predominates at low temperatures and chemisorption at higher temperatures.
  25. [25]
    THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS ...
    Replacing the Langmuir Isotherm with the Statistical Thermodynamic Fluctuation Theory. The Journal of Physical Chemistry Letters 2024, 15 (13) , 3683-3689 ...
  26. [26]
    [PDF] Kinetics of Surface Catalysed Reactions
    The Eley-Rideal mechanism does not lead to a maximum in the rate as PA and/or PB are changed, a property which can be used to distinguish this mechanism from ...
  27. [27]
    1.22: Kinetics of Catalysis - Chemistry LibreTexts
    Sep 23, 2022 · In the Eley-Rideal mechanism, only one species adsorbs onto the catalyst surface. An example of such a reaction is the partial oxidation of ...
  28. [28]
    Practical Considerations for Understanding Surface Reaction ...
    Oct 30, 2024 · (10) Sometimes, physisorption that precedes chemisorption is referred to as the “precursor state” (i.e., precursor state and physisorption may ...
  29. [29]
    Active Sites in Heterogeneous Catalytic Reaction on Metal ... - MDPI
    Oct 20, 2018 · Active sites play an essential role in heterogeneous catalysis and largely determine the reaction properties. Yet identification and study ...
  30. [30]
  31. [31]
    CO oxidation on Pt and Pd model catalysts - RSC Publishing
    Jun 7, 2017 · In a MvK–LH mechanism, CO initially adsorbs on the surface of the oxide, while in a MvK–ER mechanism CO(g) reacts directly with the lattice ...
  32. [32]
  33. [33]
    [PDF] Modern Methods in Heterogeneous Catalysis Research
    First model for the desorption rate: Polanyi-Wigner-equation. Definition: desorption order = exponent n of Θ r n. Redhead (1963). Page 6. Left: 2D gas with very ...
  34. [34]
    [PDF] “Modern Methods in Heterogeneous Catalysis” Klaus Christmann
    Principally, one has to distinguish between associative and dissociative as well as between activated and non-activated adsorption. 2. Basics: The process of ...
  35. [35]
    Analysis of Temperature-Programmed Desorption via Equilibrium ...
    We present a method to extract the distribution of adsorption energies from TPD spectra, and we rationalize the energy resolution of TPD experiments.
  36. [36]
    Metal Catalysts for Heterogeneous Catalysis: From Single Atoms to ...
    Apr 16, 2018 · In the following sections of this Review, we will summarize the catalytic applications of single atoms, nanoclusters, and nanoparticles for ...
  37. [37]
    Heterogeneous Catalysis on Metal Oxides - MDPI
    This paper attempts to review the major current industrial applications of supported and unsupported metal oxide catalysts.
  38. [38]
    Metal oxides and their roles in heterogeneous catalysis
    Jun 21, 2023 · This review focuses on heterogeneous surface catalysis, mainly with the use of metal-oxides, with emphasis on La-based perovskites, due to ...
  39. [39]
    Zeolites in catalysis: sustainable synthesis and its impact on ...
    Aug 29, 2022 · High surface area and porosity are desired properties for heterogeneous catalysts. Zeolites are outstanding candidates of porous materials ...
  40. [40]
    A Supramolecular View on the Cooperative Role of Brønsted and ...
    A systematic molecular level and spectroscopic investigation is presented to show the cooperative role of Brønsted acid and Lewis acid sites in zeolites for ...Introduction · Materials and Methods · Results and Discussion · Conclusion
  41. [41]
    Silica samurai: Aristocrat of energy and environmental catalysis
    Aug 18, 2022 · Silica (SiO2) has been widely used as a support for various heterogeneous catalysts because of its Earth abundance, low cost, and large specific ...
  42. [42]
    Dark material with a bright future: Carbon as support in future ...
    Nov 1, 2023 · Carbon supports hold great potential for future sorbents and catalysts; a personal perspective is given.
  43. [43]
    Activation of surface lattice oxygen in single-atom Pt/CeO2 for low ...
    Dec 15, 2017 · A thermally stable catalyst, atomically dispersed Pt 2+ on CeO 2 , can become active for CO oxidation at 150°C after steam treatment at 750°C.
  44. [44]
    Design strategies of Pt-based electrocatalysts and tolerance ...
    Feb 7, 2023 · Based on the above, PtRu bimetallic alloys have been considered as one of the most promising catalysts in PEMFCs. In the bifunctional mechanism, ...
  45. [45]
    Uncovering Cooperative Redox Enhancement Effects in Bimetallic ...
    Oct 21, 2025 · Typically, bimetallic systems are considered to offer enhanced catalytic performance due to an often poorly described synergistic effect that ...
  46. [46]
    Recent Developments in the Synthesis of Supported Catalysts
    Impregnation can be performed to incipient wetness, whereby only the pores of the support are filled with precursor solution, to prevent deposition on the ...Missing: seminal | Show results with:seminal
  47. [47]
    A Review of Preparation Methods for Heterogeneous Catalysts
    Aug 10, 2025 · This review presents the detailed background, principle and mechanism of each preparation method. The advantages and limitations of each method ...
  48. [48]
    “Traditional” Sol-Gel Chemistry as a Powerful Tool for the ...
    The sol-gel method is an attractive synthetic approach in the design of advanced catalytic formulations that are based on metal and metal oxide.
  49. [49]
    A Seed‐Mediated Approach for the Preparation of Modified ...
    While nanoparticles are essential components of heterogeneous catalysts, colloidal methods are rarely employed for their preparation. We have employed a seed ...
  50. [50]
  51. [51]
    Achieving volatile potassium promoted ammonia synthesis via ...
    Apr 22, 2023 · Potassium oxide (K2O) is used as a promotor in industrial ammonia synthesis, although metallic potassium (K) is better in theory.
  52. [52]
    How to Measure the Reaction Performance of Heterogeneous ...
    Dec 18, 2019 · In this Commentary, we outline the fundamentals in the reliable measurement and analysis of reactivity and kinetic data.
  53. [53]
    “Turning Over” Definitions in Catalytic Cycles | ACS Catalysis
    Nov 8, 2012 · “The standard turnover frequency, TOF°, is the TOF measured at a hypothetical 1 M standard concentration of reactants and products (or 105 Pa in ...Introduction · Typical Misuses of the TOF · “Turning Over” Definitions · Conclusion
  54. [54]
    Introduction: Operando and In Situ Studies in Catalysis and ...
    Jul 10, 2024 · A review. Heterogeneous catalysts undergo thermally and/or adsorbate-induced dynamic changes under reaction conditions, which consequently ...Author Information · Biographies · References
  55. [55]
    Theoretical Heterogeneous Catalysis: Scaling Relationships and ...
    In this review, the detailed theory behind scaling relationships is discussed, and the existence of these relationships for catalytic materials ranging from ...<|separator|>
  56. [56]
    Heterogeneous Catalyst Deactivation and Regeneration: A Review
    This review on deactivation and regeneration of heterogeneous catalysts classifies deactivation by type (chemical, thermal, and mechanical) and by mechanism.
  57. [57]
    On the mechanism of sulfur poisoning of platinum catalysts
    On the mechanism of sulfur poisoning of platinum catalysts. Author links open ... Adjustable kinetics in heterogeneous photocatalysis demonstrating the ...
  58. [58]
    State-of-the-Art Review of Fluid Catalytic Cracking (FCC) Catalyst ...
    Fluid catalytic cracking (FCC) is the workhorse of modern crude oil refinery. Its regenerator plays a critical role in optimizing the overall profitability.
  59. [59]
    Oxychlorination Redispersion of Pt Catalysts: Surface Species and ...
    Consistent with Lee and Kim's results, oxychlorination redispersion of commercial reforming catalysts is typically performed at 510–530 °C for about 4 h [19].
  60. [60]
    Sorption enhanced dimethyl ether synthesis under industrially ...
    In this work, SEDMES is demonstrated experimentally on a bench-scale reactor with pressure swing regeneration under industrially relevant conditions.
  61. [61]
    Evaluate your Catalyst Replacement Economics Part 1 | AIChE
    The decision to replace catalyst in a fixed bed reactor is a balance of costs and benefits. Replacement timing may be affected by synergies with other ...
  62. [62]
    Pyrite-induced uv-photocatalytic abiotic nitrogen fixation - Nature
    Oct 25, 2019 · It uses hydrogen and an iron-based metal catalyst, under high temperature (400–500 °C) and pressures (150–250 atm).
  63. [63]
    Zr in situ defects-created carbon nitride for efficient electrochemical ...
    For now, NH3 is predominantly prepared by a century-old Haber-Bosch process, which is operating at high temperature (400–500 °C) and pressure (200–400 atm), ...
  64. [64]
    Facilitating green ammonia manufacture under milder conditions
    Iron-based catalysts, such as magnetite and wüstite, constitute the first-generation of Haber–Bosch catalysts, modified forms of which remain in industrial use ...Missing: alternatives | Show results with:alternatives
  65. [65]
    A comparative analysis of the mechanisms of ammonia synthesis on ...
    Nov 3, 2021 · Ammonia synthesis mechanism on Ru. Ruthenium is a more active catalyst for ammonia synthesis than iron, [31] but ruthenium is more expensive, ...
  66. [66]
    Catalysts for selective hydrogenation of acetylene: A review
    A common method is to use a noble metal palladium catalyst to catalyze the selective hydrogenation of acetylene to ethylene at relatively high temperatures (100 ...
  67. [67]
    Breaking the inverse relationship between catalytic activity and ...
    The partial hydrogenation of acetylene impurities in ethylene is industrially important because acetylene acts as a catalyst poison in downstream ethylene ...
  68. [68]
    Rationally designed laterally-condensed-catalysts deliver robust ...
    Dec 10, 2024 · Rationally designed laterally-condensed-catalysts deliver robust activity and selectivity for ethylene production in acetylene hydrogenation.
  69. [69]
    Facile synthesis of highly active Rh/Al2O3 steam reforming catalysts ...
    Dec 5, 2016 · Generally, methane reforming is conducted industrially over a Ni-based catalyst at temperatures above 800 °C in a packed bed tubular reactor.
  70. [70]
    Kinetics study and modelling of steam methane reforming process ...
    Feb 2, 2017 · Kinetic rate data for steam methane reforming (SMR) coupled with water gas shift (WGS) over an 18 wt. % NiO/α-Al2O3 catalyst are presented ...
  71. [71]
    Insight into the influence of Re and Cl on Ag catalysts in ethylene ...
    Nov 6, 2024 · Commercial ethylene epoxidation catalysts consist of α-alumina supported Ag particles and usually contain a mixture of promoters.
  72. [72]
    Influence of Ag particle size and Ag : Al2O3 surface ratio in catalysts ...
    Feb 15, 2024 · Ethylene epoxidation is catalyzed by α-alumina supported silver catalysts. The influence of silver particle size has been a topic of debate.
  73. [73]
    The evolution of catalytic converters | Feature - RSC Education
    May 31, 2011 · From early smog problems to modern concerns about air pollution, catalysts pave the way in controlling the emissions from combustion engines.
  74. [74]
    Catalytic NOx Abatement Systems for Mobile Sources: From Three ...
    (109) Typically the presence of Rh is needed for the reduction of NOx, whereas the catalytic properties of Pt and Pd are useful to oxidize CO and HC. However, ...
  75. [75]
    Catalytic Converter - an overview | ScienceDirect Topics
    Catalytic converters aim to convert gaseous pollutants such as carbon monoxide (CO), nitrogen oxides (NOx), and hydrocarbons (HCs) into carbon dioxide (CO2), ...
  76. [76]
  77. [77]
    Experimental and first-principles investigation on how support ...
    Aug 1, 2024 · One class of the Ziegler–Natta catalysts (ZNC) – the TiCl4/MgCl2 having triethyl aluminum (AlEt3), has been widely utilized during ethylene ...Experimental Results · Adsorption Of Ticl On... · Alkylation Of Ticl Species...Missing: seminal | Show results with:seminal
  78. [78]
    The discovery and progress of MgCl 2 -supported TiCl 4 catalysts
    Nov 14, 2003 · TiCl3 catalysts, established by Ziegler and Natta in the 1950s, led to the births of the polyolefin industries. However, the activities and ...
  79. [79]
    [PDF] Raney® Nickel: A Life-Changing Catalyst
    Raney nickel is a heterogeneous hydrogenation catalyst. It promotes the reaction of hydrogen with organic compounds (compounds that contain carbon), and, like ...Missing: edible | Show results with:edible
  80. [80]
    [PDF] A Review of Mass Transfer Controlling the Reaction Rate in ...
    Jul 7, 2011 · Introduction. Mass transfer limitations play an important role on the rate of reaction; the rate of conversion and product formation, ...
  81. [81]
    State Of The Art Of Supported Phase Transfer‐Catalysts Onium Salt ...
    Jan 11, 2024 · This review describes the advances of heterogenous phase-transfer catalysts onium-based and their importance from a sustainable point of view.Abstract · Introduction · Heterogeneous catalysts · Supported phase-transfer...
  82. [82]
    Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell ...
    In the present paper, we have introduced a method to use density functional theory calculations to estimate the thermochemistry for electrochemical reactions ...Missing: seminal | Show results with:seminal
  83. [83]
    Simultaneous Enhancement of the Activity and Durability of the ...
    Apr 3, 2025 · The observed enhancement in the ORR performance of Pd3Mo@Pt/C likely stems from the modification of the electronic structure of the Pt surface ...Missing: seminal | Show results with:seminal
  84. [84]
    Band gap engineering of TiO2 through hydrogenation - AIP Publishing
    Nov 21, 2014 · The band gap of rutile TiO 2 is reduced remarkably through high-temperature annealing in hydrogen atmosphere, and the absorption of visible light of the ...Missing: H2 seminal
  85. [85]
    TiO2 as a Photocatalyst for Water Splitting—An Experimental and ...
    Here, we review state-of-the-art experimental and theoretical research on TiO 2 based photocatalysts and identify challenges that have to be focused on to ...Missing: seminal | Show results with:seminal
  86. [86]
    A review of advanced catalyst development for Fischer–Tropsch ...
    May 9, 2014 · This review paper summarises recent developments in FT-catalyst design with regards to optimising catalyst activity and selectivity towards synthetic fuels.1. Introduction · 2.5 Bimetallic Fe/co... · 6.4 Zeolite Supported...
  87. [87]
    Active sites for CO2 hydrogenation to methanol on Cu/ZnO catalysts
    Mar 24, 2017 · We report a direct comparison between the activity of ZnCu and ZnO/Cu model catalysts for methanol synthesis. By combining x-ray photoemission ...
  88. [88]
    Review Metal-organic framework (MOF)-derived catalysts for Fischer ...
    Review. Metal-organic framework (MOF)-derived catalysts for Fischer-Tropsch synthesis: Recent progress and future perspectives · Abstract · Graphical Abstract.
  89. [89]
    Mimicking Enzymes: The Quest for Powerful Catalysts from Simple ...
    Sep 5, 2021 · This review concerns the development of various types of enzyme mimics, namely polymeric and dendrimeric, supramol., nanoparticulate and ...