Fact-checked by Grok 2 weeks ago

Catalyst poisoning

Catalyst poisoning refers to the partial or complete deactivation of a catalyst through the strong chemisorption of impurities, reactants, or byproducts onto its active sites, thereby inhibiting the catalyst's ability to facilitate chemical reactions. This process is distinct from other deactivation mechanisms like sintering or coking, as it primarily involves chemical interactions that block or alter the catalyst's surface properties. In industrial applications, such as fluid catalytic cracking (FCC) and ammonia synthesis, catalyst poisoning can result in substantial efficiency losses and annual costs exceeding billions of dollars due to reduced productivity and the need for frequent catalyst replacement or regeneration. The mechanisms of catalyst poisoning typically involve irreversible or strongly bound adsorption that occupies active sites, modifies the electronic or geometric structure of the catalyst surface, or physically obstructs reactant access. Common poisons include sulfur compounds, carbon monoxide (CO), nitrogen oxides, heavy metals, and alkali metals, which can originate from feedstocks or form as reaction intermediates. Poisoning can be classified as reversible (e.g., temporary blockage by nitrogen compounds in FCC catalysts, recoverable through heating or time) or irreversible (e.g., permanent sulfur adsorption on metal catalysts like nickel in steam reforming), with the latter often requiring advanced regeneration techniques such as oxidation or chemical treatments. While poisoning generally diminishes catalytic activity and selectivity, it can sometimes be intentionally induced to enhance reaction specificity in selective catalytic processes. Notable examples illustrate the broad impact across catalysis types: in fuel cells, CO poisoning reduces the electrochemical surface area of platinum catalysts by strong adsorption; sulfur species deactivate nickel-based catalysts in methanation at concentrations as low as 5 ppm H₂S; and alkali metals like sodium sulfate poison vanadium-based selective catalytic reduction (SCR) catalysts for NOx abatement, though additives such as molybdenum oxide can mitigate effects. In enzymatic systems, analogous poisoning occurs when inhibitors bind to enzyme active sites, underscoring the universality of this phenomenon in biological catalysis as well. Addressing catalyst poisoning remains a key challenge in chemical engineering, driving research into poison-resistant materials, feedstock purification, and innovative regeneration methods like supercritical fluid extraction to extend catalyst lifespan and promote sustainable industrial practices.

Fundamentals

Definition and Overview

Catalyst poisoning refers to the deactivation or significant reduction in the activity of a catalyst caused by the strong adsorption of impurities, byproducts, or contaminants onto its active sites, thereby inhibiting the catalyst's ability to facilitate chemical reactions. This process typically involves trace levels of substances that bind more strongly to the catalyst surface than the intended reactants, leading to a partial or complete loss of catalytic function. In heterogeneous catalysis, which is the most common context for poisoning, these active sites are specific locations on the catalyst's surface where reaction intermediates form and transform. A key distinction exists between catalyst poisoning and other forms of deactivation, such as sintering or coking; poisoning specifically arises from the chemical interaction with trace substances that target active sites, whereas sintering involves the agglomeration of catalyst particles reducing surface area, and coking results from carbon deposition that physically covers the surface. Poisoning effects can be temporary (reversible) or permanent (irreversible), depending on whether the poison can be desorbed or removed under operational conditions—reversible cases allow recovery of activity through changes in temperature, pressure, or gas composition, while irreversible poisoning forms stable compounds that permanently alter the catalyst structure. The phenomenon of catalyst poisoning was first systematically recognized in the early 20th century during the development of the Haber-Bosch process for ammonia synthesis, where trace amounts of oxygen or sulfur impurities in the feed gas were observed to drastically diminish the efficiency of iron-based catalysts by blocking nitrogen dissociation sites. This early identification highlighted the sensitivity of industrial catalysts to feed purity and spurred advancements in purification techniques to mitigate such effects. In a basic illustration of poisoning, the catalyst surface features exposed active sites that normally adsorb reactant molecules to lower activation energies; however, poison molecules, due to their higher binding affinity, occupy these sites instead, forming a monolayer coverage that sterically or electronically hinders reactant access and reduces the overall reaction rate. This site-blocking mechanism underscores why even parts-per-million levels of poisons can profoundly impact performance.

Importance in Industrial Processes

Catalyst poisoning imposes substantial economic burdens on industrial processes reliant on catalysis, primarily through accelerated deactivation that shortens catalyst lifespan and triggers costly replacements and downtime. In petroleum refining, for instance, poisoning mechanisms such as sulfur or metal contaminants reduce operational efficiency, leading to increased raw material consumption and utility expenses as aged catalysts demand higher inputs to maintain output. The global market for refining catalysts, valued at approximately $7 billion as of 2025, underscores the scale of these investments, with deactivation contributing significantly to recurring expenditures on procurement and maintenance. Environmentally, catalyst poisoning exacerbates pollutant emissions in key applications like automotive exhaust systems, where deactivation of three-way catalytic converters allows greater release of carbon monoxide, nitrogen oxides, and unburned hydrocarbons into the atmosphere. This directly contravenes stringent regulations such as the Euro 6 emission standards, effective from 2014, which mandate ultra-low pollutant levels to mitigate air quality degradation and public health risks. In sectors like refining and power generation, unchecked poisoning similarly elevates greenhouse gas outputs, complicating compliance with global environmental targets. Operationally, poisoning presents ongoing challenges by diminishing catalytic activity and selectivity, which in turn raises energy consumption and process temperatures to compensate for reduced performance. In petrochemical industries, this often necessitates catalyst replacements every 1-3 years, depending on feedstock quality and operating conditions, resulting in frequent shutdowns that disrupt production continuity. Such inefficiencies not only amplify operational costs but also strain resource allocation in high-volume processes like hydrocracking and reforming. On a broader scale, catalyst poisoning impedes advancements in sustainable catalysis, particularly in green hydrogen production where contaminants like sulfur compounds or particulates deactivate electrocatalysts in electrolyzers, lowering efficiency and scalability. Recent studies as of 2023 highlight how poisoning challenges scalability for net-zero goals in refining by 2050. This directly affects progress toward United Nations Sustainable Development Goal 7, which aims for affordable and clean energy by 2030, as reliable catalyst performance is essential for cost-effective hydrogen generation from renewable sources.

Mechanisms

Adsorption-Based Poisoning

Adsorption-based poisoning occurs when impurities, known as poisons, adsorb onto the surface of a catalyst, thereby blocking access to active sites and reducing the catalyst's effectiveness. This mechanism is distinct from deeper structural alterations and primarily involves interactions at the catalyst surface. Poisons can adsorb through two main types: physisorption and chemisorption. Physisorption relies on weak van der Waals forces, typically with adsorption energies of 20–40 kJ/mol, making it reversible and non-specific, as the poison molecules remain intact and can form multilayers. In contrast, chemisorption involves stronger covalent or ionic bonds, with energies ranging from 100–500 kJ/mol, often leading to irreversible deactivation under operating conditions, as the poison dissociates or forms stable surface complexes specific to active sites such as metal atoms or acid centers. The site blocking model describes how poisons reduce the availability of active sites for reactants, following the principles of competitive adsorption. In this framework, the fractional coverage of the surface by the poison, θ, is given by the Langmuir isotherm adapted for poisons and reactants: \theta = \frac{K_p P_p}{1 + K_p P_p + \sum K_i P_i} where K_p is the adsorption equilibrium constant for the poison, P_p its partial pressure, and the summation term accounts for reactants (K_i P_i). This coverage directly lowers the turnover frequency by limiting reactant adsorption, with even low poison concentrations (e.g., parts per million) causing significant deactivation if K_p \gg K_i, as poisons preferentially occupy sites due to higher binding affinity. For instance, sulfur species like H₂S exhibit strong chemisorption on nickel catalysts, blocking multiple adjacent sites per poison molecule. The distribution of poison on the catalyst particle—homogeneous or heterogeneous—depends on the relative rates of poison adsorption/reaction and intraparticle diffusion, often quantified by the Thiele modulus (φ), which compares reaction rate to diffusion rate. In slow reactions (low φ, diffusion-controlled), poisons distribute uniformly (homogeneous poisoning) throughout the particle, leading to gradual activity loss proportional to poison loading. Conversely, in fast reactions (high φ, reaction-controlled), poisons accumulate near the external surface (heterogeneous or pore-mouth poisoning), causing rapid initial deactivation but potentially preserving inner sites longer. This effect is critical in porous catalysts, where slow diffusion exacerbates surface-selective blocking. Diagnostic methods for adsorption-based poisoning focus on quantifying adsorption strength and coverage. Temperature-programmed desorption (TPD) is a primary technique, where the catalyst is saturated with the poison or probe molecule, then heated linearly under inert gas flow while monitoring desorbed species via mass spectrometry or thermal conductivity detection. Desorption peaks reveal binding energies: low-temperature peaks indicate physisorbed species (reversible), while high-temperature peaks (>300°C) signify chemisorbed poisons (irreversible), allowing estimation of deactivation extent. For example, TPD of CO on platinum catalysts shows peak shifts correlating with poison-induced site loss. Other methods include infrared spectroscopy for surface species identification, but TPD provides direct insight into adsorption reversibility without altering catalyst composition.

Structural and Chemical Deactivation

Structural and chemical deactivation in catalyst poisoning involves irreversible alterations to the catalyst's composition and morphology, distinct from temporary surface occupation. Poisons, such as sulfur species, react with active sites to form stable compounds that embed into the catalyst lattice or surface, fundamentally changing its chemical nature and reducing long-term efficacy. A primary mechanism is chemical transformation, where poisons chemically bind to metal sites, creating thermodynamically stable phases. For instance, (H₂S) reacts with metallic in reforming catalysts via the sulfidation reaction Ni + H₂S → + H₂, forming that passivates the active metal and prevents further catalysis. This process is irreversible under typical operating conditions due to the high stability of the metal sulfide, often requiring high-temperature regeneration to reverse. Similar transformations occur with other metals, such as palladium forming PdS, leading to a loss of metallic character and diminished adsorption capacity for reactants. These chemical changes often induce structural effects, including lattice strain and phase transitions that promote sintering. Sulfur incorporation into the metal lattice creates compressive or tensile strain, distorting the crystal structure and altering bond lengths, which weakens the catalyst-support interactions. In palladium/alumina catalysts for methane oxidation, sulfur poisoning facilitates metal particle mobility, resulting in sintering, thereby reducing the active surface area and exposing fewer low-coordination sites essential for activity. This sintering is exacerbated by the formation of transient Pd-S phases that lower the energy barrier for atom migration, leading to coalescence of nanoparticles. Poisons also exert ensemble effects by disrupting the spatial arrangement of active sites required for concerted reactions. In hydrogenation processes, which often demand adjacent metal atoms (ensembles) for hydrogen dissociation and spillover, sulfur atoms isolate Pd sites by preferentially adsorbing at ensemble edges, fragmenting larger Pd domains into inactive monomers or dimers. This ensemble disruption is particularly detrimental in bimetallic systems, where poison-induced segregation further diminishes multi-site cooperation. Recent investigations into single-atom catalysts (SACs) highlight how poison coordination profoundly impacts . Coordinating poisons can modify the d-band of the metal , weakening reactant and deactivating the . For example, a 2024 on Au SACs demonstrated that strong adsorption of alters , reducing epoxidation activity by approximately 87% (from 80.2 to 10.4 mol PO·mol Au⁻¹·h⁻¹). These effects underscore the vulnerability of SACs, where even poisons can eliminate the benefits of atomic .

Types of Poisons

Reversible Poisons

Reversible poisons are substances that temporarily inhibit catalytic activity through weak surface interactions, allowing the catalyst to regain full functionality upon removal of the poison without structural alteration. These poisons primarily engage via physisorption, characterized by binding energies below 50 kJ/mol, which distinguish them from stronger chemisorption processes. This low-energy adsorption enables straightforward recovery through adjustments in temperature or pressure, as the desorbed species leave the active sites intact. Common reversible poisons include water vapor interacting with oxide catalysts, where adsorption occurs via hydrogen bonding or van der Waals forces. On silica (SiO₂) surfaces, water exhibits an adsorption heat of approximately 36 kJ/mol, exemplifying the physisorptive nature that permits easy desorption under elevated temperatures. Similarly, carbon monoxide (CO) can act as a reversible poison on noble metal catalysts, such as palladium or platinum, particularly at low temperatures where binding is weakened, facilitating desorption without residue. Unlike irreversible poisons that necessitate catalyst replacement due to permanent site blocking, reversible ones support operational continuity in dynamic environments. In processes like over copper-based catalysts, reversible induces temporary activity , often around 10%, as observed with certain exposures that dissipate upon feedstock adjustment. Detection of such relies on in-situ techniques, including , which reveal reversible spectral shifts—such as changes in vibrational bands—corresponding to adsorption and subsequent desorption. These methods provide insights into surface coverage , aiding optimization.

Irreversible Poisons

Irreversible poisons in catalysis are substances that cause permanent deactivation of through strong or chemical reactions that form stable compounds with the active sites, often characterized by high binding energies exceeding 100 /mol. These poisons, such as reactive species including , bind irreversibly to metal surfaces, forming stable halides like metal chlorides that block catalytic activity without the possibility of regeneration under operational conditions. In contrast to reversible poisons, which allow temporary site occupation that can be reversed by changes in or , irreversible poisons lead to non-recoverable due to their strong interactions. Common examples include sulfur compounds on platinum (Pt) or rhodium (Rh) catalysts, where sulfur chemisorbs to form stable sulfides with binding energies around 150-200 kJ/mol, effectively poisoning active sites in oxidation and reforming processes. Similarly, lead in automotive exhaust catalysts deposits as lead oxide, encapsulating noble metal particles and reducing conversion efficiency; toxicity thresholds for such metals are typically below 1 ppm in feedstreams to avoid significant deactivation. Arsenic represents another potent irreversible poison, particularly in ammonia synthesis, where it forms arsenides on iron-based catalysts through cumulative site blocking. The long-term effects of irreversible poisons involve progressive buildup over time, leading to complete catalyst deactivation as the poison accumulates and alters the surface structure, often requiring catalyst replacement rather than regeneration. This cumulative nature exacerbates operational costs in industrial settings, as seen with arsenic's persistent inhibition in syngas streams for ammonia production. Recent studies from highlight on cobalt-based catalysts, where from feedstocks forms phosphates, resulting in significant activity to coverage and altered acidity. These findings the need for advanced poison-resistant designs in high-impurity feedstocks.

Specific Examples

Palladium Catalysts in Hydrogenation

Palladium-based catalysts are widely utilized in selective hydrogenation reactions, notably for converting alkynes to alkenes, such as the semi-hydrogenation of acetylene to ethylene in industrial petrochemical processes to purify ethylene streams. These catalysts excel due to palladium's strong affinity for hydrogen dissociation and its ability to promote stereoselective addition to carbon-carbon triple bonds while minimizing over-hydrogenation to alkanes. However, Pd catalysts in these reactions are highly sensitive to poisons, including sulfur compounds (e.g., H2S or organosulfur species), carbon monoxide (CO), and halides (e.g., chlorides or bromides from impure feedstocks). Sulfur and CO adsorb preferentially on Pd active sites, competing with hydrogen and substrates, while halides form stable surface complexes that block ensembles of Pd atoms necessary for catalysis. Sulfur poisoning primarily deactivates Pd by blocking hydrogen dissociation sites, which are critical for generating surface hydrogen atoms (H*) that participate in the hydrogenation mechanism. This leads to a sharp decline in reaction rates, as the kinetics of alkyne hydrogenation follow a Langmuir-Hinshelwood model where the rate is proportional to the product of adsorbed hydrogen and substrate coverages: \text{Rate} = k [\ce{H*}] [\text{substrate}] Low sulfur coverages reduce [\ce{H*}] dramatically; for example, experimental and modeling studies on Pd surfaces show hydrogen permeation (analogous to dissociation in catalysis) drops by more than 90% with sulfur exposure equivalent to ~0.1 monolayer coverage. In the context of acetylene hydrogenation over supported Pd/C catalysts, early investigations in the late 1970s revealed CO as a potent poison acting via competitive adsorption on hydrogen-binding sites. Tracer studies demonstrated that adding 0.1–1.0 Torr of CO partially poisons Pd catalysts, reducing activity but to a lesser extent than on Rh or Ir, due to hydrogen's ability to compete for sites and limit full deactivation. The mechanism involves CO blocking H* formation without significantly hindering hydrocarbon adsorption, resulting in reversible inhibition during the reaction. Recent advances in mitigating CO poisoning include alloying Pd with Ag to alter adsorption energetics and enhance tolerance. In Pd-Ag single-atom alloys, the isolated Pd sites embedded in an Ag matrix weaken CO binding relative to pure Pd, improving overall stability and selectivity in hydrogenation of unsaturated compounds like acrolein, with activity enhancements up to several-fold under impure feeds. Such designs draw from permeation studies where Pd77Ag23 membranes exhibit reduced CO poisoning effects compared to pure Pd, blocking only ~20-50% of flux versus near-total inhibition on Pd foil.

Hydrodesulfurization Catalysts

Hydrodesulfurization (HDS) catalysts, primarily cobalt-molybdenum (CoMo) or nickel-molybdenum (NiMo) sulfides supported on alumina, play a critical role in refining processes to reduce sulfur content in diesel fuels to ultra-low levels, such as the 15 parts per million (ppm) limit established by the U.S. Environmental Protection Agency for on-road diesel starting in June 2006. These catalysts facilitate the conversion of organosulfur compounds into hydrogen sulfide and hydrocarbons under high-pressure hydrogen atmospheres, enabling compliance with stringent environmental regulations aimed at minimizing sulfur oxide emissions from combustion. The layered MoS₂ structure in these catalysts, promoted by Co or Ni at edge sites, provides the active sites for sulfur removal, but their performance is highly susceptible to poisoning from impurities in crude oil feedstocks. Key poisons in HDS processes include nitrogen-containing compounds, such as , which act as strong inhibitors by adsorbing onto acidic sites on the catalyst surface, thereby blocking access to sulfur species and significantly reducing HDS activity. Basic nitrogen compounds like compete with reactants for Brønsted acid sites on the alumina support and promoter edges, leading to temporary site occupancy that hampers both the direct desulfurization and hydrogenation pathways. Additionally, metallic impurities originating from crude oil, particularly (V) and (Ni), deposit as sulfides or oxides, with vanadium being especially detrimental as it accumulates and forms vanadium oxide species that deactivate the catalytically active edge sites of the MoS₂ slabs. These metals, present in heavy feedstocks like vacuum gas oil, can displace promoter atoms (Co or Ni) from MoS₂ edges, reducing the number of available active sites and causing irreversible structural changes. In fixed-bed reactors commonly used for industrial HDS, poisons accumulate progressively from the reactor inlet to the outlet, creating concentration gradients that lead to uneven deactivation and reduced overall throughput. Vanadium, for instance, migrates downstream under trickle-flow conditions, forming deposits that encapsulate catalyst particles and inhibit hydrogen sulfide evolution, with deactivation accelerating at significant vanadium loadings relative to molybdenum. This accumulation necessitates guard beds or frequent catalyst replacement to maintain desulfurization efficiency below 15 ppm sulfur in product streams. Recent advancements in HDS resilience, particularly for upgrading bio-oils derived from , highlight the of phenolic compounds as reversible poisons that adsorb onto surfaces but can desorb under optimized pressures and temperatures, allowing improved without severe deactivation. In these contexts, NiMo/alumina s demonstrate improved to phenolics through selective pathways that mitigate formation, offering insights applicable to conventional petroleum HDS for handling oxygenated impurities in co-processed feeds.

Electrocatalysts in Fuel Cells

In proton exchange membrane (PEM) fuel cells, platinum (Pt)-based electrocatalysts serve as the primary anode material for the hydrogen oxidation reaction (HOR), where hydrogen from reformed hydrocarbon fuels often contains impurities such as carbon monoxide (CO) at concentrations up to 100 ppm. These reformed fuels, derived from processes like steam reforming of natural gas or methanol, introduce CO as a byproduct, which preferentially adsorbs onto Pt sites, leading to catalyst poisoning. The strong adsorption of CO on Pt surfaces, with a binding energy of approximately 1.5 eV, blocks active sites essential for the HOR and oxygen reduction reaction (ORR) at the cathode, thereby increasing overpotential and reducing cell efficiency. This CO adsorption results in significant performance degradation, typically manifesting as an anode overpotential loss of 50-200 mV depending on CO concentration and operating temperature, which can halve the overall fuel cell voltage under load. Sulfur-containing impurities, such as bisulfide (HS⁻) species generated in anion exchange membrane fuel cells (AEMFCs), further exacerbate poisoning by adsorbing onto Pt or alloy surfaces, leading to a reduction in power density by up to 30% through site blockage and altered electronic structure. Sulfur acts as an irreversible poison in many cases, forming stable sulfides that resist removal under standard operating conditions. To enhance CO tolerance, Pt-Ru alloys have been widely adopted, leveraging a bifunctional mechanism where Ru facilitates water activation to form hydroxyl groups (OH) at lower potentials (~0.2-0.3 V vs. RHE) compared to pure Pt. This shifts the CO oxidation potential negatively, enabling the reaction: \text{Pt-C(O)ads} + \text{Ru-(OH)ads} \rightarrow \text{Pt} + \text{Ru} + \text{CO}_2 + \text{H}^+ + \text{e}^- allowing sustained HOR activity even with 50-100 ppm CO present. Recent advancements in 2025 have explored high-entropy alloys (HEAs), such as Pt-based multi-element compositions, which improve tolerance to CO and sulfur poisoning in low-temperature PEMFCs.

Selective Poisoning

Principles and Mechanisms

Selective poisoning in catalysis refers to the deliberate introduction of mild poisons to block unselective active sites on the catalyst surface, thereby promoting the desired reaction pathways while suppressing side reactions. This approach modifies the catalyst's surface geometry and electronic properties in a controlled manner, enhancing overall selectivity without completely deactivating the material. The primary mechanism involves site-specific adsorption of the poison, where it preferentially binds to high-coordination or edge sites that facilitate unselective processes. For instance, in the selective hydrogenation of alkynes to alkenes using palladium catalysts, lead (Pb) acts as a poison that blocks edge sites on Pd, preventing over-hydrogenation to alkanes and favoring partial reduction to the cis-alkene product, as seen in Lindlar-type catalysts. This adsorption is typically reversible under mild conditions, allowing the poison to isolate reactive ensembles without permanent damage. Theoretically, selective poisoning operates through size , where the dilutes or fragments larger clusters required for non-selective , reducing them from typically 10-20 atoms to smaller ensembles of 3-5 atoms that the pathway. This , rooted in the model of , exploits the of , ensuring that only sites with the appropriate participate effectively. The foundational of selective traces back to the , with contributions from Michel Boudart, who demonstrated its use in probing and enhancing selectivity in ammonia oxidation over catalysts by selectively deactivating structure-sensitive sites. Boudart's work established as a for understanding site requirements, influencing strategies.

Applications and Advantages

Selective poisoning finds prominent application in the industrial epoxidation of to over silver catalysts supported on α-alumina. Here, cesium acts as a promoter that neutralizes sites on the , suppressing the of to and other byproducts. This selective deactivation of unselective sites enhances the overall selectivity to by up to 10 points at comparable conversions, enabling operation at lower temperatures while maintaining activity. Another application is in the selective hydrogenation of dienes, such as 1,3-butadiene, to mono-olefins using palladium-based catalysts. Sulfur poisoning modifies the Pd surface by blocking highly active sites responsible for over-hydrogenation, thereby promoting the formation of desired partial hydrogenation products like butenes over fully saturated alkanes. This approach is particularly valuable in petrochemical refining and synthesis, where controlled selectivity prevents unwanted side reactions. The primary advantages of selective poisoning lie in its ability to boost reaction yields and minimize byproduct formation, leading to more efficient processes. In ethylene epoxidation, cesium promotion can elevate ethylene oxide selectivity from around 80% to over 90% under industrial conditions, reducing waste and downstream purification needs. Similarly, in hydrogenation reactions, poisoning strategies like those in sulfur-modified Pd systems improve mono-olefin selectivity by tuning adsorption strengths, achieving high yields for target alkenes compared to unpoisoned catalysts. In pharmaceutical synthesis, selective poisoning exemplifies these benefits through catalysts like Lindlar's Pd, poisoned with lead and quinoline, which halts hydrogenation at the cis-alkene stage and avoids over-reduction to alkanes. This delivers high stereoselectivity (>95% cis) in key steps for producing compounds such as vitamin A derivatives and prostaglandins, streamlining multi-step syntheses and cutting costs associated with separation and recycling. Overall, such techniques lower purification expenses in targeted reactions by curtailing side products, while also enabling milder operating conditions that conserve energy. Economically, selective poisoning has driven industrial advancements, notably in the 1980s adoption of Pd-based catalysts for acetylene hydrogenation to ethylene in ethylene purification streams. By replacing energy-intensive extractive distillation with selective catalytic processes, these innovations reduced overall energy consumption in olefin production, enhancing profitability in large-scale petrochemical operations. Recent explorations, such as in 2023 studies on CO2 electroreduction, have shown that trace halide additives acting as mild poisons on Cu electrodes can enhance C2 product (e.g., ethylene) selectivity over unpoisoned surfaces, pointing to emerging sustainability benefits in carbon utilization.

Prevention and Mitigation

Feedstock Purification Techniques

Feedstock purification techniques aim to eliminate or reduce catalyst poisons from input materials upstream of the catalytic process, thereby extending catalyst life and maintaining reaction efficiency. Common methods include adsorption, distillation, and hydrotreating, each targeting specific classes of impurities such as organic compounds, volatile sulfur species, and metals. These approaches are essential in industries like petroleum refining and syngas production, where contaminants like sulfur, nitrogen, and heavy metals can adsorb onto active sites and deactivate catalysts. Adsorption employs materials like activated carbon to capture organic poisons, which can form coke precursors that block catalytic sites. Activated carbon's high surface area and pore structure facilitate physical and chemical interactions with aromatics and sulfur-containing organics, such as dibenzothiophene, enabling effective removal from hydrocarbon feedstocks. For instance, oxidized activated carbons enhance adsorption through increased surface heterogeneity, improving desulfurization in liquid fuels. Distillation separates volatile poisons like hydrogen sulfide (H₂S) based on boiling point differences, often integrated into refinery streams to isolate sulfur compounds from hydrocarbons before hydroprocessing. This method achieves high recovery rates, such as over 99% CO₂ with H₂S reduced to below 25 ppmv in acid gas streams. Hydrotreating, particularly pre-hydrodesulfurization (pre-HDS), uses hydrogen and catalysts to convert metallic impurities and sulfur into removable forms, protecting downstream catalysts from permanent deactivation. Specific processes, such as guard beds filled with zinc oxide (ZnO), are widely used for deep sulfur removal in hydrogen and syngas streams. These beds operate at 200–400°C, where ZnO reacts with H₂S to form stable ZnS, reducing sulfur levels to below 0.1 ppm (or even 10–50 ppbv) to meet stringent requirements for sensitive catalysts in reforming or synthesis applications. In refinery feeds, hydrotreating achieves approximately 99% removal efficiency for sulfur and metals, converting high-sulfur feeds (10,000–20,000 ppm) to ultra-low levels (10–15 ppm), thus preventing poisoning in subsequent units. Cost considerations balance initial investments against long-term savings from reduced catalyst replacement and downtime. For example, amine scrubbing systems for removing CO₂ and H₂S from syngas production incur capital costs of $30–50 million for modular biomass-to-liquids plants, with operational expenses offset by avoiding catalyst poisoning that could increase maintenance by 20–30%. These units, using solvents like MDEA, target H₂S below 100 ppmv while enabling 96–99% overall sulfur recovery in integrated processes. Recent advances include membrane-based separations for trace metal removal in biofuel feedstocks, enhancing purity for sustainable processing. Nanofiltration membranes, optimized in 2024 studies, achieve high rejection rates (often >95%) for heavy metals like nickel and lead from biomass-derived streams, reducing energy use compared to traditional methods and supporting cleaner biofuel production.

Catalyst Design Innovations

Modern catalyst design innovations focus on engineering inherent resistance to poisons through advanced material architectures, enabling sustained performance in harsh environments such as fuel cells and hydrogenation processes. Alloying represents a foundational strategy, where bimetallic compositions modify electronic structures to weaken poison adsorption. For instance, Pt-Ru alloys enhance CO tolerance in proton exchange membrane fuel cells (PEMFCs) by facilitating bifunctional mechanisms, where Ru sites generate OH species to oxidize adsorbed CO on Pt, downshifting the Pt d-band center and reducing CO binding strength. This approach has led to improvements in CO tolerance, with Pt-Cu alloys exhibiting enhanced current densities during methanol oxidation reactions compared to pure Pt catalysts. Similarly, Pt-Sn bimetallics demonstrate superior sulfur resistance in dehydrogenation and fuel cell applications, where Sn modifies Pt active sites to inhibit S adsorption, maintaining activity under sulfur-containing feeds. Core-shell structures further protect active sites from poisons by encapsulating vulnerable cores within selective shells that permit reactant diffusion while blocking contaminants. In Pd@Au core-shell nanoclusters, the Au shell induces charge transfer from Pd core, resulting in a negative d-band center that yields weak SO2 adsorption energies (-0.67 eV), comparable to O2 and far weaker than CO, thereby suppressing sulfur poisoning during CO oxidation. Single-atom catalysts (SACs) advance this paradigm by dispersing isolated metal atoms on supports, often stabilized with ligands that lower poison affinity through tailored coordination environments. Oxygen-ligand-steered SACs, such as Ni-O-C configurations, achieve high thermodynamic stability with binding energies surpassing metal cohesive energies, enhancing resistance to aggregation and poisons via strong immobilization and altered electronic properties. These designs reduce poisoning effects by minimizing active site exposure and affinity for contaminants like CO or sulfur. Promoters and supports play crucial roles in bolstering poison tolerance; for example, Sn addition to Pt catalysts not only alloys the metal but also acts as a promoter to enhance sulfur resistance in reforming processes. Oxide supports like CeO2 provide oxygen storage capacity, facilitating in-situ regeneration by storing and releasing oxygen to oxidize poisons, as seen in CeO2-supported Pt catalysts that exhibit high sulfur tolerance in waste-to-hydrogen conversion due to improved redox properties. Bimetallic catalysts generally offer higher poison tolerance over monometallics through synergistic electronic and geometric effects. Emerging high-entropy alloys (HEAs) in fuel cells exemplify this, with compositions like IrPdPtRhRu achieving overpotentials ~40 mV lower than Pt(111) for oxygen reduction reaction (ORR) while enhancing durability against impurities, enabling operation for thousands of hours in practical conditions. Recent advancements leverage AI-optimized designs, integrating density functional theory (DFT) simulations with machine learning to predict poison binding energies and screen compositions from 2023-2025 studies. These frameworks have accelerated HEA development for electrocatalysis, identifying optimal multi-element mixes that minimize poison adsorption while maximizing activity, as in PdNiRuIrRh alloys showing 8-fold higher hydrogen oxidation mass activity than Pt/C with inherent stability. Such computational tools prioritize seminal configurations, reducing experimental iterations and enabling poison-resistant catalysts for sustainable energy applications.

Regeneration Methods

Reversible Poison Removal

Reversible poison removal involves non-destructive techniques to desorb weakly bound contaminants from catalyst surfaces, restoring activity without altering the structure. These methods are applicable to poisons like (CO) or sulfur species that exhibit reversible adsorption, allowing the to regain 80-90% of its original performance through targeted desorption processes. Key methods include thermal desorption, where the catalyst is heated to 200-400°C under an inert gas such as nitrogen to volatilize adsorbed poisons. This approach exploits the lower binding energy of reversible poisons, enabling their release without permanent site blockage. Flushing with a clean feed, such as pure hydrogen or inert gas, displaces poisons by competitive adsorption or dilution, often combined with mild heating for enhanced efficiency. In fuel cell applications, electrochemical stripping uses cyclic voltammetry—sweeping the potential between 0.1 and 1.4 V—to oxidize and desorb poisons like sulfides or CO from platinum surfaces. A representative example is the removal of CO from platinum (Pt) catalysts via air oxidation, where exposure to oxygen at around 200°C oxidizes adsorbed CO to CO₂, achieving full activity recovery within 30 minutes. This mild oxidation followed by hydrogen reduction fully restores the catalyst's hydrogenation performance, as demonstrated in studies on carbon-supported Pt systems. These techniques are often integrated in-situ during cyclic operations, such as in fluid catalytic cracking (FCC) units, where reversible poisons like basic nitrogen compounds are mitigated by periodic feed switching to clean streams or controlled oxidation in the regenerator, minimizing downtime. Despite their effectiveness, these methods incur energy costs from heating or electrochemical cycling, and incomplete desorption can occur with higher poison loadings. Recent optimizations, such as microwave-assisted heating, have improved desorption efficiency compared to conventional thermal methods through targeted volumetric heating.

Irreversible Poison Handling

Irreversible in catalysts, often caused by such as and or persistent sulfides, leads to permanent deactivation that cannot be fully reversed through standard regeneration, necessitating specialized handling to recover valuable components or safely dispose of the . Chemical represents a primary method for addressing irreversible metal , where acids dissolve and extract deposited metals from the catalyst , partial of active components like . In this , spent catalysts are typically pretreated through to convert poisons into leachable forms, followed by acid to selectively solubilize metals while minimizing . For instance, treatment of spent hydrodesulfurization (HDS) catalysts has achieved up to 90% efficiency for , allowing reuse in new catalyst formulations and reducing the need for virgin materials. Oxidative combustion serves as an effective approach for removing sulfide-based irreversible poisons, involving controlled high-temperature roasting in oxygen-rich environments to convert sulfides into volatile sulfur oxides, thereby liberating active sites or preparing the catalyst for further metal extraction. This thermal treatment not only eliminates sulfur deposits but also aids in carbon and organic residue removal, though it requires careful control to avoid sintering of the catalyst support. An application of this method includes the pretreatment of petroleum refinery catalysts, where oxidative roasting facilitates subsequent leaching steps with efficiencies exceeding 80% for sulfide elimination. For cases where recovery is uneconomical or incomplete, replacement scheduling emerges as a practical strategy, involving systematic monitoring of catalyst performance metrics—such as activity decline rates and poison accumulation—to determine optimal shutdown and substitution intervals, thereby minimizing operational disruptions. This approach integrates economic modeling to balance downtime costs against extended use, particularly in continuous processes like reforming, where irreversible poisoning accelerates deactivation. Incineration provides a targeted for handling halogen-induced irreversible , as seen in spent catalysts contaminated with or , where high-temperature volatilizes as hydrogen halides, enabling safe disposal or preprocessing for metal reclamation. This technique is particularly useful for adsorbable organic (AOX), achieving near-complete removal rates while complying with emission standards. Handling irreversible poisons faces significant challenges, including stringent environmental regulations that govern waste classification and recycling mandates. The EU Waste Framework Directive 2008/98/EC, for example, imposes requirements for waste hierarchy prioritization—favoring recycling over disposal—and tracks hazardous catalyst streams to prevent leaching of toxins into ecosystems. Additionally, technical recycling rates for noble metals in spent catalysts exceed 95%, though overall collection and processing efficiencies vary by operation. Recent advancements as of 2025 have focused on enhanced hydrometallurgical processes for recovering vanadium and nickel from petroleum-derived spent catalysts, incorporating bioleaching or optimized acid cocktails to boost extraction yields and reduce overall waste generation by up to 40% compared to traditional pyrometallurgical routes. These innovations emphasize closed-loop systems that minimize acid consumption and effluent volumes, supporting circular economy principles in catalyst management. Innovative methods like supercritical fluid extraction have also emerged for poison removal, offering solvent-based regeneration with reduced environmental impact.

References

  1. [1]
    Unlocking catalytic longevity: a critical review of catalyst deactivation ...
    Aug 14, 2025 · The deactivation of a catalyst by a chemical substance, whether partial or complete, is known as catalyst poisoning. Poisoning is a particular ...
  2. [2]
    Catalyst Poisoning - an overview | ScienceDirect Topics
    Catalyst poisoning is defined as the permanent performance loss of a catalyst due to the adsorption of impurities, such as methanol and reaction intermediates ...
  3. [3]
    17.6: Catalysts and Catalysis - Chemistry LibreTexts
    Nov 13, 2022 · Catalyst poisoning, brought about by irreversible binding of a substance to its surface, can be permanent or temporary. In the latter case the ...
  4. [4]
    None
    Summary of each segment:
  5. [5]
    Catalyst Poisoning - an overview | ScienceDirect Topics
    Catalyst poisoning occurs when contaminant molecules in the gas stream adsorb irreversibly onto active sites of the catalyst. Sulfur is a common poison for ...
  6. [6]
    Evaluate Catalyst Replacement Economics: Part 1 — Periodic Replacement
    ### Summary of Catalyst Replacement Economics in Refining
  7. [7]
    [PDF] Catalyst Deactivation Mitigation for Biomass Conversion
    Apr 7, 2023 · Impact - Review paper on catalyst deactivation in biomass conversion ... • Unlocking the Mystery of Catalyst Poisoning, https://www.energy ...
  8. [8]
    Poisoning of automotive methane oxidation catalysts by silicon ...
    Dec 5, 2023 · Siloxanes and silicon dioxide are severe catalyst poisons in biomethane. Siloxanes selectively poisoned the exhaust catalyst anchoring on active metal sites.
  9. [9]
    [PDF] Three-way catalyst performance using natural gas with two different ...
    Stoichiometric engines running on natural gas rely on three-way catalysts to meet limits e.g. Euro 6 regarding emissions of hydrocarbons (including methane) ...
  10. [10]
    Predicting catalyst lifetime - DigitalRefining
    Therefore, the catalyst of a semi-regenerative catalytic reformer is regenerated during routine shutdowns of the process every 18-24 months. Normally, the ...
  11. [11]
    The impact of pollutants on catalyst performance during hydrogen ...
    Pollutants like SO2, NOx, and organic compounds can negatively impact catalyst performance by forming sulfate, corroding electrodes, and blocking active sites.
  12. [12]
    Molecular Catalysis in “Green” Hydrogen Production - Frontiers
    In this contribution we review the journey of the development of molecular catalysts for both HER and the OER, highlighting selected systems.Introduction · Hydrogen Evolving Reaction... · Oxygen Evolution Reaction...
  13. [13]
  14. [14]
    Mechanisms of Catalyst Poisoning by Sulfur Species - ScienceDirect
    In terms of chemical changes the adsorption of gaseous impurities from the feed is the most common source of catalyst poisoning. The adspecies of major concern ...Missing: transformation sulfide
  15. [15]
    [PDF] mechanism of sulfur poisoning by h2s and so2 of - HARVEST (uSask)
    Apr 4, 2017 · Finally, the mechanism of dry reforming of methane and the poisoning of catalysts by hydrogen sulfide and sulfur dioxide during the reaction ...
  16. [16]
    Tracking Sulfur Poisoning of Pd/Al2O3 Catalysts for Methane ...
    Jul 21, 2025 · Understanding the deactivation induced by sulfur poisoning is crucial for designing more efficient palladium-based monolithic catalysts for ...
  17. [17]
    Nanoscale Analysis of Sulfur Poisoning Effects on Hydrogen ...
    Oct 15, 2025 · Ensemble-based plasmonic measurements showed that sulfur poisoning lowered the kinetic rates for hydrogen (de)sorption on poisoned NPs ...
  18. [18]
    Pd ensemble effects on oxygen hydrogenation in AuPd alloys
    May 16, 2011 · Pt-based catalysts are also susceptible to poisoning by carbon monoxide (CO) and other impurities [2]. For these reasons, there exists great ...
  19. [19]
    Nanoparticles as an antidote for poisoned gold single-atom catalysts ...
    Apr 16, 2024 · Single-atom catalysts (SACs) capture significant interest in catalysis by virtue of their maximum atom utilization, tunable electronic ...<|control11|><|separator|>
  20. [20]
    Recent advances in the development of single atom catalysts for ...
    Sep 11, 2024 · This review comprehensively covers the fundamentals, synthesis methods, and recent advances in applications of SACs for the oxygen evolution reaction (OER)
  21. [21]
    Desorption lifetimes and activation energies influencing gas–surface ...
    Mar 20, 2024 · We note that the energy range of 50 kJ mol−1 mentioned above to distinguish between physisorption and chemisorption is not necessarily ...Missing: poisons | Show results with:poisons
  22. [22]
    Poisoning (catalyst poisoning) - candcs
    Reversible poisoning, for example by water (H2O) in ammonia synthesis can be eliminated simply by removing the poison from the reaction mixture. In contrast, ...Missing: characteristics | Show results with:characteristics
  23. [23]
    Heat of Water Adsorption on Large-Pores Mesoporous Silica
    Aug 11, 2025 · The heat of water adsorption resulted to be 37–36 kJ/mol, in satisfactory agreement with the value obtained by microcalorimetry. The IR ...
  24. [24]
    Understanding Catalyst Poisoning in Precious Metal Catalysts
    Jul 24, 2025 · Catalyst poisoning refers to the loss or significant reduction of catalytic activity of a catalyst during a chemical reaction due to the ...
  25. [25]
    Catalyst deactivation Common causes - AmmoniaKnowHow
    Catalyst poisoning is a chemical effect and involves strong interaction ... Mechanisms by which a poison may affect catalytic activity are multifold as ...
  26. [26]
    Valorization of steel-work off-gases: Influence of impurities on the ...
    Sep 15, 2022 · 4, Fig. 5. The activity and methanol selectivity drop by about 10% after 100 h time-on-stream, suggesting that the S-poisoned catalyst exhibits ...
  27. [27]
    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.Missing: poisoning | Show results with:poisoning
  28. [28]
    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.
  29. [29]
    Catalyst Deactivation - an overview | ScienceDirect Topics
    Examples include sulfur dioxide and halogens, although water and ammonia can irreversibly damage strong acid catalysts. If the presence of irreversible poisons ...
  30. [30]
    Lead Accumulation on Automotive Catalysts - jstor
    of lead oxide poisoning of the platinum, both of which are very sensitive to ... "Scavenger and Lead Poisoning of Automotive. Oxidation Catalysts." SÁE ...Missing: toxicity threshold
  31. [31]
    Impact of the support material on the phosphorus poisoning on Co ...
    Nov 1, 2024 · The P loading resulted in a decrease in site-time yield and a higher hydrogenation activity leading to a reduced production of value products.
  32. [32]
    January 2024Catalysts—Watkins (ART Hydroprocessing)
    Phosphorus is found in the phospholipids present in biologically produced oils ... poisoning include rapid activity loss and an increase in pressure drop.
  33. [33]
    Mechanisms of Catalyst Poisoning in Palladium-Catalyzed ...
    Reaction paths leading to palladium catalyst deactivation during cyanation of haloarenes (eq 1) have been identified and studied.
  34. [34]
    [PDF] 1 MODELING SULFUR POISONING OF PALLADIUM MEMBRANES ...
    Catalytic poisoning by adsorbed sulfur compounds and corrosive decay produced by surface sulfidation have been reported as the two main mechanisms that reduce ...Missing: strain | Show results with:strain
  35. [35]
    Hydrogenation of acetylene over supported metal catalysts. Part 3 ...
    Addition of 0.1–1.0 Torr of carbon monoxide to the acetylene hydrogenation results in complete poisoning of the rhodium and iridium catalysts, but only partial ...Missing: 1970s Pd/
  36. [36]
    Single-Atom Alloy Pd–Ag Catalyst for Selective Hydrogenation of ...
    Here we present results for the selective hydrogenation of acrolein over a Pd–Ag bimetallic system with dilute Pd in Ag as a host metal. The situation is ...
  37. [37]
    CO and C 3 H 6 poisoning of hydrogen permeation across Pd 77 Ag ...
    Feb 15, 2022 · In this work, we compare the poisoning (inhibiting) effects of CO and C 3 H 6 on H 2 permeation across Pd 77 Ag 23 and Pd foil membranes under identical ...
  38. [38]
    Large reduction in distillate fuel sulfur content has only minor effect ...
    Feb 24, 2015 · In 2006, the limit for highway diesel was further tightened, as the EPA required no more than 15 ppm of sulfur (ultra-low-sulfur diesel) by 2010 ...Missing: regulations | Show results with:regulations
  39. [39]
    Revisions to Motor Vehicle Diesel Fuel Sulfur Transition Provisions
    Nov 22, 2005 · Amended to require recordkeeping by each facility that receives or distributes 15 ppm or 500 ppm highway diesel fuel beginning June 1, 2006.
  40. [40]
    Design and Evaluation of a Co–Mo-Supported Nano Alumina ...
    May 18, 2020 · A nanocatalyst, γ-Al 2 O 3 nanoparticle-supported CoMo, was prepared experimentally and evaluated through a hydrodesulfurization (HDS) process for removing ...<|separator|>
  41. [41]
    Poisoning effect of nitrogen compounds on the performance of ...
    The nitrogen compounds poison acidic sites on the catalyst, as confirmed by the reduced amounts of pyridine adsorbed and by the suppressed isomerization of 2,2′ ...
  42. [42]
    Inhibition effects of nitrogen compounds on the hydrodesulfurization ...
    The inhibiting effect of the non-basic compounds, indole and carbazole, was found to be comparable to that of the basic compound, quinoline. The inhibition ...
  43. [43]
    Poisoning of Hydrodesulfurization Catalysts by Vanadium
    The HDS activity of a commercial NiMo catalyst has been followed versus the amount of vanadium added to it, either by an oxidic procedure or by a sulfided route ...Missing: MoS2 | Show results with:MoS2
  44. [44]
    Catalyst poisoning and fixed bed reactor dynamics part III: Start-up ...
    Previous studies of the dynamic behavior of adiabatic fixed bed reactors under conditions of catalyst poisoning have been extended to investigation of the ...Missing: accumulation HDS
  45. [45]
    Deactivation of a Ni/Mo hydrotreating catalyst by vanadium deposits
    After resulfiding, the vanadium-containing catalysts were tested for catalytic activity for hydrodesulfurization (HDS), hydrogenation (HYD) and ...
  46. [46]
    Nickel and vanadium removal | Catalyst poison removal | Ni - Topsoe
    A final-stage catalyst for fixed-bed resid hydrodesulfurization (HDS) units. It delivers a high capacity for uptake of nickel and vanadium, together with a ...
  47. [47]
    Progress and Perspectives in the Catalytic Hydrotreatment of Bio-Oils
    This review focuses on the catalytic hydrotreatment of bio-oils, which has been shown to be an effective approach for the removal of heteroatoms at moderate ...
  48. [48]
    Hydrodeoxygenation of Phenolic Compounds and Lignin Bio-Oil ...
    The hydrodeoxygenation of a surrogate mixture of eleven phenolic and aromatic compounds was then studied by investigating the effect of reaction conditions ( ...Hydrodeoxygenation Of... · 2. Results · 2.4. Carbon/coke...
  49. [49]
    Design strategies of Pt-based electrocatalysts and tolerance ...
    Feb 7, 2023 · Two main strategies have received increasing attention in improving the CO tolerance of Pt-based electrocatalysts, including alloying Pt with a second element ...
  50. [50]
    Mechanistic insight into the CO oxidation reaction at pure, Nb-doped ...
    At other alternative binding sites (conformers c-f) the CO binding energy is even lower (around 1.5 eV or less). ... Pt18 binding energy (of the order of 1.5 eV).
  51. [51]
    Advanced Strategies for Mitigating Catalyst Poisoning in Low and ...
    Catalyst poisoning remains a persistent barrier to the efficiency and longevity of electrocatalytic energy conversion devices, namely fuel cells.
  52. [52]
    Strategies for Carbon and Sulfur Tolerant Solid Oxide Fuel Cell ...
    In this review, we first study the theoretical basis behind carbon and sulfur poisoning, before examining the strategies toward carbon and sulfur tolerance ...
  53. [53]
    Structurally Ordered High‐Entropy Intermetallics for Electrocatalysis
    Jul 29, 2025 · High-entropy intermetallic (HEI) catalysts have emerged as promising candidates for next-generation catalytic materials.
  54. [54]
    Statistical models for ensemble control by alloying and poisoning of ...
    In the present work a poison lattice model, taking into account the form of the ensembles, the ordering of the passivating atoms, and their radius of influence, ...
  55. [55]
    The role of lead in the selectivity of palladium-lead (lindlar type ...
    Surface structures and selectivities of lead-poisoned Pd or Pd/graphite catalysts were studied. Two types of surface species, adsorbed lead and bulk lead, ...
  56. [56]
    Role of Ensembles in Catalysis by Metals - SpringerLink
    Feb 22, 2016 · Among the recent ideas in catalysis by metals, the concept of ensemble appears to be one of the most promising. According to this view the ...
  57. [57]
    List of Publications and Patents of Michel Boudart
    Boudart, Ligand and Ensemble Effects in the Adsorption of Carbon ... Boudart, Model Catalysts: Reductionism for Understanding, Top. Catal., 13 (1 ...
  58. [58]
    Ethylene epoxidation over silver and copper–silver bimetallic catalysts
    It has been suggested that the presence of cesium neutralizes acid sites on the support ... promoter modifying the active sites of a working catalyst. For ...
  59. [59]
  60. [60]
    Applications of catalysis in pharmacuetical industry - OAText
    The selectivity can be improved by the use of suitable catalyst poisons which modify the activity of the metal catalyst. One of the most widely used and ...Over-the-counter and... · Improved methods for catalyst... · Catalyst recovery
  61. [61]
    Palladium–bismuth intermetallic and surface-poisoned catalysts for ...
    An increase in the Pd:Bi ratio from 7 to 1 in the Bi-poisoned catalysts decreased the hydrogenation activity due to blocking of active sites, but increased ...
  62. [62]
    Catalytic Reactions of Acetylene: A Feedstock for the Chemical ...
    The hydration of acetylene was also attempted on zeolites exchanged with ions such as Ag+, Zn2+, Cd2+, and others at temperatures between 100 and 200 °C. These ...
  63. [63]
    Selective electroreduction of CO 2 to C 2+ products on cobalt ...
    Aug 6, 2024 · Our results revealed a volcano-shaped correlation between Co loading and C2+ selectivity, with the CuCo0.4% catalyst exhibiting a 2-fold ...
  64. [64]
    How to Prevent Catalyst Poisoning at the Industrial Scale
    Apr 23, 2024 · This phenomenon occurs when chemical impurities interact with catalyst surfaces, diminishing their activity and longevity.
  65. [65]
    Importance of Structural and Chemical Heterogeneity of Activated ...
    The results showed the importance of activated carbon pore sizes and surface chemistry for the adsorption of dibenzothiophene (DBT) from liquid phase.
  66. [66]
    [PDF] H2S REMOVAL FROM CO2 BY DISTILLATION - Trimeric Corporation
    Oct 8, 2018 · This paper introduces a new process configuration developed by Air Products and Chemicals,. Inc. to efficiently separate carbon dioxide ...
  67. [67]
    None
    ### Summary of Feedstock Purification Techniques (Johnson Matthey Brochure)
  68. [68]
    Desulfurization of Biomass Syngas Using ZnO-Based Adsorbents
    Feb 24, 2020 · The guard bed should be operated at low temperatures of 200–300 °C to ensure deep sulfur removal according to the equilibrium calculations.
  69. [69]
    An Overview of Hydrotreating - AIChE
    Hydrotreating processes are standard in refineries primarily to remove sulfur from refined petroleum fuels. This helps reduce sulfur dioxide emissions that are ...Missing: pre- HDS
  70. [70]
    [PDF] Equipment Design and Cost Estimation for Small Modular Biomass ...
    The amine unit is designed to selectively remove H2S to a concentration of 100 ppmv; after an analysis of possible amine solvents, it was determined that. MDEA ...
  71. [71]
    Nanofiltration membranes for sustainable removal of heavy metal ...
    Jun 14, 2024 · This review is expected to shine a light on the fabrication of NF membranes for high-performance removal of heavy metal ions from the wastewater.Missing: biofuel | Show results with:biofuel
  72. [72]
    Design of PtSn Nanocatalysts for Fuel Cell Applications - 2024
    Oct 9, 2024 · Nanostructured PtSn is a versatile material for both anode and cathode electrodes in fuel cells showing higher resistance to CO poisoning and improved kinetics.<|separator|>
  73. [73]
    Core–shell PdAu nanocluster catalysts to suppress sulfur poisoning
    These findings provide insights into designing core–shell catalysts to suppress sulfur poisoning while optimizing catalytic behaviors.
  74. [74]
  75. [75]
    The stability of single-atom catalysts in thermocatalysis
    Jun 19, 2024 · The promising way to avoid SAC poisoning is to develop poisoning-resistant SACs by lowering the affinity between active metal atoms and poisoned ...
  76. [76]
    Sulfur-Resistant CeO2-Supported Pt Catalyst for Waste-to-Hydrogen
    The Pt-based catalyst prepared by incipient wetness impregnation showed the highest catalytic activity and sulfur tolerance due to the standout factors.
  77. [77]
    High entropy powering green energy: hydrogen, batteries ... - Nature
    May 22, 2025 · This review examines recent innovations in HEMs, focusing on hydrogen generation/storage, fuel cells, batteries, semiconductors/electronics, and catalysis.
  78. [78]
    High entropy alloys for advanced electrocatalysis with computational ...
    Sep 15, 2025 · The growing demand for sustainable energy solutions drives the exploration of high-entropy alloys (HEAs) in electrocatalysis.<|separator|>
  79. [79]
    Review on mechanisms and recovery procedures for reversible ...
    Mar 15, 2021 · Several procedures were established to recover performance losses caused by sulfide poisoning. The most common procedure is the use of CV.
  80. [80]
  81. [81]
    None
    ### Summary of CO Removal from Pt Catalysts Using Air Oxidation
  82. [82]
    Comparative Study of Microwave and Resistance Heating for the ...
    Sep 6, 2024 · The findings suggest that microwave-assisted thermal desorption is a highly effective and energy-efficient technology for soil remediation, ...
  83. [83]
    Efficient recovery of all valuable metals from spent HDS catalysts
    Spent HDS catalysts are pretreated by roasting to remove carbon, sulfur, and organics while converting metals to oxides or salts that can be easily leached.
  84. [84]
    Recovery and Resource Utilization of Spent Hydrogenation Catalysts
    The present study focuses on the efficient utilization of spent hydrogenation catalysts. To this end, it summarizes pretreatment processes, leaching methods, ...
  85. [85]
    Extraction and Recovery of Metals from Spent HDS Catalysts - MDPI
    These catalysts contain valuable metals like cobalt (Co), molybdenum (Mo), nickel (Ni), and vanadium (V). In particular, one Co-Mo catalyst was treated in order ...
  86. [86]
    Industrial-Scale Technology for Molybdic Acid Production from ... - NIH
    Aug 23, 2023 · The article describes the technology of molybdic acid recovery from spent petrochemical catalysts (HDS) developed and implemented in industrial activity.
  87. [87]
    What criteria do you use to determine end of run (EOR) catalyst life ...
    Refinery economics are typically the ultimate factor used for determining when to replace CCR Reforming catalysts. The useful life of a CCR catalyst is impacted ...
  88. [88]
    Optimal control of scheduling and production for a multienzymatic ...
    Optimal catalyst replacement decisions must account for dynamic demand, shutdown costs, and storage expenses, complicating optimization due to mixed-integer ...
  89. [89]
    Extensive comparison of methods for removal of organic halogen ...
    Nov 1, 2023 · The present study examines the performance of air stripping, steam stripping, distillation, and incineration in adsorbable organic halogens (AOX) ...
  90. [90]
    Waste Framework Directive - Environment - European Commission
    Discover the EU's Waste Framework Directive which sets the basic concepts and definitions related to waste management, such as recovery and recycling.Missing: catalyst | Show results with:catalyst
  91. [91]
    [PDF] European Catalysts Manufacturers Association - ECMA
    Within the European Union, the national regulations, implementing the EU Waste Framework · Directive, will apply and must be observed. Page 9. GUIDELINES FOR ...
  92. [92]
    Recovery of Noble Metals from Spent Catalysts: A Review
    Recycling of catalytic waste, in particular, presents a more cost-effective and environmentally sustainable approach acquiring these metals, in contrast to ...
  93. [93]
    Life Cycle Assessment of Mo, V, Ni, and Co Recovery from Spent ...
    Aug 19, 2025 · The goal of this study was to evaluate the environmental impacts of recycling a spent catalyst and recovering all of the valuable metals: Mo as ...
  94. [94]
    Sustainable extraction of valuable metals from petroleum spent ...
    Feb 19, 2025 · The petroleum spent hydroprocessing catalysts contain high concentrations of toxic and valuable metals (Ni, V, Mo, and Al).
  95. [95]
    Biological recycling of critical metals from spent hydrodesulfurization ...
    Jun 3, 2025 · Recycling of spent catalysts supports the circular economy through smelting, acid or alkali leaching, or integrated pyro-hydrometallurgical ...<|separator|>