Fact-checked by Grok 2 weeks ago

Solid acid

A solid acid is a solid material that exhibits acidic properties through surface sites capable of donating protons (Brønsted acidity) or accepting electron pairs ( acidity), enabling it to function as a heterogeneous catalyst in chemical reactions without dissolving in the reaction medium. These materials are distinguished from traditional liquid acids by their insolubility, which facilitates easier separation from reaction products, reduces , and minimizes environmental hazards associated with handling and disposal. Solid acids encompass a diverse range of structures, including zeolites (crystalline aluminosilicates with microporous frameworks), sulfated metal oxides such as sulfated zirconia (ZrO₂/SO₄²⁻), heteropolyacids supported on high-surface-area carriers like silica, polymeric ion-exchange resins, and mesoporous silicas templated with micelles for tunable acidity. Sustainable carbon-based solid acids derived from or waste materials offer additional benefits. The acidity strength and density of these sites can be precisely engineered through synthetic methods, such as or sulfonation, to optimize catalytic performance for specific reactions. In industrial applications, solid acids play a pivotal role in petroleum refining processes like and isomerization, where they enable the conversion of heavy hydrocarbons into valuable fuels and chemicals. They are also increasingly vital in valorization for producing biofuels and platform chemicals from lignocellulosic feedstocks, as well as in synthesis for reactions such as esterification, , and , promoting greener manufacturing by enhancing selectivity and reducing waste. Ongoing research focuses on developing recyclable and durable solid acids to address challenges like deactivation from formation and to expand their use in sustainable .

Introduction

Definition

Solid acids are heterogeneous materials that exhibit acidity in the solid state, primarily through proton donation (Brønsted acidity) or electron pair acceptance ( acidity), enabling catalytic reactions at their surface without dissolving into the reaction medium. These materials, often synthesized by incorporating acidic functional groups onto a solid support, facilitate surface-mediated interactions with reactants in gas or liquid phases, distinguishing them from homogeneous acids where the catalyst and reactants share the same phase. In contrast to traditional liquid acids like sulfuric or , solid acids immobilize active sites on a stable matrix, such as silica, alumina, or zeolites, allowing for straightforward separation of the catalyst from products via and promoting reusability in . This heterogeneous nature supports high turnover numbers and selectivity while minimizing and handling hazards associated with liquid counterparts. Surface acidity in solids stems from specific sites, including Brønsted acid sites like bridging hydroxyl groups (Si-OH-Al) that donate protons, and Lewis acid sites such as coordinatively unsaturated metal cations that accept pairs. These sites enable acid-base interactions tailored to various substrates, particularly in involving weakly basic hydrocarbons or oxygen-containing renewables. Solid acids contribute to principles by serving as environmentally benign alternatives to corrosive liquid acids, reducing waste generation, and enabling more sustainable manufacturing of chemicals and fuels.

Historical Development

The origins of solid acid catalysis trace back to the early , when researchers sought efficient methods for refining amid growing demand for . In , Eugene Houdry developed the Houdry , a fixed-bed catalytic cracking technique that utilized silica-alumina catalysts to convert heavy hydrocarbons into lighter fractions, marking the first industrial application of solid acids in processing. This innovation, commercialized in 1936, significantly improved yields over thermal cracking and laid the foundation for heterogeneous by demonstrating the stability and reusability of solid materials like amorphous silica-alumina. A major advancement occurred in the 1960s with the discovery of shape-selective catalysis using zeolites, pioneered by Paul B. Weisz and colleagues at Research and Development Corporation. In 1960, Weisz and V.J. Frilette published seminal work on intracrystalline catalysis in zeolite salts, revealing how the microporous structure of synthetic zeolites enabled selective hydrocarbon conversions by restricting reactant and product diffusion based on molecular size. This breakthrough, later exemplified and commercialized using zeolites like in processes such as 's (methanol-to-gasoline) in the 1970s, revolutionized petrochemical synthesis and refining by enhancing selectivity and efficiency over non-selective silica-alumina catalysts. In 1979, sulfated zirconia was developed as a solid superacid, extending the concept of extreme acidity to heterogeneous systems. Developed by researchers including Koichi Arata, this material achieved Hammett acidity (H0) values below -12, rivaling liquid s, through sulfate modification of zirconia surfaces, which generated strong Brønsted and acid sites for and reactions. George A. Olah, recipient of the 1994 for carbocation chemistry in superacids, bridged liquid and solid superacid research by highlighting the potential of supported systems like heteropolyacids and metal oxides for practical , influencing the design of environmentally benign alternatives to corrosive liquids. In the 2000s and , solid acid research shifted toward sustainable applications, with heteropolyacids emerging as versatile catalysts for due to their tunable acidity and solubility in polar media. These Keggin-type structures, such as , facilitated selective transformations in synthesis and . Concurrently, metal-organic frameworks (MOFs) gained prominence as customizable solid acids, incorporating acidic sites like sulfonic groups for valorization, enabling efficient conversions of lignocellulosic feedstocks into platform chemicals. Post-2020 developments have intensified focus on conversion, with MOF-derived and heteropolyacid-based catalysts optimizing and under mild conditions to support circular economies. As of 2025, recent advances include recyclable solid acid catalysts for one-step from waste feedstocks, further enhancing in manufacturing.

Classification

Brønsted Solid Acids

Brønsted solid acids are heterogeneous catalysts that facilitate proton donation through surface hydroxyl groups, enabling acid-base interactions in catalytic processes. These sites, often denoted as bridging hydroxyls, transfer protons to adsorbed molecules, initiating reactions such as or cracking by forming carbocations. In zeolites, for instance, the prototypical Brønsted site is the Si-OH-Al group, where aluminum in the silica generates acidic protons due to charge compensation. This proton transfer mechanism can be represented by the general equation: \text{Surface-OH} + \text{Base} \rightleftharpoons \text{Surface-O}^- + \text{BaseH}^+ where the surface hydroxyl acts as the proton donor, forming a conjugate base on the solid while protonating the adsorbate. Common materials exemplifying Brønsted solid acids include zeolites such as H-ZSM-5, which features a MFI framework with tunable Si/Al ratios that modulate site density and strength. Amorphous aluminosilicates, like silica-alumina, also provide bridging Si-OH-Al sites, offering broader pore structures compared to crystalline zeolites while maintaining comparable Brønsted acidity. Heteropolyacids (HPAs), particularly Keggin-type structures like phosphotungstic acid (H₃PW₁₂O₄₀), exhibit strong Brønsted acidity through protons associated with oxygen atoms in the polyanion framework, often surpassing mineral acids in proton donation capability. These materials' acidity strength is adjustable by varying framework composition—for example, increasing aluminum content in aluminosilicates enhances proton affinity—or by anion substitution in HPAs, influencing deprotonation energies. Structural features play a crucial role in their function; in zeolites, micropores (typically <2 nm) confine reactants and transition states, promoting shape selectivity and stabilizing carbocations through electrostatic interactions. HPAs' pseudo-liquid behavior, arising from their hydrated Keggin clusters, allows for efficient proton mobility on the surface, while aluminosilicates' disordered networks provide accessible sites without diffusion limitations. These attributes enable Brønsted solid acids to outperform liquid acids in heterogeneous systems by combining high acidity with thermal stability.

Lewis Solid Acids

Lewis solid acids are heterogeneous catalysts characterized by their ability to act as electron pair acceptors, primarily through coordinatively unsaturated metal cations on the surface of solid materials. These sites, often denoted as :M^{n+}, facilitate reactions by binding to Lewis bases, thereby polarizing adjacent bonds and enhancing substrate reactivity. Unlike Brønsted acids, which rely on proton donation, Lewis solid acids promote activation via coordination chemistry, making them suitable for processes involving non-protic substrates. The fundamental mechanism involves the coordination of a Lewis base to the metal cation, which withdraws electron density and polarizes the base's bonds for subsequent transformations. For instance, in γ-alumina (γ-Al₂O₃), Al³⁺ sites coordinate with olefins, polarizing the C=C double bond to generate a carbocation-like intermediate that enables C-C bond formation in oligomerization reactions. This coordination is particularly effective on surfaces with low-coordinated metal centers, where the electron deficiency is pronounced. Common materials exhibiting dominant Lewis acidity include simple metal oxides such as ZnO, TiO₂, and γ-alumina, where exposed metal cations like Zn²⁺, Ti⁴⁺, and Al³⁺ serve as active sites. Sulfated metal oxides, exemplified by SO₄²⁻/ZrO₂, feature enhanced Lewis acidity from unsaturated Zr⁴⁺ centers, which interact strongly with electron donors. Layered clays like montmorillonite also provide Lewis acid sites through their octahedral Al³⁺ or Fe³⁺ ions in the structure. The acidity strength of these sites correlates with the metal's coordination number; tetra- or penta-coordinated cations exhibit higher acidity than hexa-coordinated ones due to greater electron withdrawal capability. Although some materials display hybrid behavior with coexisting Brønsted and Lewis sites, the focus in Lewis solid acids remains on the predominant electron acceptance by metal centers, as seen in dehydrated metal oxides where surface hydroxyl groups are minimized.

Properties and Characterization

Acidity Metrics

The acidity of solid acids is quantified through a variety of experimental techniques that measure site density, strength, and type, enabling comparison across materials for catalytic performance. Key methods include Hammett indicator titration, which determines acid site density by colorimetric detection of protonation using indicators like n-butylamine, providing values in mmol/g for total acidic sites on surfaces such as montmorillonites. Another fundamental technique is temperature-programmed desorption (TPD) of ammonia, which assesses total acidity by adsorbing NH₃ onto acid sites and measuring desorption temperatures to quantify site density and infer strength distribution, often yielding results like 0.5–2 mmol/g for zeolites. Advanced spectroscopic methods offer detailed site identification. Fourier-transform infrared (FTIR) spectroscopy with pyridine adsorption distinguishes Brønsted and Lewis sites via characteristic peaks: the band at approximately 1545 cm⁻¹ indicates pyridinium ions on Brønsted sites, while 1450 cm⁻¹ corresponds to coordinated pyridine on Lewis sites, allowing quantification through integrated absorbance with extinction coefficients around 1.4–2.2 cm μmol⁻¹. Nuclear magnetic resonance (NMR) spectroscopy, particularly ²⁷Al and ³¹P MAS NMR with probes like trimethylphosphine oxide, probes local environments of acid sites; for instance, ³¹P chemical shifts shift downfield (e.g., 40–60 ppm) with increasing Brønsted acidity, revealing site coordination and strength in aluminosilicates. Acidity scales for solids adapt liquid-phase concepts, notably the Hammett acidity function H_0, defined as H_0 = -\log \left( \frac{[BH^+]}{[B]} \right) + \mathrm{p}K_{\mathrm{BH^+}}, where B is a weak base indicator, to evaluate protonating ability on surfaces. This function classifies superacidity in solids when H_0 < -12, surpassing pure sulfuric acid (H_0 = -12), as demonstrated in sulfated metal-organic frameworks where indicators confirm proton transfer beyond conventional limits. Quantitative metrics emphasize acid site density, typically reported as mmol/g from titration or TPD, which correlates with catalytic turnover; for example, densities above 1 mmol/g often enhance activity in acid-catalyzed reactions. Microcalorimetry measures strength distribution by differential heat of adsorption (e.g., >150 kJ/mol for strong sites) of probes like , mapping heterogeneous acidity profiles in materials such as zeolites, where initial heats exceed 200 kJ/mol for superacidic sites. These metrics collectively provide a framework for optimizing acids, prioritizing strength and density for targeted applications.

Structural and Thermal Properties

Solid acids, such as zeolites and metal oxides, typically exhibit high specific surface areas ranging from 100 to 800 m²/g, as determined by the Brunauer-Emmett-Teller (BET) method using nitrogen adsorption isotherms, which facilitates the dispersion of active sites and enhances reactant accessibility. Pore size distributions in these materials often include micropores (less than 2 nm) in zeolites and mesopores (2–50 nm) in silica-based variants, analyzed via the Barrett-Joyner-Halenda (BJH) method from desorption branches of isotherms; for instance, hierarchical zeolites may show average pore diameters around 4 nm, contributing to improved mass transfer. These textural features are critical for performance in processes requiring efficient diffusion, though excessive microporosity can lead to diffusion limitations in larger molecules. The crystalline structure of solid acids is commonly characterized by X-ray diffraction () patterns, which identify phases and reveal framework topologies, such as the aluminosilicate cages in zeolites with tunable Si/Al ratios (e.g., 10–50 in HZSM-5). Defects, including aluminum substitutions, hydroxyl nests, and groups, play a key role in modulating activity by influencing site accessibility and stability, as observed in dealuminated or acid-treated frameworks where XRD broadening indicates lattice strain. Thermal stability is assessed through (TGA) and (DSC), which track weight loss and phase transitions; zeolites, for example, maintain structural integrity up to 800°C, with initial below 200°C followed by collapse at higher temperatures. Hydrothermal stability poses challenges, particularly in zeolites under steam conditions, where dealumination and pore blockage occur at Si/Al ratios below 100, leading to reduced crystallinity and activity, though higher ratios or metal doping mitigate these issues. Mechanical properties, including and , impact industrial applicability; smaller particle sizes (e.g., <10 μm) enhance intraparticle diffusion by shortening diffusion paths, thereby reducing mass transfer resistances, while sufficient ensures durability during handling and fluidization in reactors. In sulfated zirconia or carbon-based solid acids, mechanical robustness is further supported by the inherent rigidity of oxide frameworks or cross-linked carbon matrices.

Preparation Methods

Synthesis Techniques

Solid acids are typically synthesized through methods that control the incorporation of acidic sites into crystalline or amorphous frameworks, such as and . These techniques emphasize precise control over composition, structure, and porosity to generate or essential for catalytic activity. Key approaches include for aluminosilicates, for oxides, for protonation, and precipitation for mixed systems. Hydrothermal synthesis is a primary method for preparing zeolite-based solid acids, involving the crystallization of aluminosilicate gels under aqueous, alkaline conditions at elevated temperatures and pressures. The process begins with mixing silica and alumina sources, such as sodium silicate and aluminate, to form a hydrogel, which is then aged and heated in an autoclave at 90–150 °C and 1–15 bar for 24–96 hours to yield crystalline structures like . Structure-directing agents, such as tetrapropylammonium (TPA⁺) cations, are crucial for high-silica zeolites like , guiding the formation of the MFI framework with medium-sized pores (0.45–0.6 nm) during synthesis at 150–200 °C for 24–72 hours; these templates are later removed by calcination. This method allows tuning of the Si/Al ratio, with lower ratios (≤5) favoring structures like LTA or X zeolites and higher ratios (≥5) enabling or beta types, influencing the density of acid sites. Sol-gel methods enable the synthesis of metal oxide solid acids, particularly amorphous , by controlled hydrolysis and condensation of molecular precursors. The process starts with the hydrolysis of alkoxides, such as tetraethoxysilane (TEOS) for silica and aluminum isopropoxide for alumina, in the presence of solvents like ethanol and catalysts (e.g., acetic or oxalic acid) to form a sol that gels into a network. Subsequent drying and calcination at around 500 °C densify the gel into a porous oxide with tunable Si/Al ratios and enhanced , as demonstrated in acetic acid-assisted routes yielding materials with narrow mesoporous distributions. This technique produces high-surface-area solids suitable for acid catalysis, with the hydrolysis step controlling particle uniformity and the thermal treatment stabilizing the amorphous structure. Ion exchange is employed to convert as-synthesized zeolites, often in sodium form, into protonated (H-form) solid acids by replacing exchangeable cations with ammonium ions followed by deammoniation. The zeolite is typically treated with an ammonium chloride (NH₄Cl) solution at elevated temperatures (e.g., 80–100 °C) for several hours to achieve near-complete exchange of Na⁺ with NH₄⁺, leveraging the zeolite's ion-sieving pores. Thermal deammoniation then occurs via calcination at 400–600 °C, releasing NH₃ and generating Brønsted acid sites (Si-OH-Al bridges) while preserving framework integrity. This stepwise process enhances acidity for applications like cracking, with the extent of exchange (up to 90–100%) directly impacting site density. Precipitation and co-precipitation techniques are used to prepare mixed oxide solid acids, such as silica-alumina, by inducing supersaturation in precursor solutions to form gels or precipitates. In co-precipitation, solutions of TEOS and aluminum salts (e.g., Al(NO₃)₃) are mixed and precipitated using bases like NH₄OH, with control (typically 4–8) critical for gel formation and uniform Si/Al distribution to avoid phase separation. The -swing method, for instance, alternates acidic and basic conditions during precipitation to yield mesoporous silica-alumina with high surface areas (>400 m²/g) and tunable acidity. Aging the precipitate at 50–70 °C followed by washing, drying, and at 500–600 °C produces amorphous solids with dispersed acid sites, where lower favors tetrahedral aluminum coordination for stronger Brønsted acidity.

Surface Modification

Surface modification of solid acids involves post-synthesis techniques to tailor surface properties, such as acidity strength, site density, and selectivity, by introducing functional groups or active onto existing structures. These alterations enhance catalytic without altering the bulk composition, enabling better , , and resistance to deactivation in targeted applications. Common methods include impregnation, sulfation, doping, and encapsulation, each addressing specific limitations like or poor accessibility in harsh environments. Impregnation is a widely used technique for loading heteropolyacids (HPAs) onto porous supports to improve acid dispersion and prevent aggregation. In wet impregnation, the support, such as mesoporous silica SBA-15, is contacted with an aqueous or alcoholic solution of the HPA (e.g., H₃PW₁₂O₄₀), followed by solvent evaporation and drying typically at 60–100°C to fix the acid on the surface. This method yields highly active catalysts with tunable loading levels, often 10–30 wt%, enhancing Brønsted acidity while maintaining the support's high surface area (around 600–800 m²/g). For instance, HPA-impregnated SBA-15 exhibits superior performance in acid-catalyzed reactions due to the uniform distribution of Keggin units within mesopores. Sulfation introduces sulfate groups to generate superacidic sites on metal oxides, significantly boosting proton acidity. For sulfated zirconia (SZ), this is achieved by impregnating ZrO₂ or Zr(OH)₄ with ammonium sulfate ((NH₄)₂SO₄) solution, followed by drying and calcination at approximately 600°C to decompose the precursor and graft SO₄²⁻ onto zirconium sites, forming bidentate sulfate complexes. The resulting SZ displays Hammett acidity (H₀) values as low as -12 to -16, exceeding that of concentrated sulfuric acid, with enhanced thermal stability up to 500°C. This modification transforms weakly acidic zirconia into a versatile superacid catalyst suitable for isomerization and alkylation processes. Doping incorporates metal nanoparticles into solid acid frameworks via to create bifunctional catalysts combining /dehydrogenation with sites. In this process, a metal precursor (e.g., H₂PtCl₆ for ) is added dropwise to the until the pore volume is saturated, followed by drying, reduction (often with H₂ at 300–400°C), and sometimes . For example, doping in H-Beta or zeolites at 0.5–2 wt% loading enables synergistic metal- catalysis, improving selectivity in hydroisomerization by facilitating rearrangements on sites after metal-mediated transfer. This approach enhances overall activity while minimizing formation on sites. Encapsulation employs polymer coatings to shield solid acid particles from environmental degradation, particularly in aqueous conditions. , a perfluorosulfonic acid , is commonly used in composites where inorganic acids (e.g., silica-supported HPAs or zeolites) are embedded or coated with via solution casting or , forming a protective matrix that retains acidity while improving hydrothermal stability. These composites maintain proton conductivity above 0.1 S/cm at 80°C and exhibit reduced in -saturated environments, with durability extended by 20–50% compared to uncoated acids due to the polymer's chemical inertness and water retention properties. Such modifications are crucial for applications requiring operation in humid or liquid-phase media.

Applications

Catalytic Processes

Solid acids play a pivotal role in accelerating chemical transformations through , particularly in refining and production, where they facilitate the breaking and rearranging of molecular bonds under controlled conditions. In (FCC), zeolites such as Y-type and serve as primary catalysts to convert heavy vacuum gas oils into valuable lighter fractions, including , by promoting the cracking of long-chain hydrocarbons into shorter alkenes and alkanes. This process operates in fluidized-bed reactors, where finely powdered catalyst particles (typically 60-75 μm) are suspended in an upward-flowing stream of feedstock and steam, enabling continuous regeneration of the catalyst via coke combustion in a separate stripper-regenerator unit. , often occurring concurrently in FCC, enhances octane by rearranging linear alkanes into branched isomers, with additives boosting light olefin yields like through selective cracking of C7-C13 hydrocarbons. In processes, solid acids enable the combination of with olefins such as to produce high-octane alkylate for blending, with sulfated zirconia emerging as a robust alternative to liquid acids due to its superacidity. Sulfated zirconia catalyzes this reaction via of the olefin, leading to selective formation of trimethylpentanes with minimal side products like dienes, achieving high selectivities to trimethylpentanes. For , heteropolyacids (HPAs) such as tungstophosphoric acid immobilized on mesoporous supports like SBA-15 effectively esterify free fatty acids (FFAs) in low-grade feedstocks with , converting acids like palmitic or oleic into methyl esters with yields over 90% while tolerating impurities that deactivate homogeneous catalysts. This pretreatment step reduces FFA content below 1 wt%, enabling subsequent base-catalyzed . Recent advances as of 2025 include the development of recyclable magnetic brush solid acids and one-step prepared catalysts for efficient synthesis from waste cooking oils and valorization, promoting production with high yields and reusability. The underlying reaction mechanisms in these processes predominantly involve Brønsted acid sites on solid acids, which donate protons to generate intermediates that drive skeletal rearrangements and bond cleavages. For instance, in or cracking, of a forms a , which can then undergo hydride shifts or beta-scission: \ce{R-CH=CH2 + H+ -> R-CH+-CH3 ->[isomerization] R-CH(CH3)-CH3} This carbocation pathway is central to Brønsted site activity in zeolites and sulfated oxides, where the proton affinity and site confinement influence selectivity toward branched products or lighter fractions. In HPA-catalyzed esterification, the Brønsted acidity protonates the carbonyl oxygen of FFAs, facilitating nucleophilic attack by methanol to form the ester and water. Process conditions for solid acid vary by application but typically span temperatures of 200-500°C for conversions to balance activity and deactivation. Fixed-bed reactors are favored for and esterification, offering steady-state operation with pellets packed in columns, though they risk hot spots in exothermic reactions; reactors, where is suspended in liquid feed, provide better for viscous or high-heat duties like initial cracking stages. In FCC, temperatures around 500-550°C and -to-oil ratios of 3-7 ensure optimal yields (40-50 wt%) while minimizing formation (under 1 wt%). For sulfated zirconia , lower temperatures (0-100°C) in fixed or pulsed-flow beds enhance selectivity, often with co-feeds like supercritical CO2 to suppress oligomerization. esterification proceeds efficiently at 60-150°C in batch or fixed-bed setups, with alcohol-to-acid ratios of 10-20:1 to drive toward esters.

Non-Catalytic Uses

Solid acids play a significant role in and adsorption processes, particularly through protonated sites that enable selective binding. Zeolites, as solid acids, are widely employed for via , where calcium and magnesium ions in are replaced by sodium or ions, reducing formation in industrial and household applications. For instance, zeolite-based monoliths have demonstrated effective hardness removal from , achieving up to 90% reduction in calcium content while minimizing production compared to traditional methods. Additionally, protonated zeolites facilitate CO₂ capture by adsorption, leveraging their microporous structure and acidic sites to selectively bind CO₂ molecules under ambient conditions, with capacities reaching 2-4 mmol/g at 25°C and 1 bar. In proton conduction applications, solid acids like , a perfluorosulfonic acid , serve as membranes in fuel cells due to their high proton mobility via groups. enables efficient H⁺ transport across the membrane while blocking electron conduction, supporting proton exchange membrane fuel cells (PEMFCs) with conductivities exceeding 0.1 S/cm at 80°C and 100% relative humidity. This property arises from the hydrated channels formed by the acidic sulfonate groups, which facilitate the for proton hopping. Solid acids are utilized in sensors and detectors, especially pH-sensitive materials derived from layered clays, which respond to environmental pH changes through surface charge variations and . Acid-activated clays, such as , exhibit pH-dependent swelling and intercalation, enabling their integration into electrochemical sensors for monitoring or acidity in environmental applications. For example, nanoclays modified with pH-sensitive indicators detect onset on metal surfaces by color change or potential shift, offering real-time alerts with sensitivity to pH shifts as low as 0.1 units. (LDHs) functionalized as solid acids further enhance selectivity in detection for monitoring, achieving detection limits in the micromolar range for in aquatic environments. As fillers in composite materials, solid acids improve polymer performance in non-catalytic roles, such as enhancing flame retardancy and anticorrosion properties. , a weak solid acid, is incorporated into polymer matrices like coatings to promote formation during , reducing peak heat release rates by up to 30% in flame tests. In anticorrosion applications, sulfonic acid-functionalized solid particles, such as silica-based acids, are added to resins to create pH-responsive barriers that inhibit metal , extending coating lifetimes in acidic environments by forming protective passivation layers. These composites leverage the acidic sites for cross-linking and , without relying on catalytic activity.

Advantages and Challenges

Environmental and Economic Benefits

Solid acids offer significant environmental advantages over traditional liquid acids, primarily through reduced of process equipment and minimization of waste generation. Unlike liquid acids such as , which can cause severe requiring specialized materials and frequent maintenance, solid acids operate in heterogeneous systems that limit direct contact with reactor walls, thereby extending equipment lifespan and decreasing the need for corrosive-resistant alloys. Additionally, solid acid eliminates the production of aqueous effluents common in homogeneous acid processes, as the catalyst remains insoluble and can be separated without generating liquid waste streams, aligning with principles of waste prevention in . This recyclability further enhances ; for instance, many solid acid catalysts maintain activity over multiple cycles, with some demonstrating stability for up to 5 reuse cycles without significant performance loss in esterification reactions. These materials also embody key tenets, such as , by facilitating reactions that maximize incorporation of reactants into the desired product while minimizing byproducts. Heteropolyacids (HPAs), a prominent class of solid acids, exemplify this in solvent-free syntheses, such as the for dihydropyrimidines, where they enable high yields under mild conditions without volatile solvents, reducing environmental from solvent disposal. In lignocellulosic biorefineries, solid acids promote efficient conversion to platform chemicals, minimizing harsh chemical use and waste, which supports sustainable processing of renewable resources. Recent developments as of 2025 include the application of single-atom solid acid catalysts for improved selectivity in biomass valorization. Economically, solid acids lower operational costs through simplified product-catalyst separation via rather than energy-intensive required for liquid acid recovery. This approach can reduce separation energy demands by orders of magnitude in processes like , where allows straightforward catalyst reuse. Moreover, their extended lifetimes contribute to cost savings; zeolites used in (FCC) units, for example, typically endure for about one month under continuous operation, far outlasting disposable liquid catalysts and amortizing replacement expenses over high-throughput industrial scales. Lifecycle analyses underscore these benefits, revealing that solid acid processes consume less energy for handling and disposal compared to hazardous liquid acids, which require stringent protocols and neutralization steps. One study comparing solid acid to conventional methods found the former to have approximately 3.9 times lower adverse environmental impact potential, driven by reduced energy for and . Post-2020 developments have increasingly emphasized solid acids in bio-based feedstocks, such as lignocellulosic biomass for production, enhancing overall by integrating renewable inputs with low-impact .

Limitations and Deactivation Issues

Solid acids, particularly in catalytic applications, are prone to deactivation through several mechanisms that reduce their active surface area and acidity over time. One primary cause is , where carbonaceous deposits form on active sites, blocking access for reactants and leading to a gradual loss of catalytic activity, especially in processes involving cracking over zeolites. represents another key deactivation pathway, occurring at elevated temperatures where metal particles or the catalyst framework agglomerate, resulting in a significant reduction in surface area and dispersion of active components. , often by -containing impurities in feedstocks, further exacerbates deactivation by strongly adsorbing onto acid sites and rendering them inactive, with sulfur species forming stable sulfates or sulfides that are difficult to remove. Selectivity challenges in solid acid catalysts, such as zeolites, arise from over-cracking of hydrocarbons, where strong Brønsted acid sites promote excessive C-C bond cleavage, yielding low-value light gases or instead of desired products like olefins or gasoline-range fractions. This issue is particularly pronounced in microporous structures, where confined reaction environments favor sequential cracking reactions over controlled product formation. Scalability of advanced solid acids, including metal-organic frameworks (MOFs), is hindered by high synthesis costs associated with complex ligand-metal assembly and purification steps, limiting their adoption despite promising tunability. Additionally, limitations within narrow pores restrict reactant and product egress, reducing effective catalytic rates in large-scale reactors and exacerbating deactivation by promoting localized . To mitigate these limitations, regeneration techniques such as oxidative burning of deposits at approximately 500°C restore activity by combusting carbon residues, though repeated cycles can lead to framework damage if not controlled. design strategies, including the incorporation of hierarchical pore structures, enhance anti-coking performance by improving mass transport and reducing for coke precursors, thereby extending operational lifetime in zeolite-based systems.

Notable Examples

Traditional Solid Acids

Traditional solid acids encompass a class of well-established materials that have been pivotal in industrial and ion-exchange applications for decades. These include zeolites, amorphous silica-aluminas, perfluorosulfonic acid resins like , sulfated metal oxides such as sulfated zirconia, and modified clays, each characterized by distinct structural features that confer acidity and selectivity. Their maturity stems from extensive since the mid-20th century, enabling reliable performance in large-scale processes without the complexities of emerging . Sulfated zirconia (ZrO₂/SO₄²⁻), developed in the 1980s and commercialized in the , exemplifies solid superacids with tetragonal zirconia phases stabilized by groups, providing strong Brønsted and acidity (Hammett H₀ ≈ -16) for of n-butane and reactions. Zeolites such as H-Y and H-ZSM-5 represent cornerstone examples of crystalline aluminosilicates with microporous frameworks, providing shape-selective through well-defined channels and cavities. H-Y zeolite adopts the faujasite (FAU) framework type, featuring a three-dimensional network of sodalite cages and supercages accessible via 12-membered ring windows with diameters around 0.74 nm, which facilitates the accommodation of larger molecules in acidic environments. This structure arises from the tetrahedral arrangement of SiO₄ and AlO₄ units, where aluminum substitution generates Brønsted acid sites upon proton exchange. H-Y zeolite, synthesized in 1959, has been employed in (FCC) since its commercialization in the early , with rare-earth exchanged variants widely adopted in the 1970s for upgrading heavy hydrocarbons into gasoline and olefins by promoting rearrangements and β-scission mechanisms. Additionally, H-forms of faujasite-type zeolites contribute to formulations as acid catalysts for enhancing cleaning efficiency, though their primary role remains in for . In contrast, H-ZSM-5 possesses the MFI framework type, characterized by a two-dimensional system of straight 10-membered ring channels (0.53 × 0.56 nm) intersecting sinusoidal 10-membered ring channels (0.51 × 0.55 nm), enabling high shape selectivity for medium-sized molecules. Discovered in 1969 and commercialized in the 1970s by , H-ZSM-5 has revolutionized in processes like methanol-to-gasoline conversion and aromatic due to its moderate acidity and resistance to deactivation. Amorphous silica-alumina catalysts, lacking the long-range order of zeolites, offer a complementary acidity profile suited to hydrocracking applications. These materials consist of a disordered of silica and alumina domains, typically with 10-30 wt% Al₂O₃ content, where aluminum incorporation generates both Brønsted and acid sites at the interface between SiO₄ and AlO₄ tetrahedra or octahedral alumina clusters. The amorphous structure results in broader size distributions (2-50 ) compared to zeolites, allowing better of bulky reactants like vacuum gas oils. Since the , but with optimization in the mid-20th century, these catalysts have been integral to hydrocracking units, where they facilitate ring opening and chain cleavage under pressure, often supported on carriers like γ-Al₂O₃ and promoted with metals such as and for improved activity. Nafion, a perfluorosulfonic acid resin developed by in the , exemplifies polymeric solid acids with exceptional chemical stability. Its structure comprises a backbone with perfluorovinyl ether side chains terminating in groups (-SO₃H), yielding a phase-separated morphology of hydrophilic ionic clusters (2-4 nm) embedded in a hydrophobic fluorocarbon matrix. This design imparts a high ion-exchange capacity of approximately 0.9 meq/g, primarily through the sulfonic acid moieties that provide strong Brønsted acidity (pKₐ ≈ -6). Nafion's applications extend to compositions in proton-exchange membranes for fuel cells, where it conducts protons via vehicular and Grotthuss mechanisms while blocking electrons and fuels. Acid-treated bentonites and pillared clays derive from layered minerals, offering tunable acidity via modification of their 2:1 phyllosilicate structure. Bentonites, primarily , feature sheets separated by exchangeable interlayer cations, with basal spacing around 1.0 nm in the hydrated state. Acid treatment with acids like HCl or H₂SO₄ protonates the layers, octahedral Al³⁺ and Fe³⁺ to increase surface area (up to 300 m²/g) and expose edge groups as weak acid sites. Pillared variants involve intercalation of polyoxocations (e.g., Al₁₃ or Zr₄ clusters) between layers, followed by to form pillars that prop open the interlayer space to 1.8-2.5 nm, creating mesoporous galleries while preserving the sheet integrity. This expansion enhances accessibility for larger adsorbates and generates moderate Brønsted acidity from pillar , making pillared bentonites effective solid acids for esterification and oligomerization reactions since their development in the 1970s.

Advanced and Emerging Solid Acids

Advanced solid acids represent a class of heterogeneous catalysts engineered with enhanced acidity, stability, and selectivity through nanostructuring and functionalization, addressing limitations of traditional materials like zeolites or sulfated oxides. These catalysts often incorporate tunable active sites, such as Brønsted or acid centers, into frameworks with high surface areas and porosity, enabling efficient performance in demanding reactions like valorization and synthesis. Emerging designs leverage metal-organic frameworks (MOFs), supported ionic liquids (ILs), and heteropolyacids (HPAs) to achieve superior recyclability and environmental compatibility. MOF-based solid acids have gained prominence due to their modular , allowing precise integration of acidic functionalities via metal nodes, linkers, or post-synthetic modifications. For instance, UiO-66 and MIL-101 frameworks functionalized with groups (-SO₃H) exhibit strong Brønsted acidity and large pore volumes (>1 cm³/g), facilitating the conversion of biomass-derived sugars to platform chemicals like 5-hydroxymethylfurfural (5-HMF). In one application, -modified MIL-101(Cr) achieved a 39.8% of 5-HMF from glucose in /γ-valerolactone (1:9) media, outperforming conventional solid acids by enabling shape-selective within hierarchical pores. Similarly, sulfated MOF-808 demonstrates superacidity (H₀ ≤ -14.5). For to , Ga₂O₃-UiO-66 achieves up to 32% under mild conditions (240 °C), highlighting the role of defect engineering in enhancing accessibility. These materials offer advantages over bulk solid acids, including higher recyclability (up to 5 cycles with >90% retention) and reduced , due to robust coordination bonds. Supported heteropolyacids and ionic liquids further advance solid acid by combining the strong acidity of HPAs (e.g., , H₃PW₁₂O₄₀) with the tunability of ILs on stable supports like mesoporous silica or MOFs. HPA-IL hybrids, such as [HMIm]₃[PW₁₂O₄₀] immobilized on Fe-based MOFs, provide bifunctional acidity and exhibit thermal stability up to 598 °C, enabling solvent-free acetalization of to —a additive—with 100% conversion and selectivity at . These catalysts maintain performance over 7 cycles, attributed to the ionic interactions preventing HPA leaching. In , IL-functionalized SBA-15 supports bearing yield 88.1% fatty acid methyl esters from , with 80% retention after 5 reuses, surpassing homogeneous in ease of separation and reduced . Encapsulation in UiO-66 frameworks enhances this further, achieving 95% conversion in due to confined acidic microenvironments. Magnetic nano-sized solid acids emerge as versatile platforms for facile recovery in continuous processes, often featuring grafts on ferrites. For example, AlFe₂O₄ nanoparticles functionalized with propyl-ethyl- groups display dual weak/strong sites (desorption peaks at 350 °C and 630 °C via NH₃-TPD) and (saturation magnetization 27 emu/g), allowing post-reaction. This catalyst delivers 98% yield in esterification for at 60 °C and high-efficiency sulfide oxidation under solvent-free conditions, reusable for 5 cycles with minimal activity loss. Such designs underscore the shift toward multifunctional, catalysts that integrate acidity with practical handling, reducing operational costs in industrial-scale applications.

References

  1. [1]
    [PDF] A Perspective on Catalysis in Solid Acids Raymond J. Gortea and ...
    Introduction. Solid acids are among the most important heterogeneous catalysts, finding many applications ranging from hydrocarbon cracking and alkane ...
  2. [2]
    Solid Acids for Green Chemistry | Accounts of Chemical Research
    A wide range of important organic reactions can be efficiently catalyzed by these materials, which can be designed to provide different types of acidity as well ...
  3. [3]
    Solid Acid Catalyst - an overview | ScienceDirect Topics
    Solid-acid catalysts are heterogeneous catalysts synthesized by embedding acidic functional groups on the surface of a solid matrix. Various classes of ...
  4. [4]
    Houdry Process for Catalytic Cracking - American Chemical Society
    In 1930, H. F. Sheets of the Vacuum Oil Company learned of Houdry's promising results using a catalyst to convert vaporized petroleum to gasoline, and invited ...Missing: acid | Show results with:acid
  5. [5]
    Eugene Houdry - Science History Institute
    After testing hundreds of catalysts to effect the hoped-for molecular rearrangement, Houdry began working with silica-alumina and changed his feedstock from ...Missing: solid acid
  6. [6]
    PAUL B. WEISZ | Memorial Tributes: Volume 19
    Along with several Mobil collaborators, he pioneered the use of natural and synthetic zeolites as catalysts for petroleum refining and petrochemical manufacture ...
  7. [7]
    (PDF) Paul B. Weisz - ResearchGate
    Aug 6, 2025 · Paul B. Weisz, 93, an internationally recognized expert in the areas of zeolite catalysts and diffusion died on Tuesday, September 25, 2012, in State College, ...
  8. [8]
    Periodic density functional study of superacidity of sulfated zirconia
    Introduction. Sulfate-modified zirconia (S-ZrO2), which has superacidic surface, shows a superior acid catalytic activity for many reactions ...
  9. [9]
    [PDF] George A. Olah - Nobel Lecture
    Solid superacids, possessing both Bronsted and Lewis acid sites, are of increasing significance. They range from supported or intercala- ted systems, to highly ...
  10. [10]
    Heteropoly Acid-Based Catalysts for Hydrolytic Depolymerization of ...
    Sep 24, 2020 · Heteropoly acid (HPA), as a green catalyst, seems to be more conducive to the degradation of cellulosic biomass due to its extreme acidity.Missing: MOFs 2020s
  11. [11]
    Metal organic frameworks for biomass conversion - RSC Publishing
    May 12, 2020 · This review focuses on the use of metal–organic frameworks (MOFs) and derivatives as catalysts for biomass conversion.
  12. [12]
    The Heteropolyacid-Catalyzed Conversion of Biomass Saccharides ...
    Nov 18, 2024 · Keggin heteropolyacids are efficient catalysts for obtaining saccharides from cellulose and hemicellulose and converting them into bioproducts or biofuel.
  13. [13]
    Molecular Views on Mechanisms of Brønsted Acid-Catalyzed ...
    Mar 30, 2023 · Molecular Views on Mechanisms of Brønsted Acid-Catalyzed Reactions in Zeolites ... DOI: 10.1016/S0920-5861(00)00577-0. Google Scholar. 200.
  14. [14]
    Revealing Brønsted Acidic Bridging SiOHAl Groups on Amorphous ...
    Zeolites and amorphous silica-alumina (ASA), which both provide Bronsted acid sites (BASs), are the most extensively used solid acid catalysts in the chem.
  15. [15]
    Heteropoly Acid-Based Catalysts for Hydrolytic Depolymerization of ...
    In this review, the characteristics, advantages, and applications of HPAs in different categories for cellulose degradation, especially hydrolysis hydrolytic ...
  16. [16]
    Rational Design of Metal Oxide Solid Acids for Sugar Conversion
    The catalytic performance is mainly dependent on the types of solid acids. Amorphous metal oxides possess unsaturated metal cations that function as Lewis acid ...
  17. [17]
    The Role of Lewis Acid Sites in γ-Al 2 O 3 Oligomerization
    Jun 9, 2023 · Olefin oligomerization to liquid hydrocarbons by γ-Al2O3 has recently been reported, and it was suggested that Lewis acid sites are catalytic.
  18. [18]
    Acidic Strengths of Brønsted and Lewis Acid Sites in Solid Acids ...
    For the TMP−Lewis acid complex, a linear correlation between the calculated 31P chemical shifts and corresponding binding energies was observed for the B-, Al-, ...
  19. [19]
    Copper-induced formation of Lewis acid sites enhancing sulfated ...
    The moderate Cu species additive is beneficial to release Zr Lewis acid site from sulfate and increase its Lewis acid density, which accounts for its excellent ...
  20. [20]
    Lewis Acid Rather than Brønsted Acid Sites of Montmorillonite K10 ...
    Lewis Acid Rather than Brønsted Acid Sites of Montmorillonite K10 Act as a Powerful and Reusable Green Heterogeneous Catalyst for Rapid Cyanosilylation of ...
  21. [21]
    Sulfated complex metal oxides solid acids with dual Brønsted-Lewis ...
    Feb 1, 2022 · In the present work, a series of sulfated complex metal oxides solid acids catalysts with both Lewis and Brønsted acid sites were synthesized ...
  22. [22]
    Quantitative characterization of the solid acidity of montmorillonite ...
    Therefore, in measurements using the Hammett indicator method, the n-butylamine molecule must enter the interlayer of montmorillonite for the quantitative ...
  23. [23]
    Characterization of Acid Sites Using Temperature-Programmed ...
    TPD of ammonia is a widely used method for characterization of site densities in solid acids due to the simplicity of the technique. Ammonia often overestimates ...
  24. [24]
    A Variable-Temperature Diffuse Reflectance Infrared Fourier ...
    Infrared spectroscopic investigation of pyridine-treated acidic solids can readily distinguish between Brönsted and Lewis acid sites on the surface and provide ...
  25. [25]
    Acid properties of solid acid catalysts characterized by solid-state 31 ...
    Jul 25, 2011 · Solid-state 31P NMR uses phosphorous probes to characterize acid sites, providing type, strength, distribution, and concentration of acid sites ...
  26. [26]
    a series of simple basic indicators. i. the acidity functions of mixtures ...
    Physical Organic Chemistry of Solid Acids: Lessons from in Situ NMR and Theoretical Chemistry. ... Hammett constants from density functional calculations: charge ...
  27. [27]
    Liquid Phase Calorimetric Method as a Tool for Acid Strength ... - MDPI
    This review seeks to present the fundamentals of the method and show different applications of the characterized catalysts for a variety of reactions.
  28. [28]
    An Overview of Solid Acid Catalysts in Lignocellulose Biorefineries
    Therefore, along with the strength of the acid site, the hydrothermal stability of the acid site is also a critical factor in determining the efficiency of the ...
  29. [29]
    Preparation and Characterization of Solid Acid Catalysts for ... - NIH
    Apr 24, 2022 · Surface and acidic properties of various solid acid catalysts. Hβ, CTAH, PTAH, FCHH. Surface area (m2/g), 489.39, 537.99, 360.61, 435.70.Missing: typical | Show results with:typical
  30. [30]
    [PDF] Defect Structures in Zeolite Crystals - UCL Discovery
    Aluminium framework substitutionals (the Brpnsted acid sites), hydroxyl nests and vicinal disilanols are the major defect species present in these materials.
  31. [31]
    Thermal stability studies of zeolites A and X synthesized from South ...
    Both zeolites A and X were stable up to approximately 800 °C, with X retaining structural integrity past 440 °C. TGA indicated no significant weight loss ...<|separator|>
  32. [32]
    Recent progress in the improvement of hydrothermal stability ... - NIH
    May 17, 2021 · This review attempts to summarize the recently developed strategies to improve the hydrothermal framework stability of zeolites.
  33. [33]
    Relation between Catalytic Activity and Size of Particle
    Controlling the SAPO-34 Particle Size to Enhance Reaction-Diffusion Behavior for Improving Catalytic Efficiency in the MTO Reaction.Missing: mechanical | Show results with:mechanical
  34. [34]
  35. [35]
    A Novel Method To Synthesize Amorphous Silica−Alumina ...
    In this study, a novel sol−gel method is used to synthesize amorphous silica−alumina materials with a narrow mesoporous distribution and various Si/Al molar ...
  36. [36]
    Synthesis and characterization of amorphous silica-alumina with ...
    Nov 1, 2020 · Amorphous silica-aluminas with enhanced Brønsted acidity have been successfully synthesized via a novel acetic acid based sol-gel process.
  37. [37]
  38. [38]
    Change in local coordination structure of aluminum cations in silica ...
    In this work, we prepared silica–alumina from the mixed solution of tetraethoxysilane (TEOS) and aluminum salt by coprecipitation, and investigated the ...
  39. [39]
    Synthesis of amorphous silica-alumina with enhanced specific ...
    In this work, we successfully synthesized mesoporous ASA with enhanced specific surface area, Lewis acidity and total acidity by pH-swing method.
  40. [40]
    Direct synthesis, characterization and catalytic application of SBA-15 ...
    One possible route to obtain these supported HPW catalysts is the direct impregnation of the support with a heteropoly acid solution followed by evaporation of ...
  41. [41]
    H3PW12O40/SBA-15 for the Solventless Synthesis of 3-Substituted ...
    A family of silica-supported H3PW12O40 (HPW) solid acid catalysts was prepared by wet impregnation of mesoporous SBA-15 and investigated for the solventless ...
  42. [42]
    Acidity-Reactivity Relationships in Catalytic Esterification over ...
    New insight was gained into the acidity-reactivity relationships of sulfated zirconia (SZ) catalysts prepared via (NH4)2SO4 impregnation of Zr(OH)4 for ...
  43. [43]
    The Overlooked Potential of Sulfated Zirconia: Reexamining Solid ...
    Mar 20, 2024 · Superacidic sulfated zirconia (SZrO) with Hammet Acidity Functions (H0) = -12 (i.e., stronger than 100% H ...
  44. [44]
    The effects of ammonium sulfate and sulfamic acid on the surface ...
    In this work, sulfated zirconia (SZ) was prepared by a solvent-free method, ammonium sulfate ((NH4)2SO4) and sulfamic acid (NH2SO3H) were as sulfur sources, ...
  45. [45]
    Hierarchically structured Pt/K-Beta zeolites for the catalytic ...
    The Beta-HS zeolite was further modified with alkaline K through ion-exchange to passivate the acid sites and with metal Pt through incipient wetness ...
  46. [46]
  47. [47]
    Superacidity in Nafion/MOF Hybrid Membranes Retains Water at ...
    Nov 4, 2016 · Naf-1SZM composite membranes show better durability than Naf with OCV drop of 8.4%, compared to 11% for Naf. The proton conductivity in solid ...
  48. [48]
    Composite silica/Nafion® membranes prepared by ...
    Composite silica/Nafion® membranes were prepared using a tetraethylorthosilicate (TEOS) hydrolysis sol–gel reaction followed by solution casting.
  49. [49]
    Role of Catalyst in Optimizing Fluid Catalytic Cracking Performance ...
    Mar 12, 2021 · The catalyst system containing ZSM-5 increased the LPG yield but did not have an impact on gasoline octane. It was found that the predominant ...
  50. [50]
    Zeolite Catalysts | FSC 432: Petroleum Refining
    FCC catalyst consists of a fine powder with an average particle size of 60–75 μm and a size distribution ranging from 20 to 120 μm. Four major components make ...Missing: catalyzed | Show results with:catalyzed
  51. [51]
    Performance of ZSM-5 as a Fluid Catalytic Cracking Catalyst Additive
    In petroleum catalytic cracking, ZSM 5 zeolite enhances gasoline octane no. by cracking and isomerizing low-octane no., unbranched C7-13 hydrocarbons to C3 ...
  52. [52]
    Critical review on the active site structure of sulfated zirconia ...
    Feb 25, 2019 · This article critically reviews the literature on sulfated zirconia catalysts for variety of important hydrocarbon and oxygenate reactions.Missing: paper | Show results with:paper
  53. [53]
    Activity and Regenerability of Sulfated Zirconia Superacid Catalysts ...
    Alkylation of isobutane with 1-butene was carried out in the gas phase at atmospheric pressure over sulfate-promoted zirconia superacid catalysts prepared ...
  54. [54]
    Esterification of free fatty acids to biodiesel over heteropolyacids ...
    Dec 20, 2010 · Biodiesel can also be prepared by the esterification of free fatty acid present in animal fats (ex. lard or tallow), with methanol over acid ...
  55. [55]
    Review on Biodiesel Production from Various Feedstocks Using 12 ...
    The esterification takes place between free fatty acids (RCOOH) and methanol (CH3OH), whereas transesterification takes place between glycerides (RCOOR′) and ...
  56. [56]
    Bronsted and Lewis Acid Sites | FSC 432: Petroleum Refining
    You should remember that Bronsted acid sites donate protons, while Lewis acid sites accept electrons to form carbocations from hydrocarbons. Figure 7.2 ...
  57. [57]
    [PDF] Alkylation Technology Study FINAL REPORT South Coast Air ...
    Sep 9, 2016 · In the Solid Acid Alkylation process, the catalyst is a fixed bed and thus eliminates the hazards of acid handling, transportation, and storage.
  58. [58]
    Zeolite-based monoliths for water softening by ion exchange ...
    Mar 7, 2022 · The use of this monolithic softener can improve performance and sustainability of hardness removal from tap water, reducing the production of sludge.Missing: capture | Show results with:capture
  59. [59]
    Zeolites as Selective Adsorbents for CO2 Separation
    Feb 21, 2023 · Zeolites are a very versatile class of materials that can display selective CO 2 adsorption behavior and thus find applications in carbon capture, storage, and ...
  60. [60]
    An overview of the proton conductivity of nafion membranes through ...
    Apr 15, 2016 · In this study, we performed a statistical analysis on a comprehensive and representative collection of the proton conductivity of Nafion membranes.
  61. [61]
    Nanoclay composites in electrochemical sensors - Frontiers
    Oct 2, 2024 · Nanoclays are layered structures in the nanoscale range with widespread application due to their unique properties such as swelling, cation exchange capacity, ...
  62. [62]
    [PDF] pH and electrochemical responsive materials for corrosion smart ...
    The evaluation of nanoclays modified with pH sensitive colorants as sensors for onset of corrosion of polymer coated aluminum and ferrous substrates is part ...
  63. [63]
    Flame Retardant Coatings: Additives, Binders, and Fillers - MDPI
    For example, mineral acid (i.e., boric acid) and APP can be used as flame retardants, either separately or in combination, in intumescence fabrications [58,112] ...
  64. [64]
    Solid acids: Green alternatives for acid catalysis - ScienceDirect.com
    Nov 1, 2014 · Compared with liquid acid catalysts, solid acid catalysts have distinct advantages in recycling, separation, and environmental friendliness.
  65. [65]
    Facile synthesis of highly efficient and recyclable magnetic solid ...
    Aug 13, 2013 · The catalytic activity and recyclability of the solid acid catalyst are evaluated during three typical acid-catalyzed reactions: esterification, ...
  66. [66]
  67. [67]
    Economic and Environmental Impact Analyses of Solid Acid ...
    The environmental impact assessment shows that the conventional sulfuric acid process has a 3.9 times higher adverse environmental impact potential than the ...Missing: benefits acids
  68. [68]
    Fluid catalytic cracking: recent developments on the grand old lady ...
    The main application of zeolite ZSM-5 has been in FCC operation targeting an increased propylene yield. Fig. 2 clearly shows that the introduction of zeolite ...
  69. [69]
    Catalyst deactivation - ScienceDirect.com
    Deactivation can occur by a number of different mechanisms, both chemical and physical in nature. These are commonly divided into four classes, namely poisoning ...
  70. [70]
    Sintering process and catalysis
    Sintering process is the main factor in the catalyst deactivation. The deactivation of catalyst is the dramatic loss in the active surface area. Sintering ...
  71. [71]
    A comprehensive review of anti-coking, anti-poisoning and anti ...
    Due to the low melting point and high surface mobility, sintering process is accelerated, leading to the deactivation of the catalyst. Besides the sulfidation ...
  72. [72]
    Potential and challenges of zeolite chemistry in the catalytic ...
    Dec 21, 2015 · Higher temperatures with protonated zeolites allow C–C bond cracking forming gasoline and light naphtha fractions (with especially ZSM-5), ...
  73. [73]
    Toward Intrinsic Catalytic Rates and Selectivities of Zeolites in the ...
    Toward intrinsic catalytic rates and selectivities of zeolites in the presence of limiting diffusion and deactivation.
  74. [74]
    Investigation of Metal-Organic Frameworks (MOFs): Synthesis ...
    The scalability of MOF synthesis methods presents a challenge; however, recent studies are tackling this issue by introducing innovative synthesis techniques, ...
  75. [75]
    Insights into Mass Transfer Barriers in Metal–Organic Frameworks
    Apr 29, 2022 · The results indicate that uptake rates in all of the MOFs considered here are limited by mass transfer through crystallite surfaces.
  76. [76]
    Regeneration of catalysts deactivated by coke deposition: A review
    Zeolite catalysts deactivated by coke can be regenerated through three main methods: oxidation, gasification, or hydrogenation.
  77. [77]
    Acid-Modified Hierarchical Porous Rare-Earth-Containing Y Zeolite ...
    Jul 16, 2020 · ... coke formation mainly inside the micropores; however, the structure of the catalyst could be well recovered by burning the coke by calcination.
  78. [78]
    Metal–Organic Framework-Based Solid Acid Materials for Biomass ...
    Jun 21, 2021 · Metal–organic framework (MOF)-based solid acid materials have been considered as promising catalysts in biomass transformation.
  79. [79]
    Heteropoly Ionic Liquid Functionalized MOF-Fe: Synthesis ...
    Apr 4, 2023 · Nevertheless, these acidic homogeneous catalysts have drawbacks, including using cost, toxicity, nonrecyclability, corrosivity, difficulty in ...
  80. [80]
    Recent advances in supported acid/base ionic liquids as catalysts ...
    This mini-review focuses on recent advances in supported acid/base ionic liquids to synthesize ionic liquid-functionalized materials for producing biodiesel.Missing: emerging | Show results with:emerging
  81. [81]
    Magnetic nano-sized solid acid catalyst bearing sulfonic acid groups ...
    Jan 9, 2025 · The AlFe2O4@n-Pr@Et-SO3H materials demonstrated excellent performance in both the esterification of oleic acid and the oxidation of sulfides.Biodiesel Production · Catalytic Studies · Esterification Reactions