Next Article in Journal
Use of a 3D Workpiece to Inductively Heat an Ammonia Cracking Reactor
Previous Article in Journal
Mechanochemical Solvent-Free Synthesis and Biological Profiling of Novel 2-Hydrazone-Bridged Benzothiazoles as Potent Anticancer Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Comprehensive Review of SBA-15 Mesoporous Silica: Functionalization Strategies, Diffusion Mechanisms, and Emerging Applications

1
Chemical Engineering Department, Universidad de Alicante, Apartado 99, 03690 Alicante, Spain
2
UCE-GIIP, Grupo de Investigación en Ingeniería de Procesos, Facultad de Ingeniería Química, Universidad Central del Ecuador, Quito 170129, Ecuador
*
Author to whom correspondence should be addressed.
Sustain. Chem. 2025, 6(4), 42; https://doi.org/10.3390/suschem6040042
Submission received: 23 September 2025 / Revised: 15 October 2025 / Accepted: 27 October 2025 / Published: 3 November 2025

Abstract

Mesoporous materials have attracted increasing attention due to their ordered pore systems; tunable surface chemistry; and versatile applications in catalysis, adsorption, and environmental technologies. Among them, SBA-15 stands out for its large surface area, uniform mesopores, and high hydrothermal stability, which make it a promising platform for gas adsorption and mass transport studies. This review examines the functionalization of SBA-15 through strategies such as post-synthesis grafting and co-condensation, focusing on the introduction of amines, thiols, and organometallic species that enhance selectivity, adsorption capacity, and thermal stability. The discussion integrates classical diffusion models, including Fickian and Knudsen transport, with more advanced approaches such as the Maxwell–Stefan formalism, to describe molecular transport within mesoporous networks and highlight the role of van der Waals interactions in gas capture processes. Special emphasis is placed on the relationship between structural features and diffusive behavior, supported by recent advances in computational modeling and spectroscopic validation. Applications in CO2 capture, heterogeneous catalysis, drug delivery, and environmental remediation are critically assessed to illustrate the versatility of functionalized SBA-15. This review concludes by outlining future perspectives on the rational design of hierarchical and multifunctional mesoporous materials for clean energy conversion, pollutant removal, and biomedical applications.

Graphical Abstract

1. Introduction

Mesoporous materials are a class of hybrid inorganic or organo-inorganic solids characterized by the presence of pores in the range of 2 to 50 nanometers, according to the classification of the International Union of Pure and Applied Chemistry (IUPAC) [1]. They have emerged as versatile platforms in nanoscience due to their specific surface area (600–1500 m2 g−1) depending on the synthesis method [2], typical pore volumes between 0.5 and 2.5 cm3 g−1 [3], morphology defined at the nanometer scale, wall thicknesses of ~3–6 nm depending on synthesis conditions, and their ability to form ordered structures with uniform porosity [4].
This combination of characteristics has consolidated its relevance in applications such as heterogeneous catalysis [5,6], selective separation [7], and adsorption of gaseous and liquid compounds. Among them, materials such as Santa Barbara Amorphous-15 (SBA-15) have stood out for their thermal stability up to 800 °C, thicker porous walls (3.1–6.4 nm) related to materials such as MCM-41 [8,9], and the presence of interconnecting micropores that enhance molecular transport [9], which allows precise control of the porous architecture [10], controlled drug release [11], environmental remediation [12,13] and versatility to be chemically modified.

1.1. Relevance and Applications of Mesoporous Materials

These characteristics, along with the structural tunability through synthesis parameters (surfactant type, pH, temperature, and template removal method), enable the tailoring of their physicochemical properties [4,14]. For specific application requirements, functional groups can be selectively anchored onto the surface to generate functionalized mesoporous materials.
Including specific chemical groups on the surface of the material is used for forming functionalized materials with the purpose of modifying the physicochemical properties of mesoporous materials to adapt them to requirements.
Chemical functionalization of their surfaces with groups such as amines (–NH2), sulfhydryls (–SH), carboxyls (–COOH) and phosphonates (–PO3H2) has been shown to enhance the selectivity towards target molecules in adsorption and separation processes [9,15]. These groups interact by J.D van der Waals forces, hydrogen bonds or electrostatic interactions, extending their field of application towards environmental, medical and energy technologies.
Functionalized SBA-15 has shown potential in different fields such as environmental, biomedical and energy. In the environmental field, they are used in carbon dioxide capture (up to 2.5 mmol g−1 at 1 bar and 25 °C) and heavy metal removal, reaching efficiencies above 95% for Pb2+ ions [10,12]. In the biomedical field, their compatibility and controlled porosity make them effective systems for controlled drug release, with demonstrated molecular loading efficiencies ranging from 80 to 95% for compounds such as doxorubicin and ibuprofen [11,16,17]. In the energy field, mesoporous materials such as SBA-15 stand out for their thermal stability and tunable morphology, being promising for the storage of combustible gases. For example, capacities of 150 v/v have been reported for CH4 (methane, a key vector in the energy transition) [18,19] and 2.5 wt % for H2 (hydrogen, essential for clean energies) at 77 K [20,21]. These properties make them ideal candidates for applications such as natural gas vehicle (NGV) reservoirs or hydrogen storage systems in fuel cells.
The possibility of adjusting pore size and morphology by controlling synthetic parameters such as temperature, pH, surfactant type and organic template removal methods enables the optimization of the diffusion of species within the material. This aspect is necessary to maximize efficiency in dynamic processes such as gas adsorption [10,22], substrate diffusion in heterogeneous catalysis or controlled release in biomedicine [23]. Recent advances have expanded their applications towards advanced catalysis, where they function as support for metal nanoparticles (2–5 nm) with optimal dispersion that increases catalytic activity in biomass refining reactions (yields > 90%) [8,20]. This multifunctionality is enhanced by the possibility of creating hierarchical structures combining micro-, meso- and macropores, optimizing molecular transport mechanisms [14,21]. The integration of organic groups into their inorganic matrix by hybrid synthesis methods has opened new routes for the development of smart materials with response to external stimuli such as pH, temperature or magnetic fields [4,24].
Regarding the environmental sector, mesoporous materials have been used in carbon dioxide capture and storage processes, volatile organic compounds (VOCs) removal, and treatment of toxic gases [9]. In medicine and biotechnology, mesoporous materials have been used as vehicles for active pharmaceutical ingredients, enabling protection and stimulus-responsive release such as pH or temperature. Their biocompatibility, structural integrity, and ease of functionalization with tissue-targeting groups make SBA-15 a promising platform for anti-tumor therapy, intelligent drug delivery systems, and advanced formulations [11,16].
These materials have garnered increasing research attention, with more than 15,000 publications indexed in Scopus since 2015, particularly in emerging technologies such as molecular sensors, photonics, and flexible electronics [8]. Their industrial scalability has been demonstrated in processes such as the production of heterogeneous catalysts and large-scale water purification systems [4,13,23].

1.2. Review Methodology

This review was conducted through a structured search of Scopus, Web of Science, ScienceDirect, and the MDPI platform using the keywords “SBA-15”, “functionalized mesoporous silica”, and “diffusion”, covering 2015–2025 publications (seminal earlier works were included for context). Inclusion criteria comprised peer-reviewed studies reporting synthesis/functionalization of SBA-15 or closely related mesostructures, quantitative textural/transport metrics (BET, pore size, D e f f   ), or application performance (adsorption capacity, catalytic turnover, release kinetics). Non-peer-reviewed items and reports lacking quantitative data were excluded.

2. Structural Characterization and Comparison Between Materials

These materials are characterized by a three-dimensional network of pores with controlled diameters, which endows them with unique properties. Their ordered structure results from cooperative self-assembly between surfactants (such as Pluronic P123) and inorganic precursors, followed by thermal or chemical removal of the organic template [25]. The structural versatility of these materials is manifested in three main configurations (Figure 1): (i) hexagonal (p6mm) phases such as SBA-15 and MCM-41 with unidirectional channels, (ii) cubic (Ia3d) structures such as KIT-6 with interconnected three-dimensional porosity, and (iii) lamellar systems such as FSM-16 with two-dimensional arrangement [4,26].
Table 1 presents a benchmark of the most published mesoporous materials, highlighting their structural parameters and characteristic synthesis methods [25,27,28].
Control over the synthesis of materials allows to design mesostructures with different orientations, connectivities and surface morphologies, which significantly influences the efficiency of adsorption, encapsulation or catalysis processes. In this sense, mesoporous materials act as a structural interface between microporous zeolites (<2 nm) and macroporous materials (>50 nm), offering accessible, functionalizable, and thermally stable surfaces [31].
Structural characterization is essential to correlate the physicochemical properties of the material with its functional performance. Key parameters such as surface area (600–1200 m2 g−1), pore volume, and wall thickness are strongly dependent on the type of surfactant, acidity of the medium, water/precursor ratio and thermal aging conditions [31,32].
Modification of these parameters favors accessibility to active sites, increases catalytic efficiency and improves host-guest interaction. A larger surface area facilitates the modification of surface or molecular charge [4,25]. In particular, SBA-15 presents secondary micropores (0.5–2 nm) connected to the main mesoporous channels, which increases the accessibility to active sites (>90%) and improves molecular diffusion. For the case of CO2, with reported effective diffusivities D e f f   on the order of 10 7   m 2 s 1 outperforming non-hierarchical analogues of similar composition [31].
The surface properties of mesoporous materials can be modulated by chemical modification through the targeted inclusion of functional groups such as amino (-NH2), thiol (-SH), carboxyl (-COOH) or sulfonate (-SO3H). These groups enable specific interactions with gases, metal ions or biomolecules by electrostatic interactions or J.D van der Waals forces [17,23]. The introduction of amino groups increases the adsorption capacity of CO2 up to 2.8 mmol g−1 (1 bar, 25 °C), 150% higher than non-functionalized materials. However, the heterogeneous distribution of active sites in pores <4 nm remains a challenge for large-scale applications [33]. On the other hand, adding noble metals such as Pt or Pd by impregnation methods generates catalytic centers with high surface dispersion (around 80%), improving the activity in hydrogenation or selective oxidation reactions. Mesoporous materials can be composed of pure silica as in the case of SBA-15, or incorporate heteroelements such as Al, Ti, Fe or Zr by direct synthesis or post-synthesis modification, this change in composition affects the acid-base, redox properties and thermal stability [4].
The hydrophilicity of mesoporous materials is primarily determined by the density of surface silanol groups and by the heat treatment applied during or after synthesis. This property can be tuned by functionalization with organosilanes, additional silanols, or plasma treatments, to enhance compatibility with aqueous or biological media. Controlling hydrophilicity is particularly relevant in controlled drug release or the design of chemical sensors [17,32]. The surface activity of mesopores is determined by the functional groups available after synthesis or modified subsequently.
In addition to hydrophilicity, pore geometry is another critical factor determining the efficiency of these materials. The geometry of SBA-15 favors adsorption and catalysis processes requiring unidirectional molecular transport, whereas the cubic structure of KIT-6 is used for reactions demanding multidirectional diffusion [4,25]. Recent advances have demonstrated that precise control of secondary microporosity and surface functionalization allows the design of materials with tailored properties [3,17,21,31,32].

2.1. Structure and Synthesis of SBA-15

Ordered mesoporous materials such as SBA (Santa Barbara Amorphous) are versatile platforms in materials science. In particular, SBA-15, developed by Zhao in 1998, features a structure of hexagonally ordered cylindrical pores (space group p6mm) interconnected by micropores in the walls, facilitating molecular transport [8,14]. It typically exhibits pore diameters in the range of 5–30 nm, thick walls (~3–6 nm), a high specific surface area (600–1000 m2 g−1), and a pore volume reaching up to 1.5 cm3 g−1 [22,25].
SBA-15 consists of a pure amorphous SiO2 matrix, whose three-dimensional organization is induced by the micellar self-assembly of the surfactant P123 (EO20PO70EO20) [22,25], which acts as a structural template in acidic medium. Under these conditions, PPO chains of P123 become hydrophobic while PEO chains remain hydrated, leading to the formation of cylindrical micelles. Silica precursors as tetraethyl orthosilicate (TEOS) or Na2SiO3, hydrolyze and condense around the micelles through S0H+XI+ electrostatic interactions (S0 = P123, X = Cl, I+ = silica species) [22]. A thermal aging step at 100–130 °C enhances the ordering and wall thickness (3–6 nm) [14,27]. Calcination at 500–550 °C then removes the template, leaving a thermally stable mesoporous structure up to 800 °C.
In addition to P123, the chemistry of the system can be modified by introducing co-surfactants (CTAB), adjusting the pH, temperature, or type of silica precursor, which allows modulation of the morphology, porosity, and connectivity of the final solid. Such synthetic control enables the tailoring of SBA-15 to be tailored for specific applications requiring efficient molecular transport or advanced functional selectivity.

2.2. Synthesis Methods and Their Effect on Diffusive Properties

The synthesis of SBA-15 can be tailored by a variety of techniques, from conventional methods such as sol–gel to microwave-, ultrasound- or solvothermal-assisted processes. Each variant modifies the textural properties of the material, especially the pore size, wall thickness and crystallinity of the siliceous phase, which directly affects its efficiency in mass transport processes. Table 2 summarizes the main synthesis methods, operating conditions, and associated advantages.

3. Functionalization of Mesoporous Materials: Strategies, Properties

Modifying mesoporous materials represents a paradigm shift in the design of solids with tailored surface chemistry and diffusion behavior at the nanoscale [34,35]. This section critically analyzes the functionalization strategies, their impact on physicochemical properties, and the structure-activity correlations in gaseous fluids.
Materials such as SBA-15, MCM-41 and KIT-6 have proven to be versatile platforms due to their high surface area, structural ordering, and possibility of chemical modification through covalent grafting techniques, co-condensation, or incorporation of metallic and supramolecular species [10,36,37], as well as hybrid structures such as mesoporous carbons, MOFs, and silicas doped with heteroelements. These modifications enable precise tuning of host–guest interactions and the regulation of mass transport mechanisms according to system-specific requirements.
While the discussion here focuses on SBA-15 due to its widespread use as a model material, the described strategies are equally applicable to other mesoporous silica-based materials such as MCM-41, KIT-6, or disordered systems like MSU-X [4,25,38]. The choice of method critically depends on: (i) the active site density, (ii) the pore accessibility, and (iii) the thermal and chemical stability of the resulting material [38]. The main strategies for the modification of mesoporous silicas such as SBA-15 and their impact on textural and functional properties are described below.
Post-synthesis grafting is a two-step method that involves the covalent attachment of organosilanes such as 3-aminopropyl)triethoxysilane (APTES) to surface silanol groups under strict anhydrous conditions (reflux in toluene at 110 °C under a N2 atmosphere). The reaction proceeds by nucleophilic substitution: S i O H + E t O 3 S i R S i O S I R + 3 E t O H .
This procedure preserves the hexagonal structure and enables control over the functional group density (0.5–3.0 groups/nm2), although its typically leads to a ~25% reduction in pore diameter and surface area and limited thermal stability (~250 °C) due to bond degradation [19,39]. However, it can present heterogeneous distribution in narrow pores (<4 nm) and limits its thermal stability (~250 °C) without altering the porous architecture of SBA-15 [36,40]. This method is suitable for situations requiring controlled surface modifications, such as CO2 capture.
An effective alternative is direct co-condensation, where functional precursors are simultaneously incorporated with TEOS and functional silanes such as 3-mercaptopropyltrimethoxysilane (MPTMS) in the presence of the Pluronic® P123 surfactant during the synthesis of the mesoporous material. This method has demonstrated a more homogeneous integration of active groups such as –SH or –NH2 (approximately 1.2–3.5 groups/nm2) within the silica backbone, although it typically results in a pore size reduction of approximately 25% of the BET area (1–2 nm) and presents lower thermal stability (<200 °C) [10,25,38].
Methods such as impregnation with metal nitrates (Cu2+, Fe3+, Ni2+) and subsequent calcination generate active sites that allow efficient catalytic processes, such as CH4 oxidation or Fenton-type reactions, achieving conversion rates exceeding 90% in systems such as Cu/SBA-15 [41,42,43]. Additionally, the isomorphic substitution of Si4+ by Al3+ in the network generates Brønsted acid sites without compromising the integrity of the SBA-15 framework, improving its activity in acid catalysis and selective adsorption processes [18,44]. This method does not significantly alter the Brunauer–Emmett–Teller (BET) area (–10%) and offers thermal stability (~600 °C) [45], ideal for catalytic systems that require noble metals (Pt, Pd) under severe conditions.
More recently, modification through supramolecular interactions has gained relevance, where the density of groups varies depending on the host, but reduces the BET area by 20% and has low thermal stability (<200 °C) [3]. This allows adding guest molecules such as β-cyclodextrin (β-CD), capable of forming inclusion complexes with organic contaminants, heavy metals or pesticides. Hybrid systems such as SBA-15 functionalized with β-CD have been shown to improve adsorption, as is the case with methylene blue where it reached up to 1791 (mg g−1), thanks to host–guest interactions and electrostatic bonds between amino groups and sulfonated groups, through amide linkages or covalent grafting, which improves the removal of metal ions such as Cd2+, Hg2+ or Pb2+, with efficiencies greater than 90% [37]. Furthermore, their inclusion in hybrid structures such as MOFs or PVDF membranes allow adsorbents to be recovered for up to five cycles with minimal capacity loss [36,37].
Emerging methods, such as the anchoring of β-cyclodextrin or azobenzene onto the surface of SBA-15 through photoactive modification, allow creating materials capable of responding to external stimuli such as light, pH, or redox conditions. These materials exhibit advanced properties such as high CO2/N2 selectivity (>200) or photoactivated diffusion control [3,40]. Therefore, they are used in the selective adsorption of pollutants through host–guest interactions.
Functionalization typically induces changes in textural properties, including a decrease in BET surface area and pore size; however, these effects are often offset by improved adsorption capacity, selectivity, and surface reactivity [18,36]. SBA-15 modified with β-CD and amines showed synergy in the capture of aromatic pollutants, validated both experimentally and through DFT calculations and two-dimensional nuclear magnetic resonance (NMR) spectroscopy [36].
The field of mesoporous materials has evolved toward a diversification in both the base structures and the functional groups introduced. This versatility has enabled applications across new areas, such as fine catalysis, selective environmental remediation, magnetic separation, and targeted drug release.
Figure 2 illustrates representative surface modification strategies and their functional outcomes: –NH2 (aminopropyl, APTES) grafting introduces amino groups that enhance CO2 uptake (1.2–2.8 mmol g−1) and heavy-metal binding; post-synthesis oxidation yields carboxyl-functionalized SBA-15, improving dye and metal adsorption by increasing surface polarity; incorporation of magnetic oxides (Fe3O4) enables magnetic recovery and reuse in liquid-phase separations; Schiff-base formation produces bidentate chelating sites with high Pb(II) selectivity (qmax ≈ 150 mg g−1); and introduction of hydrophobic organic groups (alkyls, phenyls) modulates surface wettability for selective adsorption of volatile organics and drug molecules.
Table 3 compiles examples of functionalized mesoporous systems, detailing support structures, chemical modifications, synthetic approaches, processing conditions, and primary applications and highlighting not only the support structure (SBA-15, MCM-41, KIT-6, CMK-3, MOFs) but also the type of chemical modification, the synthetic method used, typical operating conditions, and the associated main application. Functionalized SBA-15 consistently demonstrates enhanced CO2 adsorption capacities (qa5s ≥ 2.8 mmol g−1 at 1 bar, 25 °C) and improved acid catalysis performance (TOF: 15–40 h−1) compared to non-functionalized analogues [40,46].
This summary highlights the critical role of rational support–modifier pairing and incorporation protocols in determining material efficiency, while also reflecting a clear trend toward multifunctional and hybrid materials with enhanced reusability and structural integrity under demanding conditions.
The choice of functional group depends on the application objective. Amino groups (–NH2) favor interactions with anionic species, heavy metals and allow the immobilization of biomolecules [12,17]. Thiol groups (–SH) have an affinity for noble metals (Au, Ag, Pt) and are used in catalysis and selective detection [17,45]. Phosphonic groups (–PO3H2) improve thermal stability and introduce acid sites for acid catalysis [3,44]. And organic functional groups (phenyls, alkyls, carboxyls) adjust hydrophobicity and polarity and allow the anchoring of specific drugs or catalysts [11,51,52]. These routes allow the SBA-15 to surface to be tailored for specific applications such as metal capture, selective adsorption and targeted drug release [7,18,38].
Each route modifies the physicochemical environment of SBA-15, directly influencing textural parameters (ΔBET = –10 to –30%) and mass-transfer behavior, thus linking surface chemistry to performance in adsorption, catalysis, and separation processes.
Beyond the qualitative overview presented in Figure 2, the quantitative influence of each functionalization route on the texture, stability, and functionality of SBA-15 is summarized in Table 4. The data were compiled from high-impact studies by Pérez-Page et al., [21], Zhu et al., [51], Dudarko [53], Larki et al., [54] and Linares et al., [55].
These data show that intermediate surface coverages (≈1–2 nm−2) provide the best compromise between adsorption affinity and molecular transport, preserving at least 70% of the original BET surface area while ensuring adequate active-site density for selective interactions. For instance, APTES-grafted SBA-15 typically exhibits a surface area decrease from 830 → 640 m2 g−1 (−23%) and a CO2 uptake of 1.2–2.8 mmol g−1 at 25 °C [54].
Metal-impregnated SBA-15 retains its structural order with conversion efficiencies exceeding 90% in oxidation and hydrogenation reactions. In metal-modified systems, the long-term activity depends on nanoparticle dispersion and resistance to sintering under thermal cycling; isomorphic substitution (Si4+ → Al3+) enhances surface acidity while preserving diffusion pathways up to 600 °C [50].
Competitive multi-ion adsorption studies on NH2-SBA-15 confirm that Pb2+ and Cr3+ exhibit the highest selectivity, while the efficiency for Ni2+ and Co2+ drops to ~40% due to site competition [56].

4. Diffusion Mechanism of Substances in Mesopores: General Model Applied to SBA-15

Mass transport in ordered mesoporous materials is governed by a complex interplay of pore geometry, surface interactions, and multiple concurrent diffusion mechanisms operating under various thermodynamic conditions. This process involves coupled mechanisms of volumetric diffusion, surface diffusion, and adsorption–desorption phenomena operating at multiple spatial and temporal scales. Understanding these transport mechanisms is crucial for optimizing material performance [11,57,58].
Mathematical modeling of transport in these media is generally approached using continuous approximations based on partial differential equations (PDEs), considering pore geometry, local crystal structure, surface heterogeneity, and molecular migration and retention mechanisms. These models allow estimating operating parameters such as the effective diffusion coefficient, adsorption rate, and concentration profile, which are difficult to measure directly in complex porous media [59]. Molecular transport in SBA-15 is commonly modeled by partial differential equations that describe the temporal and spatial evolution of solute concentration [60,61].

4.1. Modeling Diffusion and Transport Mechanisms

Mass transport in ordered mesoporous materials, such as SBA-15, is a multifactorial process involving complex interactions between pore geometry, surface roughness, chemical functionalization, and the surrounding thermodynamic conditions. This behavior involves coupled volumetric and surface diffusion along with reversible adsorption phenomena, which operate at multiple spatial and temporal scales [59].
Under ideal conditions (isothermal regime (298–323 K), low pressure (P < 1 atm), and absence of convective flow), mass transport can be adequately described by Fick’s law (Equation (1)). This is valid when internal surfaces do not present active functionalization, and homogeneous concentration gradients are maintained [32]. Under these conditions, diffusion predominantly proceeds via molecular mechanisms, which allows the use of simplified one-dimensional models to estimate the effective diffusion coefficient within the mesoporous framework [57,61]. However, characterization studies using experimental techniques (NMR, X-ray scattering) and molecular simulations [62] have reported cases of anomalous or non-Fickian diffusion when the mesoporous material presents complex microstructures, irregular connectivities or molecular confinement, or interfacial forces [63]. This behavior is often attributed to factors such as: (i) irregular micropore connectivity, which induces tortuous paths and prolonged molecular residence times, and (ii) confinement effects in sub-10 nm pores, where wall collisions dominate over intermolecular collisions, favoring Knudsen type transport. Molecular confinement in pores <10 nm, molecule wall collisions dominate over intermolecular collisions, activating Knudsen mechanisms [64].
SBA-15 not only possesses primary mesopores, but also micropores within the walls that interconnect the main channels. The presence and relative proportion of these micropores can change the dominant mechanism: from a zeolitic type diffusion in micropores to a faster one in secondary mesopores [65]. Moreover, the interaction energy between adsorbate and surface significantly influences local concentration gradients and overall transport dynamics [66].
The introduction of functional groups significantly alters the mass transport mechanisms in mesoporous materials. Functionalization can either enhance diffusion or promote selective sorption, depending on the nature of the chemical group and the type of interaction with the sorbate. Adding functional groups into SBA-15 modifies the energy interaction between sorbate molecules and pore walls, thereby influencing diffusion mechanisms [20]. For instance, amine groups (–NH2) increase the adsorption energy of CO2 via carbamate formation ( E a d s ≈ 50–80 kJ mol−1), which reduces the effective diffusivity ( D e f f ) while enhancing adsorption capacity [19]. Thiol groups (–SH) promote heavy metal (Pb2+) adsorption, with surface diffusion coefficients ( D s ) governed by a thermally activated hopping model [35]. Sulfonate groups (–SO3H), on the other hand, introduce surface charges that modulate ionic diffusion through electrostatic interactions [12].

4.1.1. Molecular Diffusion (Fickian)

In pores larger than 10 nm and under low-pressure conditions (P < 1 atm), mass transport follows Fick’s law, which predominates when the pore diameter ( d p ) is significantly greater than the molecular mean free path (λ) for λ d p (pore diameter between 0.5 and 2 nm versus pores of 5 to 10 nm). In this regime, collisions between diffusing molecules are dominant over interactions with the pore walls, and the mass flux is described by:
j = D m C
When modeling transport within a cylindrical pore such as the pure SBA-15 channels (with pores > 10 nm and at moderate pressures), the diffusion equation must be formulated in cylindrical coordinates ( r , , z ). In cases with axial symmetry and no angular or axial dependence, the one-dimensional Fick is [57,61]:
Volumetric transport equation [57]:
C ( r , t ) d t = D e f f 2 C r 2 + 1 r C r

4.1.2. Knudsen Diffusion

When the mean free path is comparable to or greater than the pore diameter is λ d p . E In this regime, collisions between molecules are rare and most collisions occur against the channel wall, generating more restricted transport [64]. This regime is relevant for CO2 capture at low pressures [20,67]. The Knudsen coefficient D k is defined as:
D k = d p 3 8 R T π M

4.1.3. Combined Diffusion: Bosanquet Model

When both regimes coexist, which is common in functionalized SBA-15 operating at ambient conditions, the transport can be described by a Bosanquet resistance model (4). Transport in the channel volume is dominated by two mechanisms acting in parallel: Knudsen diffusion ( D k ) and free-phase molecular diffusion ( D m ) [57]. The overall effective diffusion coefficient ( D v ) is described as:
D v = 1 D k + 1 D m 1
Molecular diffusion:
D m = 1 3 λ v ¯ ,     v ¯ = 8 R T π M

4.1.4. Surface Diffusion

Surface diffusion occurs when adsorbed molecules migrate laterally between active sites along the internal pore surface, typically via van der Waals or hydrogen bonding [60].
D s = Γ a 2 e E a R T
This mechanism is particularly relevant in SBA-15 functionalized with –NH2, –SH or transition metal groups [3,48].

4.1.5. Multicomponent Diffusion: Maxwell–Stefan Model

In multicomponent systems, Fick’s law is often insufficient due to the presence of molecular interactions. Instead, the Maxwell–Stefan framework accounts for these interactions via chemical potential gradients [36,57,68]:
q i R T μ i = j i θ j j i θ i j j D i j + j i D o i
In the case of CO2 capture in SBA-15 functionalized with amino groups (–NH2), it has been shown that the formation of carbamates modifies the transport mechanism, reducing the effective diffusivity due to chemical retention processes at the surface. Therefore, the Maxwell-Stefan model (Equation (7)) generalizes transport by considering chemical potential gradients and intermolecular interactions ( D i j ). According to Kärger and Ruthven, the D i j coefficients can be estimated using the Vignes correlation (Equation (8)), linking the diffusivities of pure components [57]. This correlation predicts a decrease in D e f f due to the formation of stable surface complexes such as carbamates, consistent with experimental findings [38].
In complex porous systems such as functionalized SBA-15, molecular transport is governed by multiple mechanisms operating at distinct spatial scales. To describe this behavior, models that consider both molecular interactions and the structural constraints of the material are combined. Two key equations are the Vignes correlation and the coupling model for hierarchical systems (systems that combine multi-scale pores micro <2 nm, meso 2–50 nm, and macropores >50 nm into a controlled architecture [18]):
  • Vignes correlation [69,70]:
1 D = 1 D o i + 1 D i j
Coupling model for hierarchical systems (Equation (9)) where global transport integrates interfacial barriers [57].
1 D g l o b a l = 1 D m i c r o + 1 α l
In the simple case of equally spaced transport resistances (the spacing l is assumed to be small compared to the particle length L) and permeability equal to α, in SBA-15 with interconnecting micropores, D m i c r o limits overall transport if α l is large (accessible principal pores) [55].
These transport equations are conceptually analogous to electrical resistance in series, where the inverse of the diffusivity (1/D) represents a diffusive resistance. Equation (8) considers competitive resistances between molecules, while Equation (9) integrates scalar resistances (micro → macro), while the hierarchical model requires porometric characterization (N2 adsorption) to determine l and α [55].

4.1.6. Adsorption–Desorption: Kinetic Mechanism

The presence of functional groups on the walls of SBA-15 introduces nonlinear behavior into the transport dynamics due to reversible adsorption and desorption processes. The kinetics of reversible adsorption/desorption at the surface are described by the following rate expressions:
Adsorption :   r a = k a C v 1 θ
Desorption :   r d = k d θ
A two-phase model couples pore volume diffusion in the pore volume ( J ), surface adsorption ( r a ) and desorción ( r d ) kinetics as follows:
Suschem 06 00042 i001
Suschem 06 00042 i002
One application of the two-phase model is the design of adsorbents for CO2 capture under post-combustion conditions [18,19]. For APTES-functionalized SBA-15, the equations predict the volumetric CO2 concentration ( C v ), to decay exponentially along the pore according to 2 C v k a C v / D v (adsorption-controlled domain), while the surface coverage (θ) reaches 80% in less than 10 s at 50 °C, according to in situ FTIR spectroscopy data The combination of high mesopore transport rates D v ~ 10 7 m 2 s 1 and irreversible adsorption kinetics ( k d ~ 10 5 s 1 ) accounts for the high CO2 capture capacity (2.8 mmol g−1 at 1 bar) observed in APTES-functionalized SBA-15 [10]. These results, supported by in situ FTIR spectroscopy, demonstrate how coupled equations for Cv and θ enable performance predictions under real-world conditions, such as in post-combustion capture or purification systems [10,18,47].

4.1.7. Contribution of Nonpolar van der Waals Forces

In functionalized mesoporous materials, both physical and chemical adsorption mechanisms may coexist, depending on the nature of the adsorbate. This duality has been well documented, particularly in CO2 adsorption on SBA-15–NH2, where chemisorption dominates. The increased residence time of molecules on the functionalized surface results in a reduced effective diffusion coefficient D e f f [20,52,60].
In contrast, the adsorption of nonpolar gases such as N2 and CH4 is primarily governed by van der Waals forces. These interactions are physical in nature and depend on the polarizability of the molecules and the adsorbate–adsorbent distance.
Thus, incorporating van der Waals interactions is essential in diffusion models, particularly for systems involving apolar gases and where no chemical bonding occurs. Although these interactions are weaker than covalent or ionic bonds, they can dominate adsorption behavior under ideal conditions, significantly affecting gas separation or volatile compound storage processes. These interactions, although weaker than chemical ones, can dominate adsorption behavior under ideal conditions, when working with apolar gases. Their correct inclusion in the models allows to improve the predictability of diffusion and to adjust the design of materials for specific processes such as gas separation or storage of volatile compounds. Equation (14) provides a simplified representation of the Lennard-Jones potential [59,62,71], which describes non-covalent interactions between neutral molecules. In this model, the adsorption energy ( E a ) is directly proportional to the van der Waals interaction constant and inversely proportional to the sixth power of the intermolecular distance ( l p p ).
E a d s A l p p 6
This modeling approach is applicable to the design of adsorbents, heterogeneous catalysts, and controlled release systems, where the prediction of weak adsorption phenomena is essential.

4.2. Effect of Functionalization on Diffusion and Transport Mechanisms

Mesoporous Understanding diffusion and transport in mesoporous materials such as SBA-15 is essential to interpret their performance in catalysis, gas separation, energy storage, and controlled drug delivery. The functionalization of mesoporous silicas not only modifies their surface chemistry but also reshapes the internal diffusion landscape. As discussed in the preceding sections, strategies such as grafting, co-condensation, metal impregnation, and supramolecular anchoring introduce functional groups that alter pore size, connectivity, and surface energy. These structural and chemical modifications directly influence the effective diffusivity ( D e f f   ) and mass-transfer resistance within the mesoporous framework, particularly in SBA-15 with pore diameters ranging from 6 to 10 nm [9,55].
From a mechanistic standpoint, diffusion in functionalized SBA-15 results from a dynamic balance between adsorptive affinity and pore accessibility. High surface coverages enhance gas–solid interaction strength and selectivity typically following Langmuir or Freundlich adsorption isotherms but simultaneously increase tortuosity and reduce the accessible cross-sectional area for molecular transport. Consequently, D e f f   decreases as pore channels become partially blocked or energetically heterogeneous [54].
Recent molecular dynamics and Monte Carlo simulations have provided quantitative insight into CO2 adsorption mechanisms in amine-functionalized SBA-15, confirming the formation of chemisorbed carbamate species with adsorption energies near 40 kJ mol−1 and diffusion coefficients in the 10−9–10−10 m2 s−1 range [72].
Intermediate functional group densities (1–2 nm−2) generally provide the most favorable compromise between selectivity and transport efficiency, maintaining sufficient active sites while preserving 70–80% of the original BET surface area. This regime supports effective Van der Waals interactions for CO2, CH4, and volatile organic compounds while minimizing diffusion constraints within SBA-15’s ordered hexagonal network [55,73,74].
Among the various diffusion-controlled processes, CO2 capture offers a representative case study linking surface chemistry to molecular transport in functionalized SBA-15. The selective adsorption of CO2 on amine-modified SBA-15 occurs mainly through acid–base chemisorption mechanisms, forming carbamate or bicarbonate species according to reactions such as CO2 + 2 RNH2 → RNHCOO + RNH3+. The uniform mesoporous channels enhance accessibility to amine sites, while diffusion resistance within the pores governs overall uptake kinetics. Hydrophilicity and surface polarity further modulate the balance between physisorption and chemisorption, contributing to the high selectivity observed in SBA-15-based adsorbents [9,54,75,76].

5. Challenges and Future Perspectives

Mesoporous materials such as SBA-15 have proven effective in catalysis, gas separation, energy storage, and controlled drug delivery. Nonetheless, their development faces several challenges, particularly in engineering hierarchical pore architectures (micro–mesoporous, branched, or spherical morphologies) that enhance mass transport and active site accessibility. Achieving such architecture typically requires stringent synthesis conditions, including precise control of pH, temperature, and surfactant concentration. Similarly, functionalization with specific groups requires additional synthesis steps and costly reagents, which may hinder scalability [54,77].
In the energy sector, mesoporous materials are being actively investigated as components in solar cells, batteries, thermoelectric devices, and photocatalysts. According to Linares et al. [55], mesoporosity plays a critical role in clean energy technologies by improving the dispersion of the active phase and increasing surface area, which together enhance energy conversion efficiency.
In environmental catalysis and remediation, functionalized SBA-15 has shown high efficiency for the removal of heavy metals and organic pollutants. Tailoring the surface chemistry of SBA-15 increases selectivity toward specific analytes and improves regeneration capability in adsorption desorption cycles [54,78].
In biomedical applications represent an interesting frontier for mesoporous structures. The potential of SBA-15 and its functionalized derivatives for drug delivery, optical sensing, and bone regeneration scaffolds have been demonstrated, owing to their biocompatibility and ordered mesostructure [16,21].
Future developments are expected to focus on the hybridization of SBA-15 with active phases such as metal nanoparticles, conductive polymers, or photocatalysts. These hybrid materials aim to address current challenges in energy storage, contaminant detection, and selective gas adsorption [54,55].
A limitation of functionalized SBA-15 is the heterogeneous distribution of surface groups produced by post-synthesis grafting, which compromises reproducibility and reduces transport efficiency particularly in narrow pores (<4 nm). Moreover, some functional linkages are thermally or chemically labile, leading to decreased long term stability [54].
Regarding mass transport modeling, classical approaches such as Fickian or Knudsen diffusion are often inadequate for describing molecular transport in functionalized mesoporous materials because they neglect sorbate–sorbent interactions and multicomponent coupling effects. In contrast, the Maxwell–Stefan formalism provides a more comprehensive and realistic framework by integrating surface diffusion and pore confinement phenomena, thereby linking molecular-scale adsorption processes with continuum transport behavior [78]. Nevertheless, the practical application of this model remains challenging due to the complexity of accurately determining its key parameters. The cross-diffusion coefficients ( D i j ), which describe interactions between co-adsorbed species such as CO2 and NH3 within amine-functionalized SBA-15, require detailed in situ spectroscopic or calorimetric measurements [57]. The adsorption energies ( E a d s ) vary depending on the functional groups (–NH2 vs. -SO3H) and the heterogeneous distribution of active sites [19,39,79], while the surface coverage fraction (θ), dependent on functional group density and residual humidity, remains difficult to quantify in nanometric pores.
The mechanistic understanding derived from diffusion models directly reflects real applications. For instance, pollutant adsorption on functionalized SBA-15 is driven by electrostatic interactions between charged surface groups (–NH3+, –COO) and ionic contaminants, π–π stacking with aromatic molecules, and surface complexation with metal ions such as Pb2+ or Cd2+ [51,54].
Similarly, drug loading generally occurs through diffusion or capillary impregnation within mesopores, while release kinetics depend on pore size, surface polarity, and functional-group chemistry. Amino- and carboxyl-modified SBA-15 exhibit sustained-release behavior due to hydrogen bonding and electrostatic interactions between drug molecules and silanol groups. The resulting profiles typically follow Fickian or anomalous diffusion, enabling tunable release rates for antibiotics, anti-inflammatory agents, and wound-healing drugs [4].
Experimental validation of diffusion models remains challenging, primarily due to the difficulty of directly measuring effective diffusivities. These parameters are essential for predicting selectivity in gas separations and kinetics in drug release applications. However, accurate quantification often requires advanced techniques such as Pulsed Field Gradient Nuclear Magnetic Resonance (PFG-NMR) or Quartz Crystal Microbalance with Dissipation monitoring (QCM-D), which, despite their value, are limited by cost and availability [33].

6. Conclusions

This SBA-15 mesoporous silica represents a highly versatile and tunable platform with diverse applications in catalysis, gas separation, chemical sensing, and biomedicine. Its well-ordered mesostructure, high surface area (>500 m2 g−1), and controllable pore size (5–30 nm) make it particularly suitable for processes where mass transport is critical. However, alternative mesoporous materials such as KIT-6 (3D cubic structure) and MCM-41 (smaller 2–4 nm pores) offer complementary advantages depending on specific application requirements. While these systems share common functionalization approaches (grafting, co-condensation), material selection should be guided by key parameters including diffusion pathway geometry (unidirectional versus three-dimensional transport), molecular dimensions of target species, and required thermal/chemical stability thresholds.
Strategic surface functionalization of SBA-15 with chemical moieties such as amine, thiol, sulfonate, or organometallic groups has significantly enhanced its selectivity and adsorption capacity for target species such as CO2, heavy metal ions, and persistent organic pollutants. Nevertheless, several technological challenges remain, including: (1) non-uniform distribution of active sites, (2) susceptibility to degradation under extreme pH or thermal condition, and (3) difficulty in preserving structural integrity in heavily modified or hybrid configurations. Overcoming these limitations will require advanced synthesis protocols and robust characterization techniques to unlock the full potential of functionalized mesoporous systems.
While classical transport models (Fick’s Law, Knudsen diffusion) remain useful for simplified systems, they often fail to capture the nonlinear sorbate–sorbent interactions and multicomponent effects present in functionalized structures. More sophisticated approaches such as the Maxwell–Stefan model offer a rigorous alternative, yet their implementation is constrained by the need for detailed input parameters (cross-diffusivities, adsorption energies), which are difficult to obtain experimentally.
Future advances are expected in the development of hierarchical mesoporous materials through the addition of hybrid phases such as metallic nanoparticles or conductive polymers and multiscale computational methods with experimental validation. These approaches will allow the design of functional supports with tailored properties. These innovations aim to overcome the limitations of single-porosity systems by integrating macropores for enhanced molecular transport, mesopores for high adsorption capacity, and micropores for improved selectivity.
Finally, the integration of SBA-15-based materials into practical systems such as adsorption beds, membranes, or microfluidic platforms represents a critical step toward their scalable and sustainable industrial deployment.

7. Further Steps

Ongoing research on functionalized mesoporous materials like SBA-15 is paving the way for a new generation of hierarchical hybrid architectures capable of integrating multiple functionalities within a single platform. By engineering synergistic micro–mesoporous networks with precisely controlled geometries (spherical, lamellar, branched), researchers aim to optimize adsorption efficiency and active site accessibility under real operating conditions. Particularly promising is the strategic incorporation of embedded functional phases, such as metallic nanoparticles, stimuli-responsive polymers, MOF structures, and biofunctional matrices, to tailor materials for specific applications.
Future breakthroughs are anticipated in the development of advanced mathematical models to describe diffusion and reaction phenomena in mesoporous systems. The integration of artificial intelligence (AI) and machine learning (ML) is expected to transform predictive modeling, particularly in combination with multiscale Maxwell–Stefan formulations that include adsorption–desorption kinetics. These models will be validated through state-of-the-art techniques such as PFG-NMR and QENS, enabling precise in situ characterization of transport dynamics.
This combined computational–experimental approach will drive the rational design of next-generation functionalized materials for critical applications. Key focus areas include selective separation processes for valuable compounds (CO2 capture, rare-earth element recovery), targeted delivery systems with stimuli-responsive release mechanisms, and high-density storage solutions for energy carriers (H2, CH4) under extreme environmental conditions. The goal is to develop smart, multifunctional materials that push the boundaries of what’s possible in fields ranging from sustainable energy to precision medicine.

Author Contributions

Conceptualization, M.M. and M.R.; methodology, M.M. and G.M.; validation, D.F., R.N. and A.M.; formal analysis, M.R. and D.F.; investigation, M.M.; data curation, A.M.; writing—original draft preparation, M.M.; writing—review and editing, M.R. and R.N.; visualization, M.M.; supervision, A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This study did not generate any new data.

Acknowledgments

The authors thank the Process Engineering Research Group (GIIP) of Central University of Ecuador and the University of Alicante for their support. The authors have reviewed and edited the output and take full responsibility for the content of this publication.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

The following abbreviations are used in this manuscript:
SymbolDescriptionUnits
a Distance between active sites on the surfacem
A Attractive interaction constant between molecules or between a molecule and the surface of the materialJ m6
C Solute concentration in general (may vary across spatial domains)mol m−3
C ( r , t ) Concentration at a radial distance r and time tmol m−3
C v Solute concentration in the gas phase inside the porous channelmol m−3
C o Initial solute concentrationmol m−3
d p Pore diameterm
D e f f Effective diffusion coefficient in the porous mediumm2 s−1
D v Diffusion coefficient in the channel volume (combined molecular and Knudsen)m2 s−1
D m Molecular diffusion coefficient in the free phase (free motion between molecular collisions)m2 s−1
D K Knudsen diffusion coefficient (dominated by collisions with pore walls)m2 s−1
D s Surface diffusion coefficient (adsorbate migrating along the pore wall)m2 s−1
D i j Cross-diffusivity coefficient between species i and jm2 s−1
D o i Frictional diffusivity coefficient between species i and the porous solid matrixm2 s−1
D g l o b a l Overall effective diffusion coefficient in hierarchical materialsm2 s−1
E a Activation energy for surface migration (~5–30 kJ mol−1 for Van der Waals)kJ mol−1
E a d s Adsorption energykJ mol−1
J Total molar flux vectormol m−2 s−1
k a Adsorption rate constants−1
k d Desorption rate constants−1
l Interfacial spacing between porous domains (used in hierarchical transport)m
l p p Distance between two particles or between an adsorbed molecule and the material surfacem
M Molar mass of the solutekg mol−1
q i Net charge or transport factor of species idimensionless
r Radial distance from the pore axism
R Universal gas constant (8.314)J mol−1 K−1
T Absolute temperatureK
v ¯ Average molecular velocity of the solutem s−1
α Interfacial permeability coefficient in hierarchical transport modelsm s−1
λ Mean free path of the moleculem
ε Porosity of the mediumdimensionless
τ Tortuosity of the diffusive pathdimensionless
Γ Hopping frequency of adsorbate between sites (~1012 s−1)s−1
μ i Chemical potential of idimensionless
θ Surface coverage fraction (occupied sites over total available)dimensionless
θ j Mole fraction of species jdimensionless
2 Laplacian operator (gradient of the gradient)m−2
C Gradient of concentration fieldmol m−4
μ i Gradient of chemical potential of species iJ mol−1 m−1
/ t Partial derivative related to times−1
2 / r 2 Second partial derivative related to the radial coordinatem−2

References

  1. Everett, D.H. Manual of Symbols and Terminology for Physicochemical Quantities and Units, Appendix II: Definitions, Terminology and Symbols in Colloid and Surface Chemistry. Pure Appl. Chem. 1972, 31, 577–638. [Google Scholar] [CrossRef]
  2. Kresge, C.T.; Leonowicz, M.E.; Roth, W.J.; Vartuli, J.C.; Beck, J.S. Ordered mesoporous molecular sieves synthesized by a liquid-crystal template mechanism. Nature 1992, 359, 710–712. [Google Scholar] [CrossRef]
  3. Wan, Y.; Zhao, D. On the controllable soft-templating approach to mesoporous silicates. Chem. Rev. 2007, 107, 2821–2860. [Google Scholar] [CrossRef] [PubMed]
  4. Asefa, T.; Dubovoy, V. 9.09—Ordered Mesoporous/Nanoporous Inorganic Materials via Self-Assembly. In Comprehensive Supramolecular Chemistry II; Atwood, J.L., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; Volume 9, pp. 157–192. [Google Scholar] [CrossRef]
  5. Shakeri, M.; Shal, Z.K.; Van Der Voort, P. An overview of the challenges and progress of synthesis, characterization and applications of plugged sba-15 materials for heterogeneous catalysis. Materials 2021, 14, 5082. [Google Scholar] [CrossRef] [PubMed]
  6. Verma, P.; Kuwahara, Y.; Mori, K.; Raja, R.; Yamashita, H. Functionalized mesoporous SBA-15 silica: Recent trends and catalytic applications. Nanoscale 2020, 12, 11333–11363. [Google Scholar] [CrossRef]
  7. Yuan, S.; Wang, M.; Liu, J.; Guo, B. Recent advances of SBA-15-based composites as the heterogeneous catalysts in water decontamination: A mini-review. J. Environ. Manag. 2020, 254, 109787. [Google Scholar] [CrossRef]
  8. Janus, R.; Wądrzyk, M.; Lewandowski, M.; Natkański, P.; Łątka, P.; Kuśtrowski, P. Understanding porous structure of SBA-15 upon pseudomorphic transformation into MCM-41: Non-direct investigation by carbon replication. J. Ind. Eng. Chem. 2020, 92, 131–144. [Google Scholar] [CrossRef]
  9. Asensio, J.; Beltrán, M.I.; Juárez-Serrano, N.; Berenguer, D.; Marcilla, A. Study of the Decomposition of N-Nitrosonornicotine (NNN) under Inert and Oxidative Atmospheres: Effect of the Addition of SBA-15 and MCM-41. Appl. Sci. 2022, 12, 9426. [Google Scholar] [CrossRef]
  10. Sayari, A.; Belmabkhout, Y. Stabilization of Amine-Containing CO2 Adsorbents: Dramatic Effect of Water Vapor. J. Am. Chem. Soc. 2010, 132, 6312–6314. [Google Scholar] [CrossRef]
  11. Vallet-Regí, M.; Colilla, M.; Izquierdo-Barba, I.; Manzano, M. Mesoporous silica nanoparticles for drug delivery: Current insights. Molecules 2018, 23, 47. [Google Scholar] [CrossRef]
  12. Ezzeddine, Z.; Batonneau-Gener, I.; Ghssein, G.; Pouilloux, Y. Recent Advances in Heavy Metal Adsorption via Organically Modified Mesoporous Silica: A Review. Water 2025, 17, 669. [Google Scholar] [CrossRef]
  13. Diagboya, P.N.; Dikio, E.D. Silica-based mesoporous materials; emerging designer adsorbents for aqueous pollutants removal and water treatment. Microporous Mesoporous Mater. 2018, 266, 252–267. [Google Scholar] [CrossRef]
  14. Galarneau, A.; Cambon, H.; Di Renzo, F.; Fajula, F. True microporosity and surface area of mesoporous SBA-15 silicas as a function of synthesis temperature. Langmuir 2001, 17, 8328–8335. [Google Scholar] [CrossRef]
  15. Juárez-Serrano, N.; Asensio, J.; Blasco, I.; Beltrán, M.; Marcilla, A. Effect of Time, Temperature and Stirring Rate Used in the First Step of the Synthesis of SBA-15 on Its Application as Reductor of Tars in Tobacco Smoke. Catalysts 2021, 11, 375. [Google Scholar] [CrossRef]
  16. Grini, M.I.; Benbayer, C.; Saidi-Besbes, S.; Elaissari, A. Advances in mesoporous silica nanoparticles as carriers for drug delivery and other biomedical applications. Microporous Mesoporous Mater. 2025, 391, 113603. [Google Scholar] [CrossRef]
  17. Tang, F.; Li, L.; Chen, D. Mesoporous silica nanoparticles: Synthesis, biocompatibility and drug delivery. Adv. Mater. 2012, 24, 1504–1534. [Google Scholar] [CrossRef] [PubMed]
  18. Lan, K.; Zhao, D. Functional Ordered Mesoporous Materials: Present and Future. Nano Lett. 2022, 22, 3177–3179. [Google Scholar] [CrossRef]
  19. Wang, L.; Yang, R.T. Increasing selective CO2 adsorption on amine-grafted SBA-15 by increasing silanol density. J. Phys. Chem. C 2011, 115, 21264–21272. [Google Scholar] [CrossRef]
  20. Rahmat, N.; Abdullah, A.Z.; Mohamed, A.R. A review: Mesoporous Santa Barbara amorphous-15, types, synthesis and its applications towards biorefinery production. Am. J. Appl. Sci. 2010, 7, 1579–1586. [Google Scholar] [CrossRef]
  21. Pérez-Page, M.; Yu, E.; Li, J.; Rahman, M.; Dryden, D.M.; Vidu, R.; Stroeve, P. Template-based syntheses for shape controlled nanostructures. Adv. Colloid Interface Sci. 2016, 234, 51–79. [Google Scholar] [CrossRef]
  22. Zhao, D.; Huo, Q.; Feng, J.; Chmelka, B.F.; Stucky, G.D. Nonionic Triblock and Star Diblock Copolymer and Oligomeric Surfactant Syntheses of Highly Ordered, Hydrothermally Stable, Mesoporous Silica Structures. J. Am. Chem. Soc. 1998, 120, 6024–6036. [Google Scholar] [CrossRef]
  23. Juárez-Serrano, N.; Berenguer, D.; Martínez-Castellanos, I.; Blasco, I.; Beltrán, M.; Marcilla, A. Effect of reaction time and hydrothermal treatment time on the textural properties of SBA-15 synthesized using sodium silicate as a silica source and its efficiency for reducing tobacco smoke toxicity. Catalysts 2021, 11, 808. [Google Scholar] [CrossRef]
  24. Asefa, T.; MacLachlan, M.J.; Coombs, N.; Ozin, G.A. Periodic mesoporous organosilicas with organic groups inside the channel walls. Nature 1999, 402, 867–871. [Google Scholar] [CrossRef]
  25. Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G.H.; Chmelka, B.F.; Stucky, G.D. Triblock Copolymer Syntheses of Mesoporous Silica with Periodic 50 to 300 Angstrom Pores. Science 1998, 279, 548–552. [Google Scholar] [CrossRef] [PubMed]
  26. Serra, E.; Blanco, R.M.; Díaz, I. Síntesis y Caracterización de Materiales Mesoporosos Ordenados y su Aplicación como Soportes en la Inmovilización de Lipasa. An. Quím. 2008, 104, 97–103. Available online: https://dialnet.unirioja.es/servlet/articulo?codigo=2662603 (accessed on 15 April 2025).
  27. Han, W.; Jia, Y.; Xiong, G.; Yang, W. Template-free sol-gel synthesis of mesoporous materials with ZSM-5 structure walls. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 2007; Volume 165, pp. 515–518. [Google Scholar] [CrossRef]
  28. Isaacs, M.A.; Robinson, N.; Barbero, B.; Durndell, L.J.; Manayil, J.C.; Parlett, C.M.A.; D’Agostino, C.; Wilson, K.; Lee, A.F. Unravelling mass transport in hierarchically porous catalysts. J. Mater. Chem. A 2019, 7, 11814–11825. [Google Scholar] [CrossRef]
  29. Asfaha, Y.G.; Tekile, A.K.; Zewge, F. Hybrid process of electrocoagulation and electrooxidation system for wastewater treatment: A review. Clean. Eng. Technol. 2021, 4, 100261. [Google Scholar] [CrossRef]
  30. Zhao, J.; Yan, W. Microwave-assisted Inorganic Syntheses. In Modern Inorganic Synthetic Chemistry; Elsevier: Amsterdam, The Netherlands, 2011; pp. 173–195. [Google Scholar] [CrossRef]
  31. Vigón, P.V. Síntesis de Materiales Mesoporosos Compuestos Sílice/Carbono y su Empleo como Plataforma para la Fabricación de Materiales con Propiedades Avanzadas. Ph.D. Thesis, Universidad de Oviedo, Oviedo, Spain, 2013. Available online: https://dialnet.unirioja.es/servlet/tesis?codigo=82387 (accessed on 17 May 2025).
  32. Martínez, J.G. Sólidos Ordenados Desde la Nano a la Macroestructura. An. Quim. 2006, 1, 5–12, ISSN 1575-341 (Print), 2792-5250 (Electronic). Available online: https://dialnet.unirioja.es/servlet/articulo?codigo=1996641 (accessed on 17 May 2025).
  33. Williams, J.H.; Zheng, Q.; Mantle, M.D.; Sederman, A.J.; Gladden, L.F. Probing the Diffusion Mechanism of n-Alkanes in Mesoporous Confinement Using Pulsed Field Gradient NMR. J. Phys. Chem. C 2023, 127, 15326–15335. [Google Scholar] [CrossRef]
  34. Kumar, P.; Guliants, V.V. Periodic mesoporous organic–inorganic hybrid materials: Applications in membrane separations and adsorption. Microporous Mesoporous Mater. 2010, 132, 1–14. [Google Scholar] [CrossRef]
  35. Zheng, L.; Yang, Y.; Zhang, Y.; Zhu, T.; Wang, X. Functionalization of SBA-15 mesoporous silica with bis-schiff base for the selective removal of Pb(II) from water. J. Solid State Chem. 2021, 301, 122320. [Google Scholar] [CrossRef]
  36. Li, D.; Chai, K.; Yao, X.; Zhou, L.; Wu, K.; Huang, Z.; Yan, J.; Qin, X.; Wei, W.; Ji, H. β-Cyclodextrin functionalized SBA-15 via amide linkage as a super adsorbent for rapid removal of methyl blue. J. Colloid Interface Sci. 2021, 583, 100–112. [Google Scholar] [CrossRef]
  37. Chuin, L.X.; Kamaruzaman, S.; Praveena, S.M.; Yahaya, N. Recent applications of β-cyclodextrin in selective adsorption of pesticides, heavy metals, and organic pollutants from water samples: Mini review. Microchem. J. 2024, 206, 111583. [Google Scholar] [CrossRef]
  38. da Silva, F.d.C.M.; Costa, M.J.d.S.; da Silva, L.K.R.; Batista, A.M.; da Luz, G.E., Jr. Functionalization methods of SBA-15 mesoporous molecular sieve: A brief overview. SN Appl. Sci. 2019, 1, 654. [Google Scholar] [CrossRef]
  39. Shimon, D.; Chen, C.-H.; Lee, J.J.; Didas, S.A.; Sievers, C.; Jones, C.W.; Hayes, S.E. 15N Solid State NMR Spectroscopic Study of Surface Amine Groups for Carbon Capture: 3-Aminopropylsilyl Grafted to SBA-15 Mesoporous Silica. Environ. Sci. Technol. 2018, 52, 1488–1495. [Google Scholar] [CrossRef]
  40. Ojeda-López, R.; Pérez-Hermosillo, I.J.; Esparza-Schulz, J.M.; Cervantes-Uribe, A.; Domínguez-Ortiz, A. SBA-15 materials: Calcination temperature influence on textural properties and total silanol ratio. Adsorption 2015, 21, 659–669. [Google Scholar] [CrossRef]
  41. Zhang, Z.; Yin, J.; Heeres, H.J.; Melián-Cabrera, I. Thermal detemplation of SBA-15 mesophases. Effect of the activation protocol on the framework contraction. Microporous Mesoporous Mater. 2013, 176, 103–111. [Google Scholar] [CrossRef]
  42. Szegedi, Á.; Lázár, K.; Solt, H.; Popova, M. Peculiar redox properties of SBA-15 supported copper ferrite catalysts promoting total oxidation of a model volatile organic air pollutant. Surfaces Interfaces 2025, 56, 105498. [Google Scholar] [CrossRef]
  43. Gil, A.G.; Wu, Z.; Chadwick, D.; Li, K. Ni/SBA-15 Catalysts for combined steam methane reforming and water gas shift—Prepared for use in catalytic membrane reactors. Appl. Catal. A Gen. 2015, 506, 188–196. [Google Scholar] [CrossRef]
  44. Laskowski, Ł.; Laskowska, M.; Bałanda, M.; Fitta, M.; Kwiatkowska, J.; Dziliński, K.; Karczmarska, A. Mesoporous silica SBA-15 functionalized by nickel–phosphonic units: Raman and magnetic analysis. Microporous Mesoporous Mater. 2014, 200, 253–259. [Google Scholar] [CrossRef]
  45. Wu, H.; Xiao, Y.; Guo, Y.; Miao, S.; Chen, Q.; Chen, Z. Functionalization of SBA-15 mesoporous materials with 2-acetylthiophene for adsorption of Cr(III) ions. Microporous Mesoporous Mater. 2020, 292, 109754. [Google Scholar] [CrossRef]
  46. Xia, X.; Jin, Y.; Zhao, H.; Wang, G.; Huang, D. Optimization and Experiment of Hot Air Drying Process of Cyperus esculentus Seeds. Agriculture 2023, 13, 617. [Google Scholar] [CrossRef]
  47. Wang, Y.; Yang, R.T. Template Removal from SBA-15 by Ionic Liquid for Amine Grafting: Applications to CO2 Capture and Natural Gas Desulfurization. ACS Sustain. Chem. Eng. 2020, 8, 8295–8304. [Google Scholar] [CrossRef]
  48. Hoffmann, F.; Cornelius, M.; Morell, J.; Fröba, M. Silica-based mesoporous organic–inorganic hybrid materials. Angew. Chem. Int. Ed. Engl. 2006, 45, 3216–3251. [Google Scholar] [CrossRef]
  49. Taguchi, A.; Schüth, F. Ordered mesoporous materials in catalysis. Microporous Mesoporous Mater. 2005, 77, 1–45. [Google Scholar] [CrossRef]
  50. Yin, Y.; Yang, Z.-F.; Wen, Z.-H.; Yuan, A.-H.; Liu, X.-Q.; Zhang, Z.-Z.; Zhou, H. Modification of as Synthesized SBA-15 with Pt nanoparticles: Nanoconfinement Effects Give a Boost for Hydrogen Storage at Room Temperature. Sci. Rep. 2017, 7, 4509. [Google Scholar] [CrossRef]
  51. Zhu, J.; Wang, T.; Xu, X.; Xiao, P.; Li, J. Pt nanoparticles supported on SBA-15: Synthesis, characterization and applications in heterogeneous catalysis. Appl. Catal. B Environ. 2013, 130–131, 197–217. [Google Scholar] [CrossRef]
  52. Laskowski, Ł.; Laskowska, M.; Jelonkiewicz, J.; Galkowski, T.; Pawlik, P.; Piech, H.; Doskocz, M. Iron Doped SBA-15 Mesoporous Silica Studied by Mössbauer Spectroscopy. J. Nanomater. 2016, 2016, 1256851. [Google Scholar] [CrossRef]
  53. Dudarko, O.; Kobylinska, N.; Mishra, B.; Kessler, V.G.; Tripathi, B.P.; Seisenbaeva, G.A. Facile strategies for synthesis of functionalized mesoporous silicas for the removal of rare-earth elements and heavy metals from aqueous systems. Microporous Mesoporous Mater. 2021, 315, 110919. [Google Scholar] [CrossRef]
  54. Larki, A.; Saghanezhad, S.J.; Ghomi, M. Recent advances of functionalized SBA-15 in the separation/preconcentration of various analytes: A review. Microchem. J. 2021, 169, 106601. [Google Scholar] [CrossRef]
  55. Linares, N.; Silvestre-Albero, A.M.; Serrano, E.; Silvestre-Albero, J.; García-Martínez, J. Mesoporous materials for clean energy technologies. Chem. Soc. Rev. 2014, 43, 7681–7717. [Google Scholar] [CrossRef]
  56. Albayati, T.M.; Sabri, A.A.; Abed, D.B. Adsorption of binary and multi heavy metals ions from aqueous solution by amine functionalized SBA-15 mesoporous adsorbent in a batch system. Desalination Water Treat. 2019, 151, 315–321. [Google Scholar] [CrossRef]
  57. Kärger, J.; Ruthven, D.M. Diffusion in nanoporous materials: Fundamental principles, insights and challenges. New J. Chem. 2016, 40, 4027–4048. [Google Scholar] [CrossRef]
  58. Li, W.; Liu, J.; Zhao, D. Mesoporous materials for energy conversion and storage devices. Nat. Rev. Mater. 2016, 1, 16023. [Google Scholar] [CrossRef]
  59. Kärger, J.; Valiullin, R. Mass transfer in mesoporous materials: The benefit of microscopic diffusion measurement. Chem. Soc. Rev. 2013, 42, 4172–4197. [Google Scholar] [CrossRef] [PubMed]
  60. Whitaker, S. The Method of Volume Averaging; Theory and Applications of Transport in Porous Media; Springer: Dordrecht, The Netherlands, 1999; Volume 13. [Google Scholar] [CrossRef]
  61. Bird, R.B.; Earl, S.W.; Lightfoot, E.N. Fenómenos de Transporte, 2nd ed.; John Wiley & Sons: Hoboken, NJ, USA, 2010; Available online: https://books.google.com/books/about/Transport_Phenomena.html?id=L5FnNlIaGfcC (accessed on 17 May 2025).
  62. Kärger, J.; Freude, D.; Haase, J. Diffusion in nanoporous materials: Novel insights by combining MAS and PFG NMR. Processes 2018, 6, 147. [Google Scholar] [CrossRef]
  63. Zhokh, A. Size-controlled non-Fickian diffusion in a combined micro- and mesoporous material. Chem. Phys. 2019, 520, 27–31. [Google Scholar] [CrossRef]
  64. Duong, D.D. Adsorption Analysis: Equilibria and Kinetics; Imperial College Press: London, UK, 1998; Volume 1, p. 916. [Google Scholar] [CrossRef]
  65. Hoang, V.-T.; Huang, Q.; Eić, M.; Do, T.-O.; Kaliaguine, S. Structure and diffusion characterization of SBA-15 materials. Langmuir 2005, 21, 2051–2057. [Google Scholar] [CrossRef]
  66. Rusinque, H.; Brenner, G. Mass transport in porous media at the micro- and nanoscale: A novel method to model hindered diffusion. Microporous Mesoporous Mater. 2019, 280, 157–165. [Google Scholar] [CrossRef]
  67. Wang, T.; Zhang, K.; Du, S.-W.; Wu, L.-G.; Li, C.-J.; Chen, H.L.; Guo, H.-C. Transformation from Knudsen diffusion to facilitated diffusion for CO2 in confined mass transfer channels of polyimide mixed matrix membranes. Chem. Eng. J. 2023, 477, 147280. [Google Scholar] [CrossRef]
  68. Vignes, A. Diffusion in Binary Solutions Variation of Diffusion Coefficient with Composition. Ind. Eng. Chem. Fundam. 1966, 5, 189–199. [Google Scholar] [CrossRef]
  69. Geng, Z.; Zhang, Y.; Wang, L.; Gong, H. A theoretical insight into diffusion mechanism of aldol condensation of acetaldehyde in Zr-BEA zeolite. Mol. Catal. 2025, 572, 114770. [Google Scholar] [CrossRef]
  70. Israelachvili, J.N. Intermolecular and Surface Forces, 3rd ed.; Elsevier: Amsterdam, The Netherlands, 2011; pp. 1–676. [Google Scholar] [CrossRef]
  71. Subagyono, R.R.D.J.N.; Chaffee, A.L. CO2 adsorption on SBA-15: A molecular modelling. IOP Conf. Ser. Earth Environ. Sci. 2018, 144, 012045. [Google Scholar] [CrossRef]
  72. Obeso, J.L.; López-Cervantes, V.B.; Ojeda-López, R.; Yañez-Aulestia, A.; Cordero-Sánchez, S.; Sánchez-González, E.; Novelo-Peralta, O.; Morales, K.E.R.; Solís-Ibarra, D.; Ibarra, I.A.; et al. SO2 Capture and Detection by SBA-15 and Amine-Functionalized SBA-15. Ind. Eng. Chem. Res. 2024, 63, 2223–2230. [Google Scholar] [CrossRef]
  73. Lacarra-Etxarri, I.; De La Torre, U.; González-Marcos, J.A.; González-Velasco, J.R.; Pereda-Ayo, B. Polyethylenimine-functionalized SBA-15 mesoporous silica for CO2 direct air capture and conversion to methane in a coupled catalytic reactor. J. CO2 Util. 2025, 97, 103134. [Google Scholar] [CrossRef]
  74. Niu, Y.; Zou, S.; Liu, H.; Hu, M.; Cai, J.; Li, C.; Mumford, K.A.; Alivand, M.S.; Barzagli, F.; Zhang, R. Functionalized SBA-15-based catalysts for energy-efficient CO2 desorption: Bridging experimentation and machine learning to enhance amine sorbents regeneration. Chem. Eng. J. 2025, 522, 167944. [Google Scholar] [CrossRef]
  75. Yang, W.; Shirazian, S.; Soltani, R.; Zare, M.H. Bio-originated mesosilicate SBA-15: Synthesis, characterization, and application for heavy metal removal. npj Clean Water 2024, 7, 49. [Google Scholar] [CrossRef]
  76. García-Martínez, J.; Serrano-Torregrosa, E. Chemistry Education: Best Practices, Opportunities and Trends; Wiley: Hoboken, NJ, USA, 2015; ISBN 978-3-527-33605-0. [Google Scholar]
  77. Liang, B.; Zhu, P.; Gu, J.; Yuan, W.; Xiao, B.; Hu, H.; Rao, M. Advancing Adsorption and Separation with Modified SBA-15: A Comprehensive Review and Future Perspectives. Molecules 2024, 29, 3543. [Google Scholar] [CrossRef] [PubMed]
  78. Gatti, G.; Costenaro, D.; Vittoni, C.; Paul, G.; Crocellà, V.; Mangano, E.; Brandani, S.; Bordiga, S.; Cossi, M.; Marchese, L. CO2 adsorption on different organo-modified SBA-15 silicas: A multidisciplinary study on the effects of basic surface groups. Phys. Chem. Chem. Phys. 2017, 19, 14114–14128. [Google Scholar] [CrossRef]
  79. Wawrzyńczak, A.; Nowak, I.; Feliczak-Guzik, A. SBA-15- and SBA-16-Functionalized Silicas as New Carriers of Niacinamide. Int. J. Mol. Sci. 2023, 24, 17567. [Google Scholar] [CrossRef]
Figure 1. Schemes of ordered mesoporous structures. From left to right: SBA-15 (p6mm) with hexagonal cylindrical channels, KIT-6 (Ia3d) with interpenetrated cubic bicontinuous structure, SBA-16 (Im3m) with body-centered cubic geometry, FDU-12 (Fm3m) with open cubic.
Figure 1. Schemes of ordered mesoporous structures. From left to right: SBA-15 (p6mm) with hexagonal cylindrical channels, KIT-6 (Ia3d) with interpenetrated cubic bicontinuous structure, SBA-16 (Im3m) with body-centered cubic geometry, FDU-12 (Fm3m) with open cubic.
Suschem 06 00042 g001
Figure 2. Schematic representation of the main functionalization routes of SBA-15 and their associated performance effects.
Figure 2. Schematic representation of the main functionalization routes of SBA-15 and their associated performance effects.
Suschem 06 00042 g002
Table 1. Comparison of mesoporous materials.
Table 1. Comparison of mesoporous materials.
Mesoporous MaterialStructurePore Size
(nm)
Surface Area
(m2 g−1)
Typical Synthesis MethodRef.
MCM-41Hexagona
lp6mm
2–10800–1000CTAB-assisted sol–gel[4,25]
SBA-15Hexagonal p6mm5–30500–1000P123-templated sol–gel[20,25]
KIT-6Cubic
Ia3d
6–12400–900P123/butanol cosurfactant system[27,29]
FDU-12Cubic
Fm3m
5–15600–900Modified triblock copolymers[27,30]
MSU-XDisordered2–12500–700Nonionic self-assembly[29]
TUD-1Interconnected3–10600–850Surfactant-assisted sol–gel[29]
MCM-48Cubic
Ia3d
2–4800–1100CTAB-templated sol–gel[4,25]
FSM-16Lamellar2–3900–1000Surfactant intercalation[4]
Table 2. SBA-15 synthesis methods and their characteristics.
Table 2. SBA-15 synthesis methods and their characteristics.
Synthesis Method Conditions Pore Size (nm) D e f f
(m2 s−1)
Key AdvantagesRef.
Conventional sol–gelHCl, 35–40 °C, 24–72 h6–8~1.0 × 10−8High ordering, easy scalability[22]
Hydrothermal100–130 °C, 24–48 h6–101.2 × 10−8Enhanced crystallinity, structural stability[8,27]
Microwave-assisted2.45 GHz, 80–100 °C, 1–2 h5–83.5 × 10−7Rapid synthesis, uniform particle size control[3,30]
SonochemicalUltrasonic frequency, low T5–91.5 × 10−7Improved homogeneity, morphological dispersion[21]
SolvothermalOrganic solvents, high pressure5–7~1.8 × 10−8Precise shape/crystallinity/particle size control[4]
Dual-template (P123/CTAB)Mixed surfactants4–12 (bimodal)~1.0 × 10−8Hierarchical meso-macroporous structures[15]
Post-synthesis graftingGrafting con –NH2, tioles, etc.4–72.8 × 10−7Tailored surfaces for selective adsorption/catalysis[10,19]
Table 3. Functionalized mesoporous materials: synthesis methods, representative conditions, and corresponding performance outcomes.
Table 3. Functionalized mesoporous materials: synthesis methods, representative conditions, and corresponding performance outcomes.
Base MaterialFunctional Group/ModifierIncorporation MethodConditionsApplicationRef.
SBA-15–NH2
(aminopropyl, APTES)
GraftingEtOH, 60–80 °C, 12–24 hCO2 capture, drug immobilization, VOCs[10,39,47,48]
SBA-15Bis-Schiff base3-step anchoring (silanization + condensation)Organic solvent, RT–80 °CSelective removal of Pb(II) and other metals[12,35]
SBA-15–COOH (carboxylic acid)Post-synthesis oxidationHNO3, 50–80 °C, 6–12 hAdsorption of dyes, metals[17,45]
SBA-15–SH (tiol)Co-condensation or graftingpH acid, 25–50 °CAdsorption of noble metals (Au, Ag, Pt)[17,45]
SBA-15Fe3O4 (magnetic oxides)Coprecipitation in mesostructure60–90 °C, pH 8–9Magnetic separation, reuse[17,41,49]
SBA-15Pt/Pd o Ni, CuImpregnation + reduction200–300 °C, H2O o ArHeterogeneous catalysis (hydrogenation, oxidation)[35,42,44,50,51]
SBA-15Organic groups (alkyls, phenyls)Grafting or co-condensationRT–120 °C, organic solventHydrophobicity tuning, drug anchoring[11,52]
SBA-15β-Cyclodextrin
(β-CD)
Modification by supramolecular anchoringRT–50 °C, aqueous solventAdsorption of organic contaminants[36,37]
SBA-15AzobenzenePhotoactive modificationRT–60 °C Organic solvent
UV (365 nm) or visible (>450 nm) light
Photoactivated diffusion control, sensors[3,40]
MCM-41–COOH (carboxylic acidPost-synthesis oxidationAPS, HNO3, 50–80 °CAdsorption of dyes[45]
KIT-6–SH (tiol)Co-condensationpH acid, 25–50 °CAdsorption of heavy metals (Hg2+, Cd2+)[17]
CMK-3–SO3H (sulfonic acid)Reflux with H2SO480–120 °C, 6–12 hEsterification, acid catalysis[34]
MOF-5–NH2, –COOHReflux with H2SO4Solvothermal temperingSelective adsorption, sensors, catalysis[4]
Table 4. Quantitative comparison of functionalization strategies for SBA-15 and their mechanistic trade-offs.
Table 4. Quantitative comparison of functionalization strategies for SBA-15 and their mechanistic trade-offs.
StrategySurface Coverage (Groups nm−2)ΔBET (%)Thermal Stability (°C)Key ApplicationMechanistic Trade-off
Grafting –NH2 (APTES, MPTMS)0.5–3.0−20 to −30≈250CO2 capture, heavy-metal and VOC adsorptionHeterogeneous coverage in narrow pores; possible diffusion constriction
Co-condensation (–NH2–SH)1.0–3.5≈−20<200Uniform functional sites, solid-phase extractionSlight pore shrinkage; lower thermal stability of organic groups
Metal impregnation (Pt, Cu, Ni)≈−10≈600Redox catalysisAgglomeration or leaching if anchoring is weak; partial pore blocking
Isomorphic substitution (Si → Al)≈−10≈600Brønsted acidity, selective adsorptionReduced surface area but enhanced acid strength and gas-phase reactivity
Supramolecular (β-CD, azobenzene)−15 to −20<200Host–guest adsorption, photo-responsive controlLower thermal robustness; potential desorption of guest molecules
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Muñoz, M.; Flores, D.; Morillo, G.; Narváez, R.; Marcilla, A.; Rosero, M. Comprehensive Review of SBA-15 Mesoporous Silica: Functionalization Strategies, Diffusion Mechanisms, and Emerging Applications. Sustain. Chem. 2025, 6, 42. https://doi.org/10.3390/suschem6040042

AMA Style

Muñoz M, Flores D, Morillo G, Narváez R, Marcilla A, Rosero M. Comprehensive Review of SBA-15 Mesoporous Silica: Functionalization Strategies, Diffusion Mechanisms, and Emerging Applications. Sustainable Chemistry. 2025; 6(4):42. https://doi.org/10.3390/suschem6040042

Chicago/Turabian Style

Muñoz, Morayma, Diego Flores, Grace Morillo, Ricardo Narváez, Antonio Marcilla, and Marco Rosero. 2025. "Comprehensive Review of SBA-15 Mesoporous Silica: Functionalization Strategies, Diffusion Mechanisms, and Emerging Applications" Sustainable Chemistry 6, no. 4: 42. https://doi.org/10.3390/suschem6040042

APA Style

Muñoz, M., Flores, D., Morillo, G., Narváez, R., Marcilla, A., & Rosero, M. (2025). Comprehensive Review of SBA-15 Mesoporous Silica: Functionalization Strategies, Diffusion Mechanisms, and Emerging Applications. Sustainable Chemistry, 6(4), 42. https://doi.org/10.3390/suschem6040042

Article Metrics

Back to TopTop