Next Article in Journal
Exploring the Therapeutic Efficacy of Parsley (Petroselinum crispum Mill.) as a Functional Food: Implications in Immunological Tolerability, Reduction of Muscle Cramps, and Treatment of Dermatitis
Previous Article in Journal
Protein Misfolding in Pregnancy: Current Insights, Potential Mechanisms, and Implications for the Pathogenesis of Preeclampsia
Previous Article in Special Issue
Synergistic Remediation of Organic Dye by Titanium Dioxide/Reduced Graphene Oxide Nanocomposite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Porphyrin-Based Nanomaterials for the Photocatalytic Remediation of Wastewater: Recent Advances and Perspectives

by
Nirmal Kumar Shee
and
Hee-Joon Kim
*
Department of Chemistry and Bioscience, Kumoh National Institute of Technology, Gumi 39177, Republic of Korea
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(3), 611; https://doi.org/10.3390/molecules29030611
Submission received: 22 December 2023 / Revised: 8 January 2024 / Accepted: 12 January 2024 / Published: 26 January 2024

Abstract

:
Self-organized, well-defined porphyrin-based nanostructures with controllable sizes and morphologies are in high demand for the photodegradation of hazardous contaminants under sunlight. From this perspective, this review summarizes the development progress in the fabrication of porphyrin-based nanostructures by changing their synthetic strategies and designs. Porphyrin-based nanostructures can be fabricated using several methods, including ionic self-assembly, metal–ligand coordination, reprecipitation, and surfactant-assisted methods. The synthetic utility of porphyrins permits the organization of porphyrin building blocks into nanostructures, which can remarkably improve their light-harvesting properties and photostability. The tunable functionalization and distinctive structures of porphyrin nanomaterials trigger the junction of the charge-transfer mechanism and facilitate the photodegradation of pollutant dyes. Finally, porphyrin nanomaterials or porphyrin/metal nanohybrids are explored to amplify their photocatalytic efficiency.

1. Introduction

Environmental pollution has become a significant issue. Earth suffers from pollution (air and water) resulting from the rapid development of civilization and industrialization. Every year, large quantities of dangerous compounds, including synthetic dyes, cosmetics, insecticides, pharmaceutical products, and aromatic compounds, are released from the textile, leather, drug, and paper printing industries into water. These toxic pollutants not only damage the potability of water but also pose a significant risk to the entire ecosystem and have detrimental impacts on marine life [1,2]. This has led to an increasing number of scientific reports by environmentalists and researchers on developing strategies for the environmental purification of wastewater [3,4,5,6]. Various physicochemical methods, including filtration [7], adsorption [8], precipitation [9], chemical coagulation [10], electrochemical methods [11], bacterial treatment [12], and advanced oxidation processes (AOPs) [13], have been demonstrated in previous reports to remove these toxic pollutants from contaminated water. AOPs are the most adaptable techniques used in pollution control because of their uncomplicated procedures, economic affordability, and high activity for the degradation of hazardous pollutants to less hazardous H2O and CO2 without the creation of any other secondary pollutants. In AOPs, a suitable photocatalyst absorbs visible light and generates ROS in situ, thereby accelerating the removal of organic compounds from water. In particular, photocatalysis has received considerable attention, as it benefits from ecological advantages and additional economic benefits related to the utilization of freely accessible natural sunlight. Moreover, its mechanism is simple. First, light harvesting followed by exciton diffusion occurs, and after that, charge separation takes place, which is accompanied by carrier transport. Therefore, light absorption and successive electron shifts are pivotal for achieving efficient solar energy harvesting using photocatalysts. Nanoscale architectures in photofunctional compounds traditionally display typical optoelectronic properties based on their sizes and shapes. Consequently, the fabrication and design of nanoscale assemblies are important for constructing photocatalytic systems with specific properties and functions [14,15].
Considering the superiority of photocatalytic processes, several organic and inorganic-based nano- and micro-structures have emerged as new building blocks to fabricate photocatalytic systems. Various inorganic nanostructures, including metal oxide nanoparticles (MNPs) (TiO2 [16,17] and ZnO [18,19]), graphitic carbon nitride (g-C3N4) [20,21], silicates [22,23], zeolites [24,25], and bismuth-based photocatalysts [26,27], have been used to eliminate hazardous pollutants from wastewater. Titanium dioxide (TiO2) has been used as an emerging catalyst because of its high stability, moderate toxicity, economic affordability, and excellent catalytic properties for the decay of toxic contaminants in aqueous solutions. However, its high Eg~3.2 eV arising from the high strength of recombination between photogenerated pairs (hole = h+ and electron = e) limits its absorption of solar light (λ < 370 nm). This leads to a low photodegradation efficiency under visible light. It is extremely difficult to use these catalysts multiple times, and their complete recovery at the end of the reactions that they catalyze is impossible. Moreover, a large quantity of photocatalysts is required for loading in the photocatalytic method to achieve a reasonable degradation rate [28,29].
Conversely, organic nanostructures have secured unusual recognition compared with their inorganic analogs. This is because of the substantial versatility of their molecular schemes and the tremendous adjustability of their photoelectronic and solution-phase properties, which make them suitable for photocatalytic systems. Among the several organic building blocks, such as carbon quantum dots [30], fullerene [31], and graphene oxide or reduced graphene oxide [32], used to construct nano- and micro-structured organic nanomaterials with well-defined shapes and dimensions, the distinguished π-conjugated planar porphyrinoids (free porphyrin or metalloporphyrin) are of great interest for their tremendous optoelectronic properties and variable coordination features [33,34,35,36]. Owing to their large absorption coefficients in the Soret (400–440 nm) and Q bands (500–720 nm), porphyrin compounds are efficient at capturing light over a longer range (UV–vis) and initiating ROS across a spin-forbidden intersystem crossing pathway. Additionally, their rigid structural framework and intrinsic aromatic electronic qualities initiate self-aggregation in the solid and solution phases. However, the use of free or immobilized porphyrin compounds in photocatalysis is restricted because of their easy agglomeration in solutions and their deactivation, lower reusability, and difficulties in the separation of catalysts after the reaction. Therefore, new photocatalysts must be developed to overcome these challenges.
For this purpose, porphyrin-based nanomaterials have emerged as photocatalysts for the decay of hazardous pollutants in wastewater over the last two decades, and they are highly regarded by material chemistry researchers [37,38,39,40]. After aggregation, the formation of a nanoassembly of porphyrins not only increases solar absorption but also enhances stability for recycling. Moreover, their rigid framework resists hydrolysis and increases their photostability for further use. Numerous intermolecular non-covalent interactions (π-π stacking interactions; hydrogen bonding interactions; electrostatic, hydrophobic, and hydrophilic interactions; and metal–ligand coordination) are accountable for the self-aggregation of free porphyrin or metalloporphyrin compounds [41,42,43,44,45]. Unlike inorganic nanostructures, self-assembled porphyrin nanostructures are formed using several methods, including ionic self-assembly [46,47], reprecipitation [48,49], metal–ligand coordination [50,51], sonication-assisted self-assembly [52,53], and surfactant-assisted methods [54,55]. The morphologies of these organic nanostructures are not always uniform and are not suitable for numerous practical applications. Therefore, fabricating porphyrin-based nanoaggregates with specific dimensions is difficult. The fabrication of porphyrin-based nanostructures, compared to monomeric porphyrin units, not only changes the structural framework but also creates high structural modifications, including high thermodynamic stability, high porous surface area, interesting morphology, and outstanding photocatalytic degradation efficiency against pollutants under visible-light irradiation in an aqueous solution. The efficiency of harvesting solar energy in the visible region and the effective electron transfer properties of these materials make them unique compared to inorganic metal oxide nanoparticles. Therefore, the low cost, ease of synthesis, high efficiency, and high reproducibility of these nanomaterials make them outstanding candidates as visible-light photocatalysts.
In this review, we focus on the ongoing developments in this field, as well as the fabrication of distinct self-assembled porphyrin nano- and micro-structures with precise dimensions, and we highlight their photodegradation of pollutants from wastewater.

2. Fabrication of Porphyrin-Based Nanostructures

The self-assembly property of porphyrin nanostructures is based on several interactions at the molecular level, such as π-π interactions, electrostatic interactions, coordination bonding interactions, and hydrogen bonding interactions. In the case of porphyrin-based nanostructures, particular molecular interactions among the inherent constituents directly affect the conditions for self-assembly and the successive stages of self-aggregation. Based on these interaction forces, various fabrication processes have been effectively used to synthesize porphyrin-based nanoaggregates via self-assembly, including ionic self-assembly, coordination-based aggregation, and reprecipitation with and without the help of surfactants. Porphyrin-based nanostructures with various morphologies obtained using different self-assembly approaches are discussed in the following sections.

2.1. Reprecipitation Technique

Reprecipitation is the easiest method for fabricating isolated porphyrin-based nanomaterials. Porphyrin nanoaggregates are fabricated via inoculation using this method. In this process, a highly concentrated solution of porphyrin dissolved in a non-polar solvent is injected into a polar solvent with vigorous stirring. This technique is referred to as the solvent exchange method. In some cases, surfactants are used to prevent agglomeration and control the structure of assembled nanostructures. Hollow-shaped hexagonal nanoprisms comprising ZnTPyP with a uniform size and shape were fabricated using reprecipitation methods [56]. The nanostructures were fabricated by mixing a ZnTPyP solution (dissolved in DMF) with CTAB (in water) at room temperature under vigorous stirring. Scanning electron microscopy (SEM) and high-resolution TEM (HRTEM) were used to determine the surface fabric of the nanoaggregates (Figure 1). The average length of these nanoprisms ranged from 517 to 541 nm, and their width ranged from 92 to 98 nm. The fence of the nanoprisms was approximately 30 nm thick. The aspect ratio and dimensions of these nanoprisms could be controlled by regulating the chemical proportion of ZnTPyP to CTAB. The absorption spectra of the nanoprisms showed a reduction in the monomeric band at 424 nm and the J-aggregate band at 460 nm. XRD data confirmed that the axial coordination between the pyridyl-N atom and ZnII metal was responsible for this type of self-assembly. Interestingly, 1D hexagonal hollow nanoprisms can easily self-aggregate into uniform 3D architectures when the solvent evaporates.
Nguyen et al. described the use of a reprecipitation method without surfactants to construct porphyrin nanostructures from tin(IV)dipyridylporphyrin (Figure 2) [57]. An ethanolic solution of tin(IV)dipyridylporphyrin was added to water at room temperature. After several days, approximately 280 nm long oval-shaped flat crystals were obtained from this heterogeneous mixture (Figure 2a). The temperature-dependent morphology of the porphyrins was also investigated. Amorphous hollow spheres (diameter approximately ~350 nm) were obtained after injecting an ethanolic solution of tin(IV)dipyridylporphyrin into water at 90 °C (Figure 2b). This study describes a new method for tuning the extension of molecular nanocrystals in solutions under dynamic flourishing conditions, which can be assembled into substantial structures using CTAB.
Optically active covalently linked porphyrin–pentapeptide chains show a solvent-dependent morphology (Figure 3) [58]. Nanofibers were fabricated from THF (tetrahydrofuran) and n-hexane (1:3) mixtures, and nanotubes were obtained from THF and water (1:3) mixtures. In the case of the nanotubes, XRD results verified that a dimeric bilayer conformation was formed between the two conjugate molecules via intermolecular hydrogen bonding. However, a right-handed helix of porphyrin moieties was observed in the case of the nanofibers. The fabricated nanostructures exhibited good semiconductor properties.
Occasionally, solvents can control the supramolecular formation of porphyrin-based nanostructures. Zhu et al. reported that a zinc complex of benzyloxy-based porphyrin (ZnP2) displayed micrometer-sized leaves in a non-polar solvent such as n-hexane, whereas twisted nanorods were observed in methanol (Figure 4) [59]. This observation can be attributed to the different kinds of intermolecular interactions (π-π stacking interactions and hydrogen bonding interactions) arising between ZnP2 molecules in the presence of either polar or non-polar solvents. The width of the micrometer-shaped leaves was approximately 50 nm. The length of the twisted nanorods varied from 4 to 8 μm, while the width ranged from 500 to 800 nm, and the height ranged from 500 to 700 nm.
The ratio of “good solvent” and “bad solvent” can control the morphology of porphyrin nanostructures. Tetracenequinone units fused to nickel porphyrin (Ni4TQ) showed uniform flower-shaped biomorphs in mixed-solvent systems (MeOH and toluene) (Figure 5) [60]. Ni4TQ displayed regular four-leaf clover-like structures via drop casting in a mixed-solvent system (MeOH:toluene = 9:1). The diagonal length of this square clover was approximately 500 nm (Figure 5a). However, larger (~2.5 μm) flower-shaped microstructures were observed (MeOH:toluene = 8.5:1.5) as the MeOH percentage was increased (Figure 5b).
Coutsolelos et al. reported the solvent-dependent morphologies of tin(IV)porphyrin-based peptide conjugates (SnPy3P-FF) [61]. The precursor tripyridyl porphyrin macrocycles were covalently linked to the diphenylalanine dipeptide. SnPy3P-FF exhibited a solvent-dependent self-assembled morphology. In a typical procedure, SnPy3P-FF was dissolved in a “good solvent” (DCM, THF, or HFIP (1,1,1,3,3,3-hexafluoroisopropanol)), and then a “bad solvent” (MeOH or heptane) was injected to induce the self-assembly process. It exhibited spherical nanostructures in mixed-solvent systems, such as DCM/MeOH (2:8). Amorphous aggregates were obtained from THF, and agglomerated spheres were obtained from heptane/DCM (8:2) (Figure 6).

2.2. Ionic Self-Assembly (ISA)

ISA is a common method for the construction of nano- and micro-structured compounds. It involves the integration of various oppositely charged species, and they are held by electrostatic interactions. In binary porphyrin systems, anionic and cationic porphyrins promote self-assembly via electrostatic interactions between the oppositely charged molecules. So far, in porphyrin binary systems, the di-acid form of H4TPPS42− has been widely used. Shelnutt et al. reported porphyrin nanotubes from the combination of H4TPPS42− with cationic Sn(OH)2TPyP4+ [62]. High-resolution TEM images of the porphyrin nanotubes (Figure 7) revealed that their lengths varied from a few nanometers to a few micrometers, with their diameters varying from 50 to 70 nm. The wall thickness of the nanotubes was approximately 20 nm. It was confirmed that the porphyrin molecules were stacked in the J-aggregate mode, which resulted in the generation of red-shifted UV–visible absorption bands at 714 and 496 nm from the complementary bands of the starting porphyrin monomer (Figure 7).
Shelnutt et al. reported biomorph-like nanostructures with the configuration of a four-leaf clover [63]. For this, they used equal proportions of anionic ZnTPPS and cationic ZnT(N-EtOH-4-Py)P. Using SEM, the average diameter of these clovers was found to be roughly 5 ± 2 mm, and the shapes varied with temperature and other conditions (Figure 8). These self-assembled nanostructures were sufficiently stable for several months and could be utilized for effective hydrogen generation. Porphyrin clovers exhibit higher photocatalytic H2 evolution activity than individual porphyrins. This is due to the fact that, unlike monomeric porphyrins, the binary solids in clovers engage in cooperative behavior.
Sometimes, ionic self-assembly is possible by combining charged porphyrin building blocks with oppositely charged nonporphyrinic species. In 2015, Kim et al. reported porphyrin–polyoxometalate-based heterostructures using electrostatic interactions between polyoxomolybdate anions and tin(IV)porphyrin cations (Figure 9) [64]. DLS calculations confirmed that the diameter of these nanoparticles in H2O was approximately 245 nm with a slender disposition in size. These nanomaterials were exceptionally stable in both neutral and acidic media. Moreover, ionic self-assembled nanomaterials have been used as photosensitizers for H2 generation under sunlight.
Mourzina et al. prepared various porphyrin-based nanostructures by varying the ratio of cationic porphyrin [SnCl2TPyHP]4+ and anionic porphyrin H4TPPS42− in solutions (Figure 10) [65]. SEM images showed that these hollow tubular nanostructures had a length of 0.4–0.8 μm with open ends. The inner and outer diameters were 7–15 nm and 30–70 nm, respectively. These nanotube frameworks demonstrated a photoconductivity of (3.1 ± 0.9) × 10−4 s m−1 under visible light and behaved as a metal-like material.
Medforth et al. reported various nanostructures using the ionic self-assembly of anionic Fe(III)porphyrin (FeTSPP) with several cationic Fe(III)porphyrin building blocks in aqueous media [66]. The shape of the aggregates depended on the substituents on the cationic porphyrin and the pH of the solution (Figure 11). Hyperbranched structures were obtained by combining anionic FeTSPP with protonated pyridinium porphyrin (FeTHPyP) or protonated imidazolium porphyrin (FeTFHImP). In contrast, globular assemblies of sheets were obtained using N-methylpyridinium porphyrin (FeTMPyP). These fabricated nanoaggregates showed higher catalase activity than individual porphyrin molecules.

2.3. Surfactant-Assisted Self-Assembly (SASA)

Thus far, most porphyrin nanostructures are organized in aqueous media and depend on the pH, temperature, and concentration. However, surfactants act as a template in water and can influence the self-organization of nanoaggregates for practical applications. In 2012, Liu et al. reported on self-organized 1D nanofibers obtained using a surfactant-assisted reprecipitation method [67]. In a typical process, ZnTPyP was dissolved in CHCl3, and CTAB was dissolved in water. The ZnTPyP solution was then slowly injected into the CTAB solution with vigorous stirring. The diameters of the fabricated spherical nanospheres ranged from 0.1 to 1 μm. These self-assembled nanospheres were gradually converted into 1D nanofibers via aging under the appropriate conditions (Figure 12). The average length was found to be approximately 4.5 μm, and the diameter was approximately 50 nm. This result provides a novel approach for the creation of tractable porphyrin-based 1D nanoaggregates using surfactant-assisted self-assembly.
Patra et al. synthesized different surfactant-assisted assembled nanostructures of H2TCPP by varying the stirring time [68]. In a typical reaction, H2TCPP was dissolved in a dilute NaOH solution. CTAB was dissolved in a dilute HCl solution in another container. The alkaline solution of H2TCPP was then mixed with the acidic CTAB solution under persistent stirring (~1200 rpm), and the morphologies were examined after 15 min, 60 min, 6 h, and 48 h (Figure 13). Spherical nanoaggregates with a diameter of 100 nm were observed after 15 min of stirring (Figure 13a). These nanospheres were converted into rod-shaped nanoaggregates after stirring for 1 h (Figure 13b). The average length was approximately 230 nm, with a width of 73 nm. These rod-shaped aggregates were converted into flake-shaped microaggregates after stirring for 6 h (Figure 13c). The average length of the flakes was roughly 3.5 μm, and the width varied from 400 to 500 nm. At the end, flower-shaped nanoaggregates (Figure 13d) were assembled, with a size of approximately 4.5 μm after 48 h of stirring. The authors also reported that, without the surfactant CTAB, these nanoarchitectures from H2TCPP would not have been achieved.
In some cases, the morphology of nanostructures differs depending on the surfactant used for self-assembly. Bai et al. observed that surfactants were crucial for adjusting the morphology and arrangement of porphyrin molecules. They reported two different morphologies when they changed the surfactant from CTAB to SDS (sodium dodecyl sulphate) [69]. In a typical process, SnCl2TPP was dissolved in chloroform, and the surfactants (CTAB or SDS) were dissolved in water. An emulsion was formed by vigorously mixing the two solutions. After that, the solvent was vaporized from the heterogeneous system by heating at 60 °C. Octahedral arrays were obtained from the CTAB solution, nanosheets were obtained from the CTAB and emulsion, and microspheres were obtained from SDS. For a regular octahedron, the average length was estimated to range from 320 to 380 nm. However, the average length of the square-shaped nanosheets was calculated to be 400 ± 100 nm. The average diameter of the microspheres varied from 200 nm to 220 nm (Figure 14).
Fan et al. reported one-dimensional (1D) porphyrin nanomaterials using SASA [70]. The nanomaterials were constructed using H2THPP as the starting material. In a typical process, an acidic (pH 2.70) CTAB (2.5 mM) solution was added dropwise to an alkaline solution of H2THPP (0.0 mM) under vigorous stirring for 48 h. The fabricated nanowires were regular in shape. The average length of these monodispersed nanowires was approximately 4.25 μm, with a width of 110 nm (Figure 15). The 1D self-assembled nanowires of H2THPP were found to be almost 20 times better for hydrogen generation than monomeric H2THPP powder. This is due to the strong delocalization of excited-state electrons in the nanowire morphology, which improves the lifetime of photogenerated electron–hole pairs and enhances hydrogen production.
Zhu et al. reported porphyrin-based nanoleaves using the surfactant-assisted method, with the PdII complex of H2TCPP (PdTCPP) as the starting molecule and DASS as the surfactant [71]. Using a typical process, a DMF solution of PdTCPP was injected into an aqueous DASS solution and vigorously stirred at room temperature. Nanoleaves were collected by centrifuging the above solution. The nanoleaves were regular in shape and had an average size of 1.64 μm × 5.43 μm, with a thickness of approximately 200 nm (Figure 16).

2.4. Metal–Ligand Coordination-Assisted Self-Assembly

In the case of metalloporphyrins, the metal–ligand coordination bond plays a pivotal role in the conformation of molecular packing and controls the conformation of self-assembled nanomaterials. Jiang et al. explored the self-assembly of H2[DP(CH3COSC5H10O)2P] and its Zn complexes [72]. Both compounds were dissolved in two solvents (methanol and n-hexane) and then dropped onto a carbon-coated gauze for an SEM analysis. Metal-free porphyrin (M = 2H) formed nanoscale hollow spheres (diameter of ~700 nm) in methanol (Figure 17A) and nanoribbons (average width of ~18 nm) in n-hexane (Figure 17B). In contrast, the zinc complex (M = Zn) formed nanorods (50 nm in width and 800 nm in length) in methanol (Figure 17C) and nanospheres (diameter of ~40 nm) in n-hexane (Figure 17D). In these observations, it was evident that the fabrication of these nanostructures depended primarily on two factors: intermolecular metal–ligand interactions (O-Zn coordination) and π-π stacking interactions between tetra-pyrrole rings. However, for the metal-free porphyrin compound (M = 2H), π-π interactions controlled the fabrication of nanoaggregates. In contrast, metal–ligand interactions were responsible for the different nanostructures in the zinc complex (M = Zn).
Zhang et al. reported porphyrin nanoaggregates of H2THPP and its copper(II) complex CuTHPP in an aqueous solution [73]. Hollow nanospheres were obtained from H2THPP. The average particle diameter was approximately 70 nm. However, one-dimensional (1D) nanoribbons with a 10 µm length and a 70 nm width were obtained from CuTHPP (Figure 18). In the case of H2THPP, intermolecular π-π interactions in cooperation with hydrogen bonding interactions were responsible for the hollow-shaped nanospheres. In contrast, in the case of CuTHPP, intermolecular Cu–O metal–ligand coordination bonding was responsible for the ribbon-like morphology.
Yan et al. reported TPPS4-Bi nanotubes fabricated via out-of-plane coordination assisted by bismuth porphyrins [74]. These nanotubes were prepared from the reaction of meso-tetra(sulfonatophenyl)porphine (TPPS44−) with Bi(NO3)3 in water. The lengths, widths, and thickness of these nanotubes were approximately 500 ± 100 nm, 50 ± 5 nm, and 15 ± 5 nm, respectively (Figure 19). The Bi3+ metal ions placed between two TPPS44− units at a dislocation angle force a staggered arrangement of TPPS44−. This not only prohibited π-π stacking overlap but also permitted Bi-TPPS4 J-aggregation to conquer the self-quenching sequel.
Recently, Kim et al. fabricated porphyrin-based nanostructures from porphyrin triads via metal–ligand coordination self-assembly [75]. These triads were constructed by reacting Sn(IV)porphyrin with various porphyrin derivatives. The robust disposition of Sn(IV)porphyrin for the coordination of aryloxides and carboxylates was the main reason for the formation of steady coordination arrays. In triad 2, effective cooperative coordination among the 3-pyridyl N atoms of tin(IV)porphyrin with the Zn atoms of axial zinc porphyrins followed by π-π interactions led to the fabrication of regular nanofibers with an average width varying from 10 to 22 nm. The lengths of these nanofibers varied from nanometers to micrometers. It was interesting that the absence of Zn(II) metal in triad 1 resulted in non-uniform nanostructures, as only π-π stacking associations were dominant for aggregation. Notably, in the presence of an external ligand (pyrrolidine), the aggregation of the nanofibers was disrupted (Figure 20).
In another study, Kim et al. constructed porphyrin-based nanostructures based on metal–ligand coordination self-assembly [76]. The addition of silver salt (AgOTf) to a solution of the two tin(IV)porphyrin precursors led to the formation of different nanoaggregates. The axial ligation of tin(IV)porphyrin centers not only transformed the conformations of the supramolecular arrays but also generated high structural configurations, including a high porosity and chemical stability and an interesting morphology. Without axial ligation, the fused nanoparticles hooked on to each other, with an average length varying from 200 nm to 400 nm. With axial ligation, a rectangular flake-shaped nanoaggregate formed. The average width of these nanoflakes was approximately 25 nm, with the average length ranging from 250 to 600 nm (Figure 21).

2.5. Miscellaneous Processes

Porphyrin-containing nanoaggregates can also be constructed using other unfamiliar techniques. Hasobe et al. fabricated supramolecular porphyrin-based nanoaggregates via sonication [52]. A toluene solution of the porphyrin was mixed nine times with acetonitrile (v/v), and the solution was sonicated for half an hour at 288 K to fabricate porphyrin-based macroscopic materials (Figure 22). No nanostructures formed in the absence of sonication.
Choi et al. prepared 1D porphyrin-based nanotubes using a vaporization–condensation–recrystallization method [77]. In this process, H2TPyP solid powder was heated to 450 °C on a carbon-coated Si(100) substance in an inert gas atmosphere. The rectangular nanotubes were organized in the end domain of the furnace temperature (~350 °C). Crystallographic data confirmed that the abovementioned nanotubes formed by the self-stacking of H2TPyP building blocks via π-π, H-π, and hydrogen bonding interactions (Figure 23).
Purello et al. demonstrated a rare self-aggregation technique for the fabrication of porphyrin nanoaggregates [78]. In this method, a tetra-cationic porphyrin (CuT4) was assembled on a poly-glutamate template. Afterwards, a tetra-anionic porphyrin (H4TPPS2−) was assembled in the pre-templated aggregates. Self-assembled porphyrin nanoaggregates were obtained by removing the template (Figure 24).
Zhu et al. prepared porphyrin-based nanowires from naphthalimide-based linear metalloporphyrin ZnD(p-NI)PP using a drop-casting method [79]. Porphyrin nanowire films were fabricated by dropping a CHCl3 solution onto a silica substrate and copper grids and then drying them in air. The widths of these nanowires varied from 100 nm to 250 nm, with an average length ranging from 1.5 to 5.0 μm. In this case, ZnD(p-NI)PP was easily aggregated into nanowires with a distinct fabric via the cooperative π-π stacking interactions between the naphthalimide and porphyrin frameworks (Figure 25).

3. Application of Porphyrin-Based Nanostructures for Photocatalytic Remediation of Wastewater

In a photocatalytic reaction, photocatalysts are triggered by the absorption of adequate photon energy (hv) (≥bandgap energy (Eg)) [80]. This process will create a charge separation due to the boosting of e from the lowest VB energy level to the highest CB energy level, thus creating a pair of holes. A photocatalyzed reaction can be favored if the recombination of hole pairs is prevented. This is because the activated electrons react with dye to generate reduced species, and the generated holes react with dye to produce an oxidized product. Alternatively, the photogenerated electrons may react with dissolved O2 and generate superoxide radical anions (O2−•). Furthermore, photogenerated holes may react with H2O or OH and oxidize them into hydroxyl radicals (OH). Sometimes, other oxidants such as peroxide radicals may be generated for photodecomposition. The resulting hydroxyl radicals and superoxide radical anions are strong oxidizing agents that can degrade toxic dyes to lower-molecular-weight compounds and finally mineralize them to CO2 and H2O. The appropriate reactions on the catalyst surface for the decay of toxic pollutants can be expressed as follows:
P + hv → P*[h+(VB) + e(CB)]
H2O + P*[h+(VB)] → P + OH + H+
OH + P*[ h+(VB)] → P + OH
O2 + P*[e(CB)] → P + O2−•
O2−• + H+ → HO2
hVB+ + pollutant → oxidation products
eCB + pollutant → reduction products
Pollutant + OH → degraded products
O2−• + pollutant → degraded products
Moreover, the photocatalytic activity of a photocatalyst bank depends on several important factors, such as the surface area, bandgap, recombination energy of hole pairs, surface morphology, active sites, crystallinity, phase composition, penetration of light through the photocatalysts, and adsorption capability of dyes on the surface of the photocatalysts. Owing to their vast morphologies and functional properties, porphyrin-based self-assembled nanomaterials have been extensively investigated for the separation of pollutant dyes from wastewater [75,76]. This is because porphyrin nanomaterials can be readily produced, and their morphology can be easily managed. Tuning the size and morphology of nanomaterials has a favorable effect on their properties, such as the light-harvesting properties, surface area, active sites and defect sites, movement of hole pairs along the morphology, and absorbance of dyes. These robust materials can directly harvest photon energy to produce radicals, prevent the degradation of energetic species, and increase their recyclability. Porphyrin-based nanostructures show a higher photocatalytic ability than monomeric amorphous porphyrins. This is because of the strong π-π interactions in J- or H-type interactions. Porphyrin nanostructures have excellent semiconductor properties and can improve charge separation during photocatalytic process. Generally, porphyrin-based nanomaterials absorb photon energy in the visible range, and electrons in the VB jump to the CB to generate electron–hole pairs. A strong electronic delocalization of these charged species occurs on the surface of porphyrin-based nanostructures. This delays the recombination of electron–hole pairs and thus enhances their photodegradation performance. Photogenerated electron–hole pairs participate in the degradation process to generate reactive oxygen species and oxidize pollutants to nontoxic CO2 and H2O.
Each porphyrin-based nanostructure is different in terms of morphology. Each morphology has its own surface properties, such as surface energy, surface area, light-harvesting ability, and number of active and defect sites. This affects the delocalization of photogenerated charge species on the surface of the catalyst and alters the recombination energy of electron–hole pairs. Therefore, photocatalytic performance is controlled by the morphology of porphyrin-based nanostructures [75,76].
Reactive oxygen species generated during the photocatalytic process are evident in radical trapping experiments using scavengers or in electron spin resonance experiments, such as when tert-butanol (tBuOH) is employed to seize OH, NaN3 (sodium azide) is employed to seize 1O2, p-BQ (para-benzoquinone) is employed to seize O2−•, and Na2-EDTA (ethylenediaminetetraacetic acid disodium) is employed to seize h+ during photocatalytic processes [81,82].
In one example, various morphologies of H2TCPP showed different efficiencies for the degradation of water-soluble RhB (Figure 13) [68]. The photodegradation rates were estimated as 56% (nanospheres), 71% (nanoflowers), 79% (nanoflakes), and 81% (nanorods) (Figure 26). Fluorescence decay data revealed that the H2TCPP molecules in the nanoflakes and nanoflowers exhibited both J- and H-aggregations. In contrast, in the case of the nanorod aggregates, the H2TCPP molecules showed only J-aggregation, which increased the electronic delocalization of charge carriers and facilitated the degradation reaction (Figure 26). The J-aggregated structures improved coherent electronic delocalization due to the strong intermolecular π electronic coupling that occurred between the coherently aligned chromophores and became efficient photosemiconductors.
In another study, SnCl2TPP nanocrystals exhibited attractive morphology-dependent photocatalytic performance for the decomposition of MO dye in the presence of sunlight (Figure 14) [69]. The tin porphyrin nanosheets exhibited 20% MO degradation after 160 min. In contrast, the photodegradation performance of MO exceeded 60% for nanooctahedra and 90% for nanospheres (Figure 27).
Fan et al. reported on nanostructures derived from H2TPyP and ZnTPyP via surfactant-assisted reprecipitation [83]. The morphologies varied from nanooctahedra for H2TPyP to nanowires for ZnTPyP. By attentively controlling the assembly behavior, ZnTPyP nanowires exhibited higher photocatalytic performance towards MO dye than H2TPyP nanooctahedra (Figure 28).
Bhosale et al. reported on the morphology-dependent photodegradation of RhB using TTOP nanostructures [84]. The dimensions of the nanoaggregates could be regulated simply by adjusting the THF/H2O ratio. The photodegradation rate constant of the RhB dye was observed to be 0.0043 min−1 for the TTOP monomer. However, the photodegradation rate constants of the RhB dye for the aggregates fabricated with the THF/H2O ratio were found to be 0.678 h−1 for THF/H2O (3:7), 0.156 h−1 for THF/H2O (2:8), and 0.21 h−1 THF/H2O (1:9) (Figure 29).
Kim et al. reported on the morphology-controlled photodegradation of MB using Sn(IV)porphyrin-based nanomaterials (Figure 20) [75]. The photodegradation rate of the MB dye obtained when using nanofibers derived from triad 4 (1.68 h−1) was higher than that obtained when using nanocomposites from triad 3 (0.67 h−1), nanoparticles from triad 2 (0.54 h−1), and nanocomposites from triad 1 (0.48 h−1) (Figure 30). These observations suggest that the catalytic photodegradation performance of the MO dye was affected by the morphology of the photocatalyst. It is likely that different morphologies can affect catalytic activity because they provide different shapes, permanent porosities, surface energies, and numbers of active and defect sites. Moreover, strong electronic delocalization of the photogenerated charged species occurred on the surface of the catalysts. This improves the delay in the recombination of electron–hole pairs and thus results in photodegradation activity. Overall, the morphology of the catalyst controls the charge-transfer process and photocatalytic activity.
In another study, Kim et al. reported the morphology-dependent photodegradation of RhB in various nanoaggregates derived from metalloporphyrin-based triads [85]. These triads exhibited divergent nanoaggregates because of the various side chains found in the axial zinc porphyrin segments. The RhB dye degradation rate constant (first order) of T3 (1.0 × 10−2 min−1) was slightly higher than that of T1 (4.8 h−1), T2 (0.54 h−1), T4 (0.42 h−1), T5 (0.30 h−1), and T6 (0.36 h−1) (Figure 31).
The same group observed the morphology-dependent photodegradation of MO using metalloporphyrin-based triads under sunlight [51]. The reaction of ZnL (meso-5-(4-hydroxyphenyl)-10,15,20-tris(4-octoxyphenyl)porphyrinato)zinc(II)) with trans-dihydroxo-[5,15-bis(2-pyridyl)-10,20-bis(phenyl)porphyrinato] tin(IV), trans-dihydroxo-[5,15-bis(3-pyridyl)-10,20-bis(phenyl)porphyrinato] tin(IV), and trans-dihydroxo-[5,15-bis(4-pyridyl)-10,20-bis(phenyl)porphyrinato] tin(IV) in anhydrous toluene under an Ar atmosphere led to the formation of T2, T3, and T4 respectively. The positions of the pyridyl nitrogen atoms in the tin(IV)porphyrin building blocks differed. The photodegradation rate of MO was calculated to be 76–94% under sunlight within 100 min. Nanoflake-shaped T4 and nanosphere-shaped T2 exhibited lower catalytic performances than nanorod-shaped T3 (Figure 32). It is likely that porphyrin nanostructures increase solar light absorption, owing to the formation of aggregated architectures as compared to monomeric units. Extensive delocalization occurs on the surface of the nanostructures through the π-π interactions. This lowers charge recombination by stabilizing the photogenerated electron–hole pairs, subsequently enhancing lifetime and improving degradation performance. Therefore, morphology controls the photocatalytic activity of porphyrin-based nanostructures.
Kim et al. synthesized several tin-porphyrin-containing ionic compounds by reacting several acids (HCl, HNO3, H2SO4, H2CO3, CF3COOH, and H3PO4) with Sn(OH)2TPhPyP (1) [86]. These ionic complexes showed morphology-dependent photocatalytic activity against the decay of MG dye in sunlight, with efficiencies ranging from 50% to 95% (Figure 33) in an aqueous solution. The rate of the photodegradation of MG dye by photocatalyst 1 was 0.84 h−1. However, the rate constants of the ionic compounds obtained from the reaction of photocatalyst 1 with HNO3, HCl, H2SO4, H2CO3, CF3COOH, and H3PO4 were 0.9 h−1, 1.08 h−1, 0.96 h−1, 1.32 h−1, 1.92 h−1, and 2.34 h−1, respectively.
The same group reported two porphyrin-based molecular square arrays: 2 and 3 [76]. These compounds were constructed from the 72 h refluxing of Re(CO)5Cl in mixed solvents (THF + toluene) with trans-dihydroxo-[5,15-bis(4-pyridyl)-10,20-bis(phenyl)porphyrinato]tin(IV) and trans-dihydroxo-[5,15-bis(4-pyridyl)-10,20-bis(4-tert-butylphenyl)-porphyrinato]tin(IV), respectively [87]. The structural configuration of these frameworks was rigid, owing to the locked 2D tetrameric array. Changing the side chain on the metalloporphyrin building blocks not only generated structural motifs but also led to the formation of precisely divergent nanoaggregates. These non-aggregated materials showed effective photodegradation of EBT dye in sunlight. The removal efficiencies were 88% for molecular array 2 and 95% for molecular array 3 within 90 min. It was observed that the photodegradation rate of EBT dye obtained when using molecular array 3 (1.92 h−1) was higher than that obtained when using molecular array 2 (1.38 h−1) (Figure 34).

4. Future Perspectives

Thus far, our discussion has been limited to self-assembled nanomaterials fabricated from various porphyrin derivatives and has examined the photocatalysis of pollutant dyes as free-standing organic semiconductors. During the photocatalysis process, photoinduced charge pairs (e and h+) experience either recombination or reach the interface of the photocatalyst and facilitate several photodegradation reactions. To improve photocatalytic performance in terms of surface area, light-harvesting capacity, reusability, and coherent charge separation, porphyrins or nanostructured porphyrins can be combined with other nanomaterials, such as various organic (graphene, graphene oxide, or reduced graphene oxide) or inorganic semiconductors (TiO2, ZnO, g-C3N4, etc.).
Inorganic semiconductors such as TiO2 can be integrated with metalloporphyrin aggregates to permit light accumulation in a broad range of sunlight. This improves charge dissociation and photodegradation efficiency. Ji et al. demonstrated the improved photodegradation activity of various metalated (Cu, Mn, Fe, and Co) complexes of meso-tetraphenylporphyrin adsorbed on the surface of TiO2 (P-25) [88]. It should be mentioned that copper porphyrin-immobilized TiO2 (CuP-TiO2) exhibited outstanding photodegradation performance for the decay of MO over other metalloporphyrins under UV light irradiation (Figure 35). The high photocatalytic activity of CuP-TiO2 can be explained by the fact that Cu2+ readily gains an electron to reach a steady state, causing the isolation of electrons with holes compared to other metal ions in the composites. Moreover, Cu2+ (0.73 Å) and Ti4+ (0.64 Å) have similar ionic radii. Therefore, Cu2+ can easily penetrate TiO2 as a deep acceptor in conjunction with the neighboring O vacancies of Ti4+.
Lee et al. demonstrated that silica-immobilized tin porphyrin (silica/SnP or s-SnP) photosensitizers facilitate the degradation of microcystins (MCs) despite the low generation of photosensitized singlet oxygen [89]. An equivalent study using Rose Bengal (RB), C60 aminofullerene, and s-SnP suggested that the RR-MC decomposition rate is directly related to the photosensitizing properties (Figure 36). This study also suggested that direct electron movement from MCs to the triplet state of SnP facilitates the photodegradation of MCs under visible-light irradiation. Moreover, the incorporation of tin porphyrin not only facilitates direct photo-initiated electron transfer but also creates large surface modifications for effective MC sorption on the surface of hybrid catalysts. Singlet oxygen is no longer a reactive species in the degradation of MCs. This was confirmed by the negligible effects of singlet oxygen scavengers such as azide and D2O on the MC-RR degradation rate.
Graphene nanoparticles (GNPs) are one of the most common types of carbon-based compounds that are integrated with the porphyrin nanostructure due to their excess electron mobility and large surface area. During the photocatalytic reaction, graphene effectively transfers the photogenerated reactive pairs to the adsorbed dyes on the surface and enhances the degradation rate. La and co-authors demonstrated the fabrication of a graphene-incorporated porphyrin nanofiber (GNPs@H2TCPP). With the support of D-arginine, self-assembled H2TCPP was fabricated on the surface of graphene nanoparticles accompanied by the fabrication of GNPs@H2TCPP [90]. It was observed that this hybrid nanomaterial showed remarkable photocatalytic degradation performance against methyl orange dye (removal rate ~80%) after 180 min of visible-light irradiation (Figure 37). This could be attributed to graphene increasing the lifetime of the charged pairs by preventing the recombination of photogenerated reactive pairs generated by the porphyrin nanofiber.
Min et al. demonstrated that effective binding between porphyrinoids and optoelectronic materials is important for improving the photodegradation of toxic dyes. To this end, they synthesized different heterocomplexes based on the carboxyphenyl side chains of various porphyrins and anatase TiO2 [91]. All these hybrid materials exhibited enhanced photodegradation performance compared to pure TiO2 (Figure 38). The photodegradation activity of MB dye was 95%, compared to only 55% for pure TiO2, in the presence of sunlight within 1 h. Additionally, the authors observed that zinc-metalated porphyrins showed higher photocatalytic performance than the metal-free carboxyphenyl derivatives of porphyrins. This may be due to the electronic coupling between the zinc carboxyphenyl derivatives of the porphyrins and the CB of TiO2. Therefore, chemical surface modification with carboxyphenyl porphyrin and TiO2 not only strengthened the bonds between them but also improved the photocatalytic performance and prevented the separation of porphyrins from the TiO2 surface.
Fajardo et al. reported a hybrid microparticle (Zn(II)Pr@PAA) for the photodegradation of toxic pollutants under sunlight (where PAA = polyacrylic acid) [92]. Zn(II)-porphyrin was easily incorporated into the PAA matrix in situ to form Zn(II)Pr@PAA. Various analytical and microscopic methods confirmed that the incorporated metalloporphyrin influenced the optical properties of the Zn(II)Pr@PAA microstructures. The resulting photocatalysts exhibited improved photocatalytic performance with respect to the decomposition of nitrobenzene (NB) and MB dyes compared to the starting Zn(II)porphyrin (Figure 39). The degradation rate constants of all organic pollutants varied from 0.03 to 0.05 min−1.
Vo et al. demonstrated the improved photodegradation activity of ZnO@H2TCPP nanofibers [93]. The ZnO@H2TCPP nanofiber hybrid complex was forged via reprecipitation self-assembly. The results confirmed the consistent incorporation of the ZnO nanoparticles (ZnO-NPs) into the self-assembled H2TCPP nanofiber matrix. The bandgap energies of the TCPP nanofibers (2.6 eV), ZnO nanoparticles (3.21 eV), and ZnO@H2TCPP (3.12 and 2.43 eV) were determined using the Tauc plot method. The fabrication of porphyrin nanofibers on the ZnO nanoparticles lowered the bandgap energy and enhanced their light-harvesting properties. Due to the synergetic effect of the ZnO and H2TCPP nanofibers, the ZnO@H2TCPP composite showed outstanding photocatalytic activity for RhB degradation (rate constant = 0.02 min−1). Approximately 97% of the dye was removed after 180 min of visible-light irradiation (Figure 40).
Xu et al. fabricated a hybrid porphyrin-based heterojunction (SA-TCPP/O-CN) incorporated in situ via the reaction of H2TCPP nanoaggregates (SA-TCPP) and oxygen-doped g-C3N4 (O-CN) nanosheets [94]. In TEM and HRTEM studies, it was observed that O-CN exhibited a nanosheet shape with a length of a few hundred nanometers. In contrast, SA-TCPP showed nanocrystals, with an average diameter varying from 5 to 20 nm. It was found that the self-assembled porphyrin nanoaggregates were regularly fabricated on the surface of oxygen-doped g-C3N4 in SA-TCPP/O-CN, confirming the construction of a 0D/2D assembled hybrid structure. The formed 0D/2D heterostructures were successfully assembled via π-π interactions. The homogeneous fabrication of SA-TCPP/O-CN facilitated electron delocalization and promoted interfacial charge transfer. The energy-harvesting property of O-CN was significantly enhanced by the wide-spectrum reactive capacity of SA-TCPP. O-CN and SA-TCPP synergistically contributed to the enhanced photodegradation of phenol and 2,4-dinitrophenol (Figure 41).
Recently, Kim et al. fabricated two heterostructures (MCM-41@SnP and SiO2@SnP) derived from the chemisorption of SnP on SiO2 nanoparticles and mesoporous structured MCM-41 [95]. The immobilization of SnP on SiO2 and MCM-41 not only increased the surface area but also modified the morphology of the composites. These factors facilitated the photodegradation of RhB, erioglaucine, and m-cresol purple dyes in the presence of sunlight. The photodegradation rates of these dyes obtained when using the abovementioned photocatalysts varied from 87 to 95% under sunlight. MCM-41@SnP showed higher photocatalytic activity towards the decay of pollutant dyes than SiO2@SnP and SnP. MCM-41@SnP exhibited outstanding photostability and reusability (Figure 42).
In another study, the same group reported an efficient nanomaterial, SnP/AA@TiO2 (photocatalyst 4, AA = adipic acid), constructed by exploiting the coordination interaction between AA and the axial hydroxyl group in SnP on the TiO2 surface [96]. The SnP molecule was firmly attached to the TiO2 surface through an AA interconnection in photocatalyst 4, as verified by several spectroscopic techniques. Under visible-light irradiation, photocatalyst 4 exhibited outstanding photocatalytic RhB degradation. Photocatalyst 4 achieved 95% RhB dye removal within 80 min. The first-order decomposition rate constant was observed to be 3.66 × 10−2 min−1. The high degradation rate and greater stability of SnP-anchored photocatalysts make them more coherent than SnP@TiO2 or TiO2 (Figure 43).
Kim et al. reported two hybrid catalysts, SnPy@ZnO (5) and SnPy/AA@ZnO (6). The reaction of ZnO with trans-dihydroxo [5,10,15,20-tetrakis(4-pyridyl)porphyrinato]tin(IV) (SnPy) led to the formation of photocatalyst 5. Photocatalyst 6 was prepared via the reaction of ZnO nanoparticles with pretreated adipic acid (AA) [97]. In the case of photocatalyst 5, ZnO and SnPy were probably connected by a coordinative interaction between the Zn atom of the ZnO surface and the pyridyl-N atom of SnPy. In contrast, in the case of photocatalyst 6, the SnPy molecule was firmly combined with the ZnO nanostructure via adipic acid anchors. Photocatalyst 6 showed significantly improved photodegradation performance for the decay of AM dye under sunlight compared to SnPy, ZnO, and photocatalyst 5. The first-order photodegradation rates of AM dye were observed to be 0.06 h−1 (SnPy), 0.036 h−1 (ZnO), 0.048 h−1 (photocatalyst 5), and 2.88 h−1 (photocatalyst 6) (Figure 44).
Surface modification not only enhances the delocalization of the excited state of photogenerated electron–hole pairs and improves photocatalytic performance, but it also results in the generation of active sites for pollutants that are larger than those of TiO2, ZnO, or SnP.

5. Conclusions

This article provides an overview of the significant progress made in porphyrin-based nanomaterials for the photodegradation of contaminants under sunlight. Porphyrin-based self-assembled nanostructures are constructed using several methods, such as ionic self-assembly, reprecipitation, sonication-assisted, metal–ligand coordination, and surfactant-assisted methods. Various interactions, including π-π interactions, hydrogen bonding interactions, coordination bonding interactions, and electrostatic interactions, are commonly involved in the self-assembly of porphyrin molecules to fabricate nanostructures in a solid state or solution. All these porphyrin nanomaterials show remarkable photocatalytic performance compared to monomeric amorphous porphyrins for the removal of toxic pollutants under visible-light irradiation. This is because porphyrin nanoaggregates with ordered molecular structures enhance charge separation compared with monomeric porphyrins. The synthetic utility of porphyrins permits the organization of porphyrin building blocks into nanostructures, which can remarkably improve their light-harvesting properties and photostability. The tunable functionalization and distinctive structure of porphyrin nanomaterials trigger the junction of the charge-transfer mechanism and facilitate the photodegradation of pollutant dyes. However, despite the advances made in integrated methodologies to fabricate various porphyrin-based nano- or micro-structures, the photodegradation activity of these systems is still limited by the rapid recombination of photogenerated reactive pairs. Various porphyrin nanoaggregates have been incorporated with other nanoparticles, including graphene oxide, TiO2, ZnO, and g-C3N4, to enhance photodegradation activity in terms of light-harvesting capacity, permanent porosity, recyclability, and significant charge separation. Therefore, a composite of optoelectronic functional materials and porphyrins can contribute to the fabrication of nanohybrid materials for visible-light photocatalysis. Further investigation of porphyrin nanomaterials or semiconductor/porphyrin composites is required to improve their photocatalytic efficiencies. Thus far, discussions have been limited to the photodegradation of pollutants in water. The fabrication of novel porphyrin-based nanoaggregates in conjunction with the high harvesting efficiency of solar light could also be a major point for future research and applications in many important areas. The advantages of porphyrin-based nanoaggregates have been exploited in several important areas, such as fuel cells [98], sensing [99], H2 production [100], biomedical applications [101], CO2 reduction [102], cancer treatment [103], catalysis [104], and gas storage and separation [105].
In conclusion, we demonstrated that porphyrin-based nanomaterials acquired via self-assembly are ideal photocatalytic compounds for harvesting solar energy from a wide range of light spectra. These nanoaggregates are easily fabricated and modified, and they exhibit high photodegradation performance. Their capacity to absorb solar energy in the visible region and their effective electron transfer properties make them unique compared to metal oxide nanoparticles. Versatile porphyrin derivatives conjugated with various semiconducting materials will likely expand the degradation performance of environmentally friendly, low-cost, robust, and highly recyclable photocatalysts for water treatment in the future.

Author Contributions

Conceptualization, data curation, formal analysis, software, validation, and writing—original draft, N.K.S.; funding acquisition, project administration, resources, review, and writing—review & editing, H.-J.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Research Foundation of Korea (grant no. 2022R1F1A1074420) funded by the Korean government (MSIT).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AOPAdvanced oxidation process
AMAmaranth
CBConduction band
CTABCetyltrimethylammonium bromide
DCMDichloromethane
DASS1-Decanesulfonic acid sodium salt
DLSDynamic light scattering
DMFN, N-dimethylformamide
EBTEriochrome Black T
EgBand gap energy
H2TCPP5,10,15,20-Tetrakis(4-carboxyphenyl) porphyrin
H2TPyP5,10,15,20-Tetrakis(4-pyridyl)porphyrin
H2THPP5,10,15,20-Tetrakis(4-(hydroxyl)phenyl)porphyrin
H4TPPS42−5,10,15,20-Tetrakis(sulfonatophenyl)porphyrin
MeOHMethanol
MBMethylene blue
MOMethyl orange
MCM-41Mobil Composition of Matter No. 41
RhBRhodamine B
ROSReactive oxygen species
SEMScanning electron microscopy
SnCl2TPPDichloro-(5,10,15,20-tetraphenylporphyrinato)tin(IV)
SnPtrans-Dihydroxo(5,10,15,20-tetraphenylporphyrinato)tin(IV)
TEMTransmission electron microscopy
THFTetrahydrofuran
VBValence band
XRDX-ray diffraction
ZnTPyP[5,10,15,20-Tetrakis(4-pyridyl)porphyrinato]zinc(II)

References

  1. Wang, Z.; Walker, G.W.; Muir, D.C.G.; Nagatani-Yoshida, K. Toward a Global Understanding of Chemical Pollution: A First Comprehensive Analysis of National and Regional Chemical Inventories. Environ. Sci. Technol. 2020, 54, 2575–2584. [Google Scholar] [CrossRef] [PubMed]
  2. Naidu, R.; Biswas, B.; Willett, I.R.; Cribb, J.; Kumar Singh, B.; Paul Nathanail, C.; Coulon, F.; Semple, K.T.; Jones, K.C.; Barclay, A.; et al. Chemical Pollution: A Growing Peril and Potential Catastrophic Risk to Humanity. Environ. Int. 2021, 156, 106616. [Google Scholar] [CrossRef] [PubMed]
  3. Boelee, E.; Geerling, G.; van der Zaan, B.; Blauw, A.; Vethaak, A.D. Water and health: From environmental pressures to integrated responses. Acta Trop. 2019, 193, 217–226. [Google Scholar] [CrossRef] [PubMed]
  4. Baaloudj, O.; Assadi, I.; Nasrallah, N.; El Jery, A.; Khezami, L.; Assadi, A.A. Simultaneous removal of antibiotics and inactivation of antibiotic-resistant bacteria by photocatalysis: A review. J. Water Process. Eng. 2021, 42, 102089. [Google Scholar] [CrossRef]
  5. Huang, H.M.; Jiang, L.B.; Yang, J.J.; Zhou, S.Y.; Yuan, X.Z.; Liang, J.; Wang, H.; Wang, H.; Bu, Y.Q.; Li, H. Synthesis and modification of ultrathin g-C3N4 for photocatalytic energy and environmental applications. Renew. Sust. Energy Rev. 2023, 173, 113110. [Google Scholar] [CrossRef]
  6. Parvulescu, V.I.; Epron, F.; Garcia, H.; Granger, P. Recent Progress and Prospects in Catalytic Water Treatment. Chem. Rev. 2022, 122, 2981–3121. [Google Scholar] [CrossRef] [PubMed]
  7. Hollender, J.; Zimmermann, S.G.; Koepke, S.; Krauss, M.; McArdell, C.S.; Ort, C.; Singer, H.; von Gunten, U.; Siegrist, H. Elimination of Organic Micropollutants in a Municipal Wastewater Treatment Plant Upgraded with a Full-Scale Post-Ozonation Followed by Sand Filtration. Environ. Sci. Technol. 2009, 43, 7862–7869. [Google Scholar] [CrossRef]
  8. Zango, Z.U.; Jumbri, K.; Sambudi, N.S.; Ramli, A.; Abu Bakar, N.H.H.; Saad, B.; Rozaini, M.N.H.; Isiyaka, H.A.; Jagaba, A.H.; Aldaghri, O.; et al. A Critical Review on Metal-Organic Frameworks and Their Composites as Advanced Materials for Adsorption and Photocatalytic Degradation of Emerging Organic Pollutants from Wastewater. Polymers 2020, 12, 2648. [Google Scholar] [CrossRef]
  9. Thamer, B.M.; Aldalbahi, A.; Moydeen, A.M.; El-Newehy, M.H. In Situ Preparation of Novel Porous Nanocomposite Hydrogel as Effective Adsorbent for the Removal of Cationic Dyes from Polluted Water. Polymers 2020, 12, 3002. [Google Scholar] [CrossRef]
  10. Xu, Q.; Huang, Q.-S.; Luo, T.-Y.; Wu, R.-L.; Wei, W.; Ni, B.-J. Coagulation removal and photocatalytic degradation of microplastics in urban waters. Chem. Eng. J. 2021, 416, 129123. [Google Scholar] [CrossRef]
  11. Łuba, M.; Mikołajczyk, T.; Pierożyński, B.; Smoczyński, L.; Wojtacha, P.; Kuczyński, M. Electrochemical Degradation of Industrial Dyes in Wastewater through the Dissolution of Aluminum Sacrificial Anode of Cu/Al Macro-Corrosion Galvanic Cell. Molecules 2020, 25, 4108. [Google Scholar] [CrossRef]
  12. Saratale, R.G.; Saratale, G.D.; Chang, J.S.; Govindwar, S.P. Bacterial decolorization and degradation of azo dyes: A review. J. Taiwan Inst. Chem. Eng. 2011, 42, 138–157. [Google Scholar] [CrossRef]
  13. Bartolomeu, M.; Neves, M.G.P.M.S.; Faustino, M.A.F.; Almeida, A. Wastewater chemical contaminants: Remediation by advanced oxidation processes. Photochem. Photobiol. Sci. 2018, 17, 1573–1598. [Google Scholar] [CrossRef] [PubMed]
  14. Cardoso, I.M.F.; Cardoso, R.M.F.; da Silva, J.C.G.E. Advanced Oxidation Processes Coupled with Nanomaterials for Water Treatment. Nanomaterials 2021, 11, 2045. [Google Scholar] [CrossRef] [PubMed]
  15. Rana, A.G.; Tasbihi, M.; Schwarze, M.; Minceva, M. Efficient Advanced Oxidation Process (AOP) for Photocatalytic Contaminant Degradation Using Exfoliated Metal-Free Graphitic Carbon Nitride and Visible Light-Emitting Diodes. Catalysts 2021, 11, 662. [Google Scholar] [CrossRef]
  16. Bakbolat, B.; Daulbayev, C.; Sultanov, F.; Beissenov, R.; Umirzakov, A.; Mereke, A.; Bekbaev, A.; Chuprakov, I. Recent Developments of TiO2-Based Photocatalysis in the Hydrogen Evolution and Photodegradation: A Review. Nanomaterials 2020, 10, 1790. [Google Scholar] [CrossRef] [PubMed]
  17. Zhou, S.; Jiang, L.; Wang, H.; Yang, Y.; Yuan, X.; Wang, H.; Liang, J.; Li, X.; Li, H.; Bu, Y. Oxygen Vacancies Modified TiO2/O-Terminated Ti3C2 Composites: Unravelling the Dual Effects between Oxygen Vacancy and High-Work-Function Titanium Carbide. Adv. Funct. Mater. 2023, 33, 2307702. [Google Scholar] [CrossRef]
  18. Franco, P.; Sacco, O.; De Marco, I.; Vaiano, V. Zinc Oxide Nanoparticles Obtained by Supercritical Anti solvent Precipitation for the Photocatalytic Degradation of Crystal Violet Dye. Catalysts 2019, 9, 346. [Google Scholar] [CrossRef]
  19. Spoială, A.; Ilie, C.-I.; Trușcă, R.-D.; Oprea, O.-C.; Surdu, V.-A.; Vasile, B.Ș.; Ficai, A.; Ficai, D.; Andronescu, E.; Dițu, L.-M. Zinc Oxide Nanoparticles for Water Purification. Materials 2021, 14, 4747. [Google Scholar] [CrossRef]
  20. Ong, W.J.; Tan, L.L.; Ng, Y.H.; Yong, S.T.; Chai, S.P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer to Achieving Sustainability. Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  21. Wei, R.; Tang, N.; Jiang, L.; Yang, J.; Guo, J.; Yuan, X.; Liang, J.; Zhu, Y.; Wu, Z.; Li, H. Bimetallic nanoparticles meet polymeric carbon nitride: Fabrications, catalytic applications and perspectives. Coord. Chem. Rev. 2022, 462, 214500. [Google Scholar] [CrossRef]
  22. McIntyre, H.M.; Hart, M.L. Photocatalytic Porous Silica-Based Granular Media for Organic Pollutant Degradation in Industrial Waste-Streams. Catalysts 2021, 11, 258. [Google Scholar] [CrossRef]
  23. Zainudin, N.F.; Sam, S.T.; Wong, Y.S.; Ismail, H.; Walli, S.; Inoue, K.; Kawamura, G.; Tan, W.K. Degradation of Diazo Congo Red Dye by Using Synthesized Poly-Ferric-Silicate-Sulphate through Co-Polymerization Process. Polymers 2023, 15, 237. [Google Scholar] [CrossRef] [PubMed]
  24. Guang, H.; Yang, J.; Duan, X.; Farnood, R.; Yang, C.; Yang, J.; Liu, W.; Liu, Q. Recent developments and challenges in zeolite-based composite photocatalysts for environmental applications. Chem. Eng. J. 2021, 417, 129209. [Google Scholar]
  25. Pouramini, Z.; Mousavi, S.M.; Babapoor, A.; Hashemi, S.A.; Lai, C.W.; Mazaheri, Y.; Chiang, W.-H. Effect of Metal Atom in Zeolitic Imidazolate Frameworks (ZIF-8 & 67) for Removal of Dyes and Antibiotics from Wastewater: A Review. Catalysts 2023, 13, 155. [Google Scholar]
  26. Subhiksha, V.; Kokilavani, S.; Khan, S.S. Recent advances in degradation of organic pollutant in aqueous solutions using bismuth based photocatalysts: A review. Chemosphere 2022, 290, 133228. [Google Scholar] [CrossRef] [PubMed]
  27. Baaloudj, O.; Nasrallah, N.; Bouallouche, R.; Kenfoud, H.; Khezami, L.; Assadi, A.A. High efficient Cefixime removal from water by the sillenite Bi12TiO20: Photocatalytic mechanism and degradation pathway. J. Clean. Prod. 2022, 330, 129934. [Google Scholar] [CrossRef]
  28. Linsebigler, A.L.; Lu, G.Q.; Yates, J.T. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  29. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef]
  30. Heng, Z.W.; Chong, W.C.; Pang, Y.L.; Koo, C.H. An Overview of the Recent Advances of Carbon Quantum Dots/metal Oxides in the Application of Heterogeneous Photocatalysis in Photodegradation of Pollutants towards Visible-Light and Solar Energy Exploitation. J. Environ. Chem. Eng. 2021, 9, 105199. [Google Scholar] [CrossRef]
  31. Yao, S.; Yuan, X.; Jiang, L.; Xiong, T.; Zhang, J. Recent Progress on Fullerene-Based Materials: Synthesis, Properties, Modifications, and Photocatalytic Applications. Materials 2020, 13, 2924. [Google Scholar] [CrossRef] [PubMed]
  32. Sandhu, I.S.; Chitkara, M.; Rana, S.; Dhillon, G.; Taneja, A.; Kumar, S. Photocatalytic performances of stand-alone graphene oxide (GO) and reduced graphene oxide (rGO) nanostructures. Opt. Quantum Electron. 2020, 52, 359. [Google Scholar] [CrossRef]
  33. Drain, C.M.; Varotto, A.; Radivojevic, I. Self-Organized Porphyrinic Materials. Chem. Rev. 2009, 109, 1630–1658. [Google Scholar] [CrossRef] [PubMed]
  34. Beletskaya, I.; Tyurin, V.S.; Tsivadze, A.Y.; Guilard, R.; Stern, C. Supramolecular chemistry of metalloporphyrins. Chem. Rev. 2009, 109, 1659–1713. [Google Scholar] [CrossRef] [PubMed]
  35. Hasobe, T. Photo- and electro-functional self-assembled architectures of porphyrins. Phys. Chem. Chem. Phys. 2012, 14, 15975–15987. [Google Scholar] [CrossRef] [PubMed]
  36. Vithanage, B.C.N.; Xu, J.X.; Zhang, D. Optical Properties and Kinetics: New Insights to the Porphyrin Assembly and Disassembly by Polarized Resonance Synchronous Spectroscopy. J. Phys. Chem. B 2018, 122, 8429–8438. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, Y.; Li, A.; Huang, Z.-H.; Wang, L.-N.; Kang, F. Porphyrin-Based Nanostructures for Photocatalytic Applications. Nanomaterials 2016, 6, 51. [Google Scholar] [CrossRef]
  38. La, D.D.; Ngo, H.H.; Nguyen, D.D.; Tran, N.T.; Vo, H.T.; Nguyen, X.H.; Chang, S.W.; Chung, W.J.; Nguyen, M.D. Advances and prospects of porphyrin-based nanomaterials via self-assembly for photocatalytic applications in environmental treatment. Coord. Chem. Rev. 2022, 463, 214543. [Google Scholar] [CrossRef]
  39. Shee, N.K.; Kim, H.-J. Sn(IV)-Porphyrin-Based Nanostructures Featuring Pd(II)-Mediated Supramolecular Arrays and Their Photocatalytic Degradation of Acid Orange 7 Dye. Int. J. Mol. Sci. 2022, 23, 13702. [Google Scholar] [CrossRef]
  40. Shee, N.K.; Kim, H.-J. Morphology-controlled self-assembled nanostructures of complementary metalloporphyrin triads through intermolecular coordination tuning and their photocatalytic degradation for Orange II. Inorg. Chem. Front. 2022, 9, 5538–5548. [Google Scholar] [CrossRef]
  41. Lehn, J.-M. Supramolecular chemistry-scope and perspectives molecules, supermolecules, and molecular devices (Nobel Lecture). Angew. Chem. Int. Ed. Engl. 1988, 27, 89–112. [Google Scholar] [CrossRef]
  42. Medforth, C.J.; Wang, Z.; Martin, K.E.; Song, Y.; Jacobsen, J.L.; Shelnutt, J.A. Self-assembled porphyrin nanostructures. Chem. Commun. 2009, 29, 7261–7277. [Google Scholar] [CrossRef] [PubMed]
  43. Durot, S.; Taesch, J.; Heitz, V. Multiporphyrinic Cages: Architectures and Functions. Chem. Rev. 2014, 114, 8542–8578. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, S.-P.; Lin, W.; Wang, X.; Cen, T.-Y.; Xie, H.; Huang, J.; Zhu, B.-Y.; Zhang, Z.; Song, A.; Hao, J.; et al. Controllable hierarchical self-assembly of porphyrin-derived supra-amphiphiles. Nat. Commun. 2019, 10, 1399–1411. [Google Scholar] [CrossRef] [PubMed]
  45. Shee, N.K.; Lee, C.-J.; Kim, H.-J. Crystal structure of bis (benzoato-κO)[5,15-di-phenyl-10,20-bis(pyridin-4-yl)porphyrinato-κ4N,N′,N″,N‴]tin(IV). IUCrData 2019, 4, x190787. [Google Scholar] [CrossRef]
  46. Tian, Y.; Busani, T.; Uyeda, G.H.; Martin, K.E.; van Swol, F.; Medforth, C.J.; Montan, G.A.; Shelnutt, J.A. Hierarchical cooperative binary ionic porphyrin nanocomposites. Chem. Commun. 2012, 48, 4863–4865. [Google Scholar] [CrossRef] [PubMed]
  47. Rodríguez-Abreu, C.; Kolen’ko, Y.V.; Kovnir, K.; Sanchez-Dominguez, M.; Shrestha, R.G.; Bairi, P.; Ariga, K.; Shrestha, L.K. 1D materials from ionic self-assembly in mixtures containing chromonic liquid crystal mesogens. Phys. Chem. Chem. Phys. 2020, 22, 23276–23285. [Google Scholar] [CrossRef]
  48. Gong, X.; Milic, T.; Xu, C.; Batteas, J.D.; Drain, C.M. Preparation and Characterization of Porphyrin Nanoparticles. J. Am. Chem. Soc. 2002, 124, 14290–14291. [Google Scholar] [CrossRef]
  49. Shao, S.; Rajendiran, V.; Lovell, J.F. Metalloporphyrin nanoparticles: Coordinating diverse theranostic functions. Coord. Chem. Rev. 2019, 379, 99–120. [Google Scholar] [CrossRef]
  50. Tian, X.; Lin, C.; Zhong, Z.; Li, X.; Xu, X.; Liu, J.; Kang, L.-T.; Chai, G.; Yao, J. Effect of axial coordination of iron porphyrin on their nanostructures and photocatalytic performance. Cryst. Growth Des. 2019, 19, 3279–3287. [Google Scholar] [CrossRef]
  51. Shee, N.K.; Kim, H.-J. Three Isomeric Zn(II)-Sn(IV)-Zn(II) Porphyrin-Triad-Based Supramolecular Nanoarchitectures for the Morphology-Dependent Photocatalytic Degradation of Methyl Orange. ACS Omega 2022, 7, 9775–9784. [Google Scholar] [CrossRef]
  52. Hasobe, T.; Oki, H.; Sandanayakaa, A.S.D.; Murata, H. Sonication-assisted supramolecular nanorods of meso-diaryl-substituted porphyrins. Chem. Commun. 2008, 6, 724–726. [Google Scholar] [CrossRef] [PubMed]
  53. Dang, L.H.; Vu, M.T.; Chen, J.; Nguyen, C.K.; Bach, L.G.; Tran, N.Q.; Le, V.T. Effect of Ultrasonication on Self-Assembled Nanostructures Formed by Amphiphilic Positive-Charged Copolymers and Negative-Charged Drug. ACS Omega 2019, 4, 4540–4552. [Google Scholar] [CrossRef]
  54. Lu, J.; Li, Z.; An, W.; Liu, L.; Cui, W. Tuning the Supramolecular Structures of Metal-Free Porphyrin via Surfactant Assisted Self-Assembly to Enhance Photocatalytic Performance. Nanomaterials 2019, 9, 1321. [Google Scholar] [CrossRef] [PubMed]
  55. Tashiro, K.; Murafuji, T.; Sumimoto, M.; Fujitsuka, M.; Yamazaki, S. The formation mechanism of ZnTPyP fibers fabricated by a surfactant-assisted method. New J. Chem. 2020, 44, 13824–13833. [Google Scholar] [CrossRef]
  56. Hu, J.S.; Guo, Y.G.; Liang, H.P.; Wan, L.J.; Jiang, L. Three-dimensional self-organization of supramolecular self-assembled porphyrin hollow hexagonal nanoprisms. J. Am. Chem. Soc. 2005, 127, 17090–17095. [Google Scholar] [CrossRef] [PubMed]
  57. Lee, S.J.; Malliakas, C.D.; Kanatzidis, M.G.; Hupp, J.T.; Nguyen, S.T. Amphiphilic porphyrin nanocrystals: Morphology tuning and hierarchical assembly. Adv. Mater. 2008, 20, 3543–3549. [Google Scholar] [CrossRef]
  58. Wang, Q.; Chen, Y.; Ma, P.; Lu, J.; Zhang, X.; Jiang, J. Morphology and chirality controlled self-assembled nanostructures of porphyrin–pentapeptide conjugate: Effect of the peptide secondary conformation. J. Mater. Chem. 2011, 21, 8057–8065. [Google Scholar] [CrossRef]
  59. Song, F.; Jia, J.; Yu, J.; Chen, C.; Feng, J.; Zhu, P. Morphology Controlled Nano-Structures of a Porphyrin Dendrimer Complex: Solvent Effect on the Self-Assembly Behavior. Inorg. Chem. Commun. 2015, 61, 149–151. [Google Scholar] [CrossRef]
  60. Chen, C.; Li, D.; Cao, G.; Qin, Z.; Xu, Y.; Liu, X.; Li, Q.; Xie, Y.; Li, C. Solvent-regulated biomorphs from the intense π,π-mediated assemblies of tetracenequinone fused porphyrin. CrystEngComm 2021, 23, 7565–7569. [Google Scholar] [CrossRef]
  61. Nikoloudakis, E.; Pigiaki, M.; Polychronaki, M.N.; Margaritopoulou, A.; Charalambidis, G.; Serpetzoglou, E.; Mitraki, A.; Loukakos, P.A.; Coutsolelos, A.G. Self-Assembly of Porphyrin Dipeptide Conjugates toward Hydrogen Production. ACS Sustain. Chem. Eng. 2021, 9, 7781–7791. [Google Scholar] [CrossRef]
  62. Wang, Z.; Medforth, C.J.; Shelnutt, J.A. Porphyrin nanotubes by ionic self-assembly. J. Am. Chem. Soc. 2004, 126, 15954–15955. [Google Scholar] [CrossRef] [PubMed]
  63. Tian, Y.; Martin, K.E.; Shelnutt, J.Y.T.; Evans, L.; Busani, T.; Miller, J.E.; Medforth, C.J.; Shelnutt, J.A. Morphological families of self-assembled porphyrin structures and their photosensitization of hydrogen generation. Chem. Commun. 2011, 47, 6069–6071. [Google Scholar] [CrossRef] [PubMed]
  64. Li, C.; Park, K.-M.; Kim, H.-J. Ionic assembled hybrid nanoparticle consisting of tin(IV) porphyrin cations and polyoxomolybdate anions, and photocatalytic hydrogen production by its visible light sensitization. Inorg. Chem. Commun. 2015, 60, 8–11. [Google Scholar] [CrossRef]
  65. Koposova, E.A.; Offenhäusser, A.; Ermolenko, Y.E.; Mourzina, Y.G. Photoresponsive Porphyrin Nanotubes of Meso-tetra(4-Sulfonatophenyl)Porphyrin and Sn(IV) meso-tetra(4-pyridyl)porphyrin. Front. Chem. 2019, 7, 351. [Google Scholar] [CrossRef] [PubMed]
  66. Rebelo, S.L.; Neves, C.M.; de Almeida, M.P.; Pereira, E.; Simões, M.M.; Neves, M.G.P.; de Castro, B.; Medforth, C.J. Binary ionic iron(III) porphyrin nanostructured materials with catalase-like activity. Appl. Mater. Today 2020, 21, 100830. [Google Scholar] [CrossRef]
  67. Guo, P.; Chen, P.; Liu, M. Porphyrin assemblies via a surfactant-assisted method: From nanospheres to nanofibers with tunable length. Langmuir 2012, 28, 15482–15490. [Google Scholar] [CrossRef]
  68. Mandal, S.; Nayak, S.K.; Mallampalli, S.; Patra, A. Surfactant-assisted porphyrin based hierarchical nano/micro assemblies and their efficient photocatalytic behavior. ACS Appl. Mater. Inter. 2013, 6, 130–136. [Google Scholar] [CrossRef]
  69. Zhong, Y.; Wang, Z.; Zhang, R.; Bai, F.; Wu, H.; Haddad, R.; Fan, H. Interfacial Self-Assembly Driven Formation of Hierarchically Structured Nanocrystals with Photocatalytic Activity. ACS Nano 2014, 8, 827–833. [Google Scholar] [CrossRef]
  70. Zhang, N.; Wang, L.; Wang, H.; Cao, R.; Wang, J.; Bai, F.; Fan, H. Self-Assembled One-Dimensional Porphyrin Nanostructures with Enhanced Photocatalytic Hydrogen Generation. Nano Lett. 2018, 18, 560–566. [Google Scholar] [CrossRef]
  71. Cao, R.; Wang, G.; Ren, X.; Duan, P.-C.; Wang, L.; Li, Y.; Chen, X.; Zhu, R.; Jia, Y.; Bai, F. Self-Assembled Porphyrin Nanoleaves with Unique Crossed Transportation of Photogenerated Carriers to Enhance Photocatalytic Hydrogen Production. Nano Lett. 2022, 22, 157–163. [Google Scholar] [CrossRef] [PubMed]
  72. Gao, Y.; Zhang, X.; Ma, C.; Li, X.; Jiang, J. Morphology-Controlled Self-Assembled Nanostructures of 5,15-Di [4-(5-Acetylsulfanylpentyloxy)Phenyl]Porphyrin Derivatives. Effect of Metal-Ligand Coordination Bonding on Tuning the Intermolecular Interaction. J. Am. Chem. Soc. 2008, 130, 17044–17052. [Google Scholar] [CrossRef] [PubMed]
  73. Lu, G.; Zhang, X.; Gai, X.; Jiang, J. Tuning the morphology of self-assembled nanostructures of amphiphilic tetra(p-hydroxyphenyl)porphyrins with hydrogen bonding and metal–ligand coordination bonding. J. Mater. Chem. 2009, 19, 2417–2424. [Google Scholar] [CrossRef]
  74. Zhao, Q.; Wang, Y.; Xu, Y.; Yan, Y.; Huang, J. Out-of-Plane Coordinated Porphyrin Nanotubes with Enhanced Singlet Oxygen Generation Efficiency. Sci. Rep. 2016, 6, 31339. [Google Scholar] [CrossRef]
  75. Shee, N.K.; Kim, M.K.; Kim, H.-J. Supramolecular porphyrin nanostructures based on coordination driven self-assembly and their visible light catalytic degradation of methylene blue dye. Nanomaterials 2020, 10, 2314. [Google Scholar] [CrossRef] [PubMed]
  76. Shee, N.K.; Kim, H.-J. Coordination framework materials fabricated by the self-assembly of Sn(IV) porphyrins with Ag(I) ions for the photocatalytic degradation of organic dyes in wastewater. Inorg. Chem. Front. 2022, 9, 1270–1280. [Google Scholar] [CrossRef]
  77. Yoon, S.M.; Hwang, I.-C.; Kim, K.S.; Choi, H.C. Synthesis of Single-Crystal Tetra(4-pyridyl)porphyrin Rectangular Nanotubes in the Vapor Phase. Angew. Chem. Int. Ed. 2009, 48, 2506–2509. [Google Scholar] [CrossRef]
  78. D’Urso, A.; Fragalà, M.E.; Purrello, R. From self-assembly to noncovalent synthesis of programmable porphyrins’ arrays in aqueous solution. Chem. Commun. 2012, 48, 8165–8176. [Google Scholar] [CrossRef]
  79. Bodedla, G.B.; Huang, J.; Wong, W.-Y.; Zhu, X. Self-Assembled Naphthalimide-Substituted Porphyrin Nanowires for Photocatalytic Hydrogen Evolution. ACS Appl. Nano Mater. 2020, 3, 7040–7046. [Google Scholar] [CrossRef]
  80. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  81. Ning, L.; Xu, J.; Lou, Y.; Pan, C.; Wang, Z.; Zhu, Y. A 3D/0D cobalt-embedded nitrogen-doped porous carbon/supramolecular porphyrin magnetic-separation photocatalyst with highly efficient pollutant degradation and water oxidation performance. J. Mater. Sci. Technol. 2022, 124, 53–64. [Google Scholar] [CrossRef]
  82. Wang, X.; Tang, W.; Jiang, L.; Feng, J.; Yang, J.; Zhou, S.; Li, W.; Yuan, X.; Wang, H.; Wang, J. Mechanism insights into visible light-induced crystalline carbon nitride activating periodate for highly efficient ciprofloxacin removal. Chem. Eng. J. 2023, 471, 144521. [Google Scholar] [CrossRef]
  83. Wang, J.; Zhong, Y.; Wang, L.; Zhang, N.; Cao, R.; Bian, K.; Alarid, L.; Haddad, R.E.; Bai, F.; Fan, H. Morphology-Controlled Synthesis and Metalation of Porphyrin Nanoparticles with Enhanced Photocatalytic Performance. Nano Lett. 2016, 16, 6523–6528. [Google Scholar] [CrossRef]
  84. Jadhav, R.W.; La, D.D.; Truong, T.N.; Khalap, S.V.; Quang, D.V.; Bhosale, S.V. The controllable nanostructure and photocatalytic behaviour of 5, 10, 15, 20-tetra-(3, 4, 5 trimethoxyphenyl) porphyrin through solvophobic supramolecular self-assembly. New J. Chem. 2020, 44, 18442–18448. [Google Scholar] [CrossRef]
  85. Shee, N.K.; Kim, H.-J. Self-Assembled Nanomaterials Based on Complementary Sn(IV) and Zn(II)-Porphyrins, and Their Photocatalytic Degradation for Rhodamine B Dye. Molecules 2021, 26, 3598. [Google Scholar] [CrossRef] [PubMed]
  86. Shee, N.K.; Kim, H.-J. Sn(IV) Porphyrin-Based Ionic Self-Assembled Nanostructures and Their Application in Visible Light Photo-Degradation of Malachite Green. Catalysts 2022, 12, 799. [Google Scholar] [CrossRef]
  87. Shee, N.K.; Kim, H.-J. Supramolecular squares of Sn(IV)porphyrins with Re(I)-corners for the fabrication of self-assembled nanostructures performing photocatalytic degradation of Eriochrome Black T dye. Inorg. Chem. Front. 2022, 10, 174–183. [Google Scholar] [CrossRef]
  88. Zhou, X.-T.; Ji, H.-B.; Huang, X.-J. Photocatalytic Degradation of Methyl Orange over Metalloporphyrins Supported on TiO2 Degussa P25. Molecules 2012, 17, 1149–1158. [Google Scholar] [CrossRef]
  89. Yoo, H.-Y.; Yan, S.; Ra, J.W.; Jeon, D.; Goh, B.; Kim, T.-Y.; Mackeyev, Y.; Ahn, Y.-Y.; Kim, H.-J.; Wilson, L.J.; et al. Tin porphyrin immobilization significantly enhances visible-light-photosensitized degradation of Microcystins: Mechanistic implications. Appl. Catal. B Environ. 2016, 199, 33–44. [Google Scholar] [CrossRef]
  90. La, D.D.; Hangarge, R.V.; Bhosale, S.V.; Ninh, H.D.; Jones, L.A.; Bhosale, S.V. Arginine-Mediated Self-Assembly of Porphyrin on Graphene: A Photocatalyst for Degradation of Dyes. Appl. Sci. 2017, 7, 643. [Google Scholar] [CrossRef]
  91. Min, K.S.; Kumar, R.S.; Lee, J.H.; Kim, K.S.; Lee, S.G.; Son, Y.A. Synthesis of New TiO2/Porphyrin-Based Composites and Photocatalytic Studies on Methylene Blue Degradation. Dye. Pigment. 2019, 160, 37–47. [Google Scholar] [CrossRef]
  92. Mota, H.P.; Quadrado, R.N.; Iglesias, B.A.; Fajardo, A.R. Enhanced photocatalytic degradation of organic pollutants mediated by Zn (II)-porphyrin/poly(acrylic acid) hybrid microparticles. Appl. Catal. B 2020, 277, 119208. [Google Scholar] [CrossRef]
  93. Vo, H.T.; Nguyen, A.T.; Tran, C.V.; Nguyen, S.X.; Tung, N.T.; Pham, D.T.; Nguyen, D.D.; La, D.D. Self-Assembly of Porphyrin Nanofibers on ZnO Nanoparticles for the Enhanced Photocatalytic Performance for Organic Dye Degradation. ACS Omega 2021, 6, 23203–23210. [Google Scholar] [CrossRef]
  94. Xu, J.; Gao, Q.Z.; Wang, Z.P.; Zhu, Y. An all-organic 0D/2D supramolecular porphyrin/g-C3N4 heterojunction assembled via π-π interaction for efficient visible photocatalytic oxidation. Appl. Catal. B Environ. 2021, 291, 120059. [Google Scholar] [CrossRef]
  95. Shee, N.K.; Park, B.-H.; Kim, H.-J. Hybrid Composite of Sn(IV)-Porphyrin and Mesoporous Structure for Enhanced Visible Light Photocatalytic Degradation of Organic Dyes. Molecules 2023, 28, 1886. [Google Scholar] [CrossRef] [PubMed]
  96. Shee, N.K.; Kim, H.-J. Sn(IV)porphyrin-Anchored TiO2 Nanoparticles via Axial-Ligand Coordination for Enhancement of Visible Light-Activated Photocatalytic Degradation. Inorganics 2023, 11, 336. [Google Scholar] [CrossRef]
  97. Shee, N.K.; Kim, H.-J. Surface Modification of ZnO with Sn(IV)-Porphyrin for Enhanced Visible Light Photocatalytic Degradation of Amaranth Dye. Molecules 2023, 28, 6481. [Google Scholar] [CrossRef] [PubMed]
  98. Trogadas, P.; Coppens, M.-O. Nature-Inspired Electrocatalysts and Devices for Energy Conversion. Chem. Soc. Rev. 2020, 49, 3107–3141. [Google Scholar] [CrossRef]
  99. Magna, G.; Mandoj, F.; Stefanelli, M.; Pomarico, G.; Monti, D.; Di Natale, C.; Paolesse, R.; Nardis, S. Recent Advances in Chemical Sensors Using Porphyrin-Carbon Nanostructure Hybrid Materials. Nanomaterials 2021, 11, 997. [Google Scholar] [CrossRef]
  100. O’Neill, J.S.; Kearney, L.; Brandon, M.P.; Pryce, M.T. Design Components of Porphyrin-Based Photocatalytic Hydrogen Evolution Systems: A Review. Coord. Chem. Rev. 2022, 467, 214599. [Google Scholar] [CrossRef]
  101. Rabiee, N.; Yaraki, M.T.; Garakani, S.M.; Garakani, S.M.; Ahmadi, S.; Lajevardi, A.; Bagherzadeh, M.; Rabiee, M.; Tayebi, L.; Tahriri, M. Recent advances in porphyrin-based nanocomposites for effective targeted imaging and therapy. Biomaterials 2020, 232, 119707. [Google Scholar] [CrossRef] [PubMed]
  102. Gotico, P.; Halime, Z.; Aukauloo, A. Recent Advances in Metalloporphyrin-Based Catalyst Design Towards Carbon Dioxide Reduction: From Bio-Inspired Second Coordination Sphere Modifications to Hierarchical Architectures. Dalton Trans. 2020, 49, 2381–2396. [Google Scholar] [CrossRef] [PubMed]
  103. Montaseri, H.; Kruger, C.A.; Abrahamse, H. Recent Advances in Porphyrin-Based Inorganic Nanoparticles for Cancer Treatment. Int. J. Mol. Sci. 2020, 21, 3358. [Google Scholar] [CrossRef] [PubMed]
  104. Limosani, F.; Remita, H.; Tagliatesta, P.; Bauer, E.M.; Leoni, A.; Carbone, M. Functionalization of Gold Nanoparticles with Ru-Porphyrin and Their Selectivity in the Oligomerization of Alkynes. Materials 2022, 15, 1207. [Google Scholar] [CrossRef] [PubMed]
  105. Shee, N.K.; Lee, C.-J.; Kim, H.-J. Hexacoordinated Sn(IV) porphyrin-based square-grid frameworks exhibiting selective uptake of CO2 over N2. Bull. Korean Chem. Soc. 2022, 43, 103–109. [Google Scholar] [CrossRef]
Figure 1. Morphology of ZnTPyP: SEM (A,B) and TEM (C). Reproduced from Ref. [56].
Figure 1. Morphology of ZnTPyP: SEM (A,B) and TEM (C). Reproduced from Ref. [56].
Molecules 29 00611 g001
Figure 2. SEM image of tin dipyridylporphyrin under various conditions: room temperature (a) and 90 °C (b). Adapted from Ref. [57].
Figure 2. SEM image of tin dipyridylporphyrin under various conditions: room temperature (a) and 90 °C (b). Adapted from Ref. [57].
Molecules 29 00611 g002
Figure 3. Self-assembled nanostructures derived from n-hexane and THF (3:1): HRTEM (A) and SEM (B) of nanofibers. Nanostructures fabricated from THF and water (1:3): HRTEM (C) including an enlargement (c) and SEM (D) including an enlargement (d) of nanotubes. Adapted from Ref. [58].
Figure 3. Self-assembled nanostructures derived from n-hexane and THF (3:1): HRTEM (A) and SEM (B) of nanofibers. Nanostructures fabricated from THF and water (1:3): HRTEM (C) including an enlargement (c) and SEM (D) including an enlargement (d) of nanotubes. Adapted from Ref. [58].
Molecules 29 00611 g003
Figure 4. Solvent-induced SEM image of ZnP2: n-hexane (A) and methanol (B). Adapted from Ref. [59].
Figure 4. Solvent-induced SEM image of ZnP2: n-hexane (A) and methanol (B). Adapted from Ref. [59].
Molecules 29 00611 g004
Figure 5. Solvent-regulated SEM image of Ni4TQ microstructures acquired in (a) MeOH: toluene (9:1) and (b) MeOH: toluene (8.5:1.5). Adapted from Ref. [60].
Figure 5. Solvent-regulated SEM image of Ni4TQ microstructures acquired in (a) MeOH: toluene (9:1) and (b) MeOH: toluene (8.5:1.5). Adapted from Ref. [60].
Molecules 29 00611 g005
Figure 6. SEM images of nanostructures derived from different solvents: (a) DCM/MeOH (2:8), (b) MeOH/HFIP (8:2), (c) heptane/DCM (8:2), and (d) MeOH/THF (8:2). Adapted from Ref. [61].
Figure 6. SEM images of nanostructures derived from different solvents: (a) DCM/MeOH (2:8), (b) MeOH/HFIP (8:2), (c) heptane/DCM (8:2), and (d) MeOH/THF (8:2). Adapted from Ref. [61].
Molecules 29 00611 g006
Figure 7. TEM images of ionic self-assembled nanostructures fabricated from cationic Sn(OH)2TPyP4+ and anionic H4TPPS42−. Adapted from Ref. [62].
Figure 7. TEM images of ionic self-assembled nanostructures fabricated from cationic Sn(OH)2TPyP4+ and anionic H4TPPS42−. Adapted from Ref. [62].
Molecules 29 00611 g007
Figure 8. Morphology of ionic self-assembled nanomaterials constructed from cationic ZnT(N-EtOH-4-Py)P and anionic ZnTPPS at several temperatures. (a) 10 °C, (b) 23 °C, (c) 60 °C, (d) 80 °C. Adapted from Ref. [63].
Figure 8. Morphology of ionic self-assembled nanomaterials constructed from cationic ZnT(N-EtOH-4-Py)P and anionic ZnTPPS at several temperatures. (a) 10 °C, (b) 23 °C, (c) 60 °C, (d) 80 °C. Adapted from Ref. [63].
Molecules 29 00611 g008
Figure 9. Morphology (SEM) of nanostructures fabricated from cationic tin(IV)porphyrin and anionic polyoxomolybdate. Adapted from Ref. [64].
Figure 9. Morphology (SEM) of nanostructures fabricated from cationic tin(IV)porphyrin and anionic polyoxomolybdate. Adapted from Ref. [64].
Molecules 29 00611 g009
Figure 10. Morphology of fabricated nanoaggregates obtained from the combination of [SnCl2TPyHP]4+ with H4TPPS42− in solution at pH = 2.0. Concentration ratio of [SnCl2TPyHP]4+: H4TPPS42−; 1:1 (A,B), 1:5 (C,D), and 5:1 (E,F). Adapted from Ref. [65].
Figure 10. Morphology of fabricated nanoaggregates obtained from the combination of [SnCl2TPyHP]4+ with H4TPPS42− in solution at pH = 2.0. Concentration ratio of [SnCl2TPyHP]4+: H4TPPS42−; 1:1 (A,B), 1:5 (C,D), and 5:1 (E,F). Adapted from Ref. [65].
Molecules 29 00611 g010
Figure 11. SEM images of ionic self-assembled nanostructures constructed from the combination of anionic FeTSPP with various cationic iron porphyrins in solution: (a,c,e) at pH = 1. (b,d,f) at pH = 2. Adapted from Ref. [66].
Figure 11. SEM images of ionic self-assembled nanostructures constructed from the combination of anionic FeTSPP with various cationic iron porphyrins in solution: (a,c,e) at pH = 1. (b,d,f) at pH = 2. Adapted from Ref. [66].
Molecules 29 00611 g011
Figure 12. SEM images of the formed porphyrin nanoarchitectures from ZnTPyP with different aging times: (A) nanospheres (t = 0), (B) heterostructures (t = 5 h), (C) nanofibers (t = 10 days). Adapted from Ref. [67].
Figure 12. SEM images of the formed porphyrin nanoarchitectures from ZnTPyP with different aging times: (A) nanospheres (t = 0), (B) heterostructures (t = 5 h), (C) nanofibers (t = 10 days). Adapted from Ref. [67].
Molecules 29 00611 g012
Figure 13. SEM images of various morphologies as a function of different stirring times. (a) 15 min, (b) 60 min, (c) 6 h, (d) 48 h. Adapted from Ref. [68].
Figure 13. SEM images of various morphologies as a function of different stirring times. (a) 15 min, (b) 60 min, (c) 6 h, (d) 48 h. Adapted from Ref. [68].
Molecules 29 00611 g013
Figure 14. Different surfactant-assisted morphologies as a function of surfactant. SEM (a) and TEM (b) morphologies of octahedral arrays from CTAB. SEM (c) and TEM (d) morphologies of nanosheets from CATB and emulsion. SEM (e) and TEM (f) morphologies of nanospheres from SDS. Adapted from Ref. [69].
Figure 14. Different surfactant-assisted morphologies as a function of surfactant. SEM (a) and TEM (b) morphologies of octahedral arrays from CTAB. SEM (c) and TEM (d) morphologies of nanosheets from CATB and emulsion. SEM (e) and TEM (f) morphologies of nanospheres from SDS. Adapted from Ref. [69].
Molecules 29 00611 g014
Figure 15. Surfactant-assisted nanowire morphology of H2THPP. SEM (A) and TEM (B). Adapted from Ref. [70].
Figure 15. Surfactant-assisted nanowire morphology of H2THPP. SEM (A) and TEM (B). Adapted from Ref. [70].
Molecules 29 00611 g015
Figure 16. Surfactant-assisted PdTCPP nanoleaves from DASS (15 mM) and porphyrin (1 mM). Morphology: (a) SEM and (b) TEM. Adapted from Ref. [71].
Figure 16. Surfactant-assisted PdTCPP nanoleaves from DASS (15 mM) and porphyrin (1 mM). Morphology: (a) SEM and (b) TEM. Adapted from Ref. [71].
Molecules 29 00611 g016
Figure 17. SEM images of porphyrin nanostructures: (A,C) in methanol, (B,D) in n-hexane. Zoom-in image of (B,D). Adapted from Ref. [72].
Figure 17. SEM images of porphyrin nanostructures: (A,C) in methanol, (B,D) in n-hexane. Zoom-in image of (B,D). Adapted from Ref. [72].
Molecules 29 00611 g017
Figure 18. SEM images of porphyrin nanostructures derived from water; TEM (A,C) and SEM (B,D). Zoom-in image of (A,D). Adapted from Ref. [73].
Figure 18. SEM images of porphyrin nanostructures derived from water; TEM (A,C) and SEM (B,D). Zoom-in image of (A,D). Adapted from Ref. [73].
Molecules 29 00611 g018
Figure 19. SEM images of porphyrin nanostructures derived from TPPS44− with the addition of Bi(NO3)3 in water (pH = 3.6). SEM; low (a) and high (b). TEM; low (c) and high (d). [Bi3+] = 0.2 mM, [TPPS44−] = 0.1 mM. Adapted from Ref. [74].
Figure 19. SEM images of porphyrin nanostructures derived from TPPS44− with the addition of Bi(NO3)3 in water (pH = 3.6). SEM; low (a) and high (b). TEM; low (c) and high (d). [Bi3+] = 0.2 mM, [TPPS44−] = 0.1 mM. Adapted from Ref. [74].
Molecules 29 00611 g019
Figure 20. SEM images of porphyrin nanostructures derived from the drop casting of a toluene solution of porphyrin triad molecules onto copper tapes. M = 2H, (a); M = Zn, (b). Adapted from Ref. [75].
Figure 20. SEM images of porphyrin nanostructures derived from the drop casting of a toluene solution of porphyrin triad molecules onto copper tapes. M = 2H, (a); M = Zn, (b). Adapted from Ref. [75].
Molecules 29 00611 g020
Figure 21. SEM images of porphyrin nanostructures derived from the addition of AgOTf to a solution of Sn(IV)porphyrin precursors. Without axial ligation (a,b), with axial ligation (c,d). Adapted from Ref. [76].
Figure 21. SEM images of porphyrin nanostructures derived from the addition of AgOTf to a solution of Sn(IV)porphyrin precursors. Without axial ligation (a,b), with axial ligation (c,d). Adapted from Ref. [76].
Molecules 29 00611 g021
Figure 22. SEM images of sonication-assisted porphyrin nanostructures. Adapted from Ref. [52].
Figure 22. SEM images of sonication-assisted porphyrin nanostructures. Adapted from Ref. [52].
Molecules 29 00611 g022
Figure 23. SEM images of nanotubes of H2TPyP prepared using a vaporization–condensation–recrystallization method. Cross-section (a), high resolution (b). Adapted from Ref. [77].
Figure 23. SEM images of nanotubes of H2TPyP prepared using a vaporization–condensation–recrystallization method. Cross-section (a), high resolution (b). Adapted from Ref. [77].
Molecules 29 00611 g023
Figure 24. Template-assisted self-assembly of porphyrin nanoaggregates. Adapted from Ref. [78].
Figure 24. Template-assisted self-assembly of porphyrin nanoaggregates. Adapted from Ref. [78].
Molecules 29 00611 g024
Figure 25. The morphology of the nanowire is derived from ZnD(p-NI)PP. SEM (a,b) and TEM (c) Adapted from Ref. [79].
Figure 25. The morphology of the nanowire is derived from ZnD(p-NI)PP. SEM (a,b) and TEM (c) Adapted from Ref. [79].
Molecules 29 00611 g025
Figure 26. (I) Morphology-dependent photocatalytic activity of different H2TCPP aggregates; (II) degradation kinetics; (III) fluorescence decay plots: (a) nanosphere, (b) nanorod, (c) nanoflower, and (d) monomeric H2TCPP. Adapted from Ref. [68].
Figure 26. (I) Morphology-dependent photocatalytic activity of different H2TCPP aggregates; (II) degradation kinetics; (III) fluorescence decay plots: (a) nanosphere, (b) nanorod, (c) nanoflower, and (d) monomeric H2TCPP. Adapted from Ref. [68].
Molecules 29 00611 g026
Figure 27. Photodegradation of MO dye by various SnCl2TPP nanostructures. No catalyst (a), TiO2 (P-25) (b), nanosheets (c), octahedra (d), and nanospheres (e). Adapted from Ref. [69].
Figure 27. Photodegradation of MO dye by various SnCl2TPP nanostructures. No catalyst (a), TiO2 (P-25) (b), nanosheets (c), octahedra (d), and nanospheres (e). Adapted from Ref. [69].
Molecules 29 00611 g027
Figure 28. Photodegradation of MO dye by various nanostructures. No catalyst (a), (b) H2TPyP nano octahedra. (c) Intermediate nanoparticles. (d) ZnTPyP nanowires. Adapted from Ref. [83].
Figure 28. Photodegradation of MO dye by various nanostructures. No catalyst (a), (b) H2TPyP nano octahedra. (c) Intermediate nanoparticles. (d) ZnTPyP nanowires. Adapted from Ref. [83].
Molecules 29 00611 g028
Figure 29. Photodegradation of RhB dye by different morphologies of TTOP fabricated using THF with different H2O fractions of 70% (a), 80% (b), 90% (c), and 95% (d). Adapted from Ref. [84].
Figure 29. Photodegradation of RhB dye by different morphologies of TTOP fabricated using THF with different H2O fractions of 70% (a), 80% (b), 90% (c), and 95% (d). Adapted from Ref. [84].
Molecules 29 00611 g029
Figure 30. Photocatalytic degradation of MB dye by triads with different morphologies in the presence of visible light (a) and first-order rate kinetics (b). Adapted from Ref. [75].
Figure 30. Photocatalytic degradation of MB dye by triads with different morphologies in the presence of visible light (a) and first-order rate kinetics (b). Adapted from Ref. [75].
Molecules 29 00611 g030
Figure 31. Photodegradation of RhB dye by various morphologies obtained from metalloporphyrin-based triads under visible-light irradiation. T1 (1), T2 (2), T3 (3), T4 (4), T5 (5), and T6 (6). Adapted from Ref. [85].
Figure 31. Photodegradation of RhB dye by various morphologies obtained from metalloporphyrin-based triads under visible-light irradiation. T1 (1), T2 (2), T3 (3), T4 (4), T5 (5), and T6 (6). Adapted from Ref. [85].
Molecules 29 00611 g031
Figure 32. Morphology-controlled degradation of MO dye by several metalloporphyrin-based triads under visible-light irradiation. ZnL (a), T2 (b), T3 (c), and T4 (d). Adapted from Ref. [51].
Figure 32. Morphology-controlled degradation of MO dye by several metalloporphyrin-based triads under visible-light irradiation. ZnL (a), T2 (b), T3 (c), and T4 (d). Adapted from Ref. [51].
Molecules 29 00611 g032
Figure 33. Morphology-controlled photodegradation of MG dye by various ionic complexes. Adapted from Ref. [86].
Figure 33. Morphology-controlled photodegradation of MG dye by various ionic complexes. Adapted from Ref. [86].
Molecules 29 00611 g033
Figure 34. Morphology-controlled photodegradation of EBT dye by 2 and 3. Adapted from Ref. [87].
Figure 34. Morphology-controlled photodegradation of EBT dye by 2 and 3. Adapted from Ref. [87].
Molecules 29 00611 g034
Figure 35. Morphology of TiO2 (a) and CuP-TiO2 (b). Photocatalytic degradation of MO dye (c). Adapted from Ref. [88].
Figure 35. Morphology of TiO2 (a) and CuP-TiO2 (b). Photocatalytic degradation of MO dye (c). Adapted from Ref. [88].
Molecules 29 00611 g035
Figure 36. Photochemical degradation of RR-MC by silica/SnP-based photosensitizers (inset: photodegradation of 2,4,6-trimethylphenol) and heterogeneous sanitizers (RB, aminoC60, s-SnP). Reproduced from Ref. [89].
Figure 36. Photochemical degradation of RR-MC by silica/SnP-based photosensitizers (inset: photodegradation of 2,4,6-trimethylphenol) and heterogeneous sanitizers (RB, aminoC60, s-SnP). Reproduced from Ref. [89].
Molecules 29 00611 g036
Figure 37. Morphology (SEM) of (A) GNPs, (B) free-standing H2TCPP nanoaggregates, and (C) GNPs@H2TCPP. Photocatalytic performance (D) and kinetic simulation curve (E) for MO dye; (a) blank, (b) GNPs, (c) H2TCPP nanoaggregates (free-standing), (d) GNPs@H2TCPP. Adapted from Ref. [90].
Figure 37. Morphology (SEM) of (A) GNPs, (B) free-standing H2TCPP nanoaggregates, and (C) GNPs@H2TCPP. Photocatalytic performance (D) and kinetic simulation curve (E) for MO dye; (a) blank, (b) GNPs, (c) H2TCPP nanoaggregates (free-standing), (d) GNPs@H2TCPP. Adapted from Ref. [90].
Molecules 29 00611 g037
Figure 38. Various TiO2-porphyrin photocatalysts for photodegradation of MB dye. Adapted from Ref. [91].
Figure 38. Various TiO2-porphyrin photocatalysts for photodegradation of MB dye. Adapted from Ref. [91].
Molecules 29 00611 g038
Figure 39. Morphology (SEM) of (a) PAA and (b) Zn(II)Pr@PAA. Photodegradation of (c) MB and (d) NB. Adapted from Ref. [92].
Figure 39. Morphology (SEM) of (a) PAA and (b) Zn(II)Pr@PAA. Photodegradation of (c) MB and (d) NB. Adapted from Ref. [92].
Molecules 29 00611 g039
Figure 40. Morphology of ZnO-NPs: (a) SEM, (c) TEM. Morphology of ZnO@H2TCPP nanofibers: (b) SEM, (d) TEM. (e) Degradation of RhB and (f) kinetic study. Adapted from Ref. [93].
Figure 40. Morphology of ZnO-NPs: (a) SEM, (c) TEM. Morphology of ZnO@H2TCPP nanofibers: (b) SEM, (d) TEM. (e) Degradation of RhB and (f) kinetic study. Adapted from Ref. [93].
Molecules 29 00611 g040
Figure 41. Spatial conformation of SA-TCPP/O-CN. Morphology (TEM) of O-CN (a), SA-TCPP (b), and SA-TCPP/O-CN-40% (c). HRTEM of SA-TCPP/O-CN-40% (d). (e) Photodegradation rate constants of organic pollutants. Adapted from Ref. [94].
Figure 41. Spatial conformation of SA-TCPP/O-CN. Morphology (TEM) of O-CN (a), SA-TCPP (b), and SA-TCPP/O-CN-40% (c). HRTEM of SA-TCPP/O-CN-40% (d). (e) Photodegradation rate constants of organic pollutants. Adapted from Ref. [94].
Molecules 29 00611 g041
Figure 42. Synthesis scheme for SiO2@SnP and MCM-41@SnP. Adapted from Ref. [95].
Figure 42. Synthesis scheme for SiO2@SnP and MCM-41@SnP. Adapted from Ref. [95].
Molecules 29 00611 g042
Figure 43. Surface-modified photocatalyst 4 for photocatalytic degradation of RhB dye. Adapted from Ref. [96].
Figure 43. Surface-modified photocatalyst 4 for photocatalytic degradation of RhB dye. Adapted from Ref. [96].
Molecules 29 00611 g043
Figure 44. Surface-modified photocatalysts for photodegradation of AM dye. Adapted from Ref. [97].
Figure 44. Surface-modified photocatalysts for photodegradation of AM dye. Adapted from Ref. [97].
Molecules 29 00611 g044
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

Shee, N.K.; Kim, H.-J. Porphyrin-Based Nanomaterials for the Photocatalytic Remediation of Wastewater: Recent Advances and Perspectives. Molecules 2024, 29, 611. https://doi.org/10.3390/molecules29030611

AMA Style

Shee NK, Kim H-J. Porphyrin-Based Nanomaterials for the Photocatalytic Remediation of Wastewater: Recent Advances and Perspectives. Molecules. 2024; 29(3):611. https://doi.org/10.3390/molecules29030611

Chicago/Turabian Style

Shee, Nirmal Kumar, and Hee-Joon Kim. 2024. "Porphyrin-Based Nanomaterials for the Photocatalytic Remediation of Wastewater: Recent Advances and Perspectives" Molecules 29, no. 3: 611. https://doi.org/10.3390/molecules29030611

Article Metrics

Back to TopTop