Next Article in Journal
Effect of Hydrogen-Donor of Heavy Crude Oil Catalytic Aquathermolysis in the Presence of a Nickel-Based Catalyst
Next Article in Special Issue
Fluoride-Doped TiO2 Photocatalyst with Enhanced Activity for Stable Pollutant Degradation
Previous Article in Journal
Efficient Catalysts of Ethanol Steam Reforming Based on Perovskite-Fluorite Nanocomposites with Supported Ni: Effect of the Synthesis Methods on the Activity and Stability
Previous Article in Special Issue
Zinc–Acetate–Amine Complexes as Precursors to ZnO and the Effect of the Amine on Nanoparticle Morphology, Size, and Photocatalytic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tailoring Structure: Current Design Strategies and Emerging Trends to Hierarchical Catalysts

1
Department of Chemical, Materials and Production Engineering, University of Naples Federico II, Piazzale Tecchio 80, Fuorigrotta, 80125 Naples, Italy
2
Department of Civil, Architectural and Environmental Engineering, University of Naples Federico II, Via Claudio 21, Fuorigrotta, 80125 Naples, Italy
3
CSGI—Center for Colloid and Surface Science, Via Della Lastruccia 3, 50019 Florence, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1152; https://doi.org/10.3390/catal12101152
Submission received: 3 September 2022 / Revised: 22 September 2022 / Accepted: 27 September 2022 / Published: 1 October 2022
(This article belongs to the Special Issue Structured Semiconductors in Photocatalysis)

Abstract

:
Nature mimicking implies the design of nanostructured materials, which can be assembled into a hierarchical structure, thus outperforming the features of the neat components because of their multiple length scale organization. This approach can be effectively exploited for the design of advanced photocatalysts with superior catalytic activity for energy and environment applications with considerable development in the recent six years. In this context, we propose a review on the state of the art for hierarchical photocatalyst production. Particularly, different synthesis strategies are presented, including template-free structuring, and organic, inorganic, and hybrid templating. Furthermore, emerging approaches based on hybrid and bio-waste templating are also highlighted. Finally, a critical comparison among available methods is carried out based on the envisaged application.

1. Introduction

“Learning from Nature” is always a winning strategy for materials scientists. Indeed, natural, and biological systems exhibit a multitude of properties and functions, often tailored for a specific application. Thus, they may be source of imitation for the design of high performance materials, able to meet rising ambitious industrial needs, following a biomimetic approach [1,2,3,4]. What makes bioavailable materials so interesting is that they often share a hierarchical structure made by a self-organization of molecular building blocks that are assembled with other phases, which in turn are self- organized at increasing size levels ranging from the nanoscale up to the macroscale [2,5,6,7]. The complex interplay between structure, morphology, and surface chemistry leads to a wide range of outstanding properties including superhydrophobicity (ex-lotus leaves [8,9]), improved mechanical performances (ex-bird feathers [10,11]), and unique optical behavior (ex-butterfly wings [12,13]). The appealing perspective to translate nature examples into industrial and technological devices steers research towards the synthesis of hierarchical nanomaterials. Indeed, intriguing features of nanostructured systems can be further improved if they are built up into hierarchically structured materials (HSPM), which exhibit a porous architecture, consisting of interconnected pores of different length scales from micro- (<2 nm), meso- (2–50 nm), to macropores (>50 nm), following bimodal or even trimodal pore size distributions [14]. Morphological features such as interconnected multilevel hierarchical porosity, high surface area, and large accessible space are responsible for enhanced light harvesting, electron and ion transport, mass loading, and diffusion [14,15].
Such outstanding properties confer to HSPM a pivotal role in a great number of technological fields including energy storage and conversion [16], catalysis [17] and photocatalysis [18], water remediation [19], separation [20], sensing [21], and biomedicine [22]. Indeed, the variety of physico-chemical characteristics exhibited by these nature-inspired materials can be modified by tuning surface chemistry and morphological properties such as surface area and pore distribution. Notably, a three-dimensional multimodal extended pore structure would be ideal for application in heterogeneous catalysis. The combination of different levels of intraparticle porosity, as well as the presence of mesopores, improve the diffusion of reactants and products towards the active sites and the external surface of the catalyst, respectively, and reduce the length of pore channels, resulting in decreased diffusion barriers in the solid volume [23,24,25,26]. Moreover, the large surface area (from 100 to thousands m2/g [26,27]) allows for better accessibility and uniformity of active sites in the catalysts, causing an improvement in the surface reaction kinetics [15]. Building up a deeply extended multimodal porous architecture in semiconductors paves the way to the design of more efficient photocatalysts [28] by overcoming the typical limits that obstacle their performances such as short lifetime of photogenerated electron-hole pairs [29], partial utilization of solar light [30], low safety, high cost, and poor chemical stability [31]. Widespread application of photocatalysis accounts for huge technological potential of hierarchical photocatalysts, prompting scientific interest towards the design and use of these systems. Accordingly, a great number of outstanding studies has been published on this subject in the last years, which have been overviewed in different reviews [15,32,33,34,35,36,37,38].
Indeed, the number of articles on hierarchical photocatalysts have been constantly growing so far and so has the sensitivity towards environmental energetic issues and sustainability [39,40,41,42,43].
This has prompted the flourishing of several systems produced according to sustainable logics, some of them with exotic structures and unusual compositions, in the attempt to improve photocatalytic performance compared to state of the art [43,44,45,46].
Thus, we believe that a review summarizing the most recent advances and strategies in the design of hierarchical photocatalysts is timely and could enable more conscious development of these systems. First, the advantages of hierarchical photocatalysts are shortly illustrated. Then, the most promising strategies for their fabrication are systematically reviewed focusing on the most outstanding studies published in the period between 2015 and 2021. Particularly, as a strong element of novelty with respect to previous reviews published on this subject, emerging manufacturing approaches are highlighted, including the use of hybrid templates, such as metallorganic frameworks (MOFs). Furthermore, special emphasis is placed on the use of bioavailable structure directing agents, even derived from bio-waste valorization, according to a circular approach. Notably, the mechanism accounting for improved photocatalytic activity has been described to highlight the strength of each synthesis approach and provide scientists involved in this field with clear indication on the key parameters that can be easily manipulated to optimize catalytic performance. To this purpose, modification strategies, including doping and co-doping, the use of co-catalysts, and heterostructures formation have been illustrated. In addition, the use of emerging non-conventional energy sources in the synthesis of hierarchical nanostructures has been shortly overviewed. Furthermore, main applications in environmental and energetic fields, devoted to pollutants removal and hydrogen production, are reviewed and tables reporting collected results on photocatalytic performance were created to enable straight comparison among systems produced using similar or different synthesis strategies. These have been analyzed highlighting pros and cons to get interested scientists more conscious with the opportunities offered by each approach. Moreover, main unaddressed issues and future challenges for both fundamental investigation and technological application have been envisaged. We hope that the review can be useful to stimulate more focused and fruitful research in this field and can open new routes in the development of more effective photocatalytic systems.

2. Hierarchical Photocatalysts

The expression hierarchical photocatalysts usually refers to nanostructured semiconductors having multidimensional domains at different size levels and multimodal pore structure [16].
Since photocatalysis involves a series of complicated interconnected phenomena including light absorption, charge excitation/separation, charge migration, transport and recombination, and charge utilization, the overall efficiency of a photocatalyst is calculated as the product of the individual efficiency of each phenomenon [15]:
ηc = ηabs ∙ηcs ∙ηcmt ∙ηcu,
It appears evident that a decrease in any singular efficiency would contribute to the decay of the overall efficiency. Hierarchical photocatalysts exhibit a series of advantages with respect to traditional photocatalysts, which are summarized in Figure 1 and include higher surface area, enhanced molecular diffusion/transfer, and improved light absorption, among others (Figure 1).
Indeed, properly tuning the pore hierarchy can increase the number of light traveling paths within the catalyst, resulting in multiple reflections as in hollow structures or enhanced light scattering phenomena, thus improving the absorption efficiency (Figure 2) [15,47].
In addition, high surface area enables heterojunction formation and promotes surface reactions as well as charge separation. Furthermore, nanosized building blocks reduce migration distance of charge carriers (Figure 1). To date, pristine and doped TiO2 [28,48,49,50,51,52,53,54,55], ZnO [56,57,58,59,60,61], CeO2 [62,63], and graphene-blended hierarchical photocatalysts [64,65,66] have been synthesized in multiple shapes and morphologies such as nanospheres [67,68], nanoflowers [69,70,71,72,73,74,75,76], nanosheets [77,78,79,80,81], urchin-like nanostructures [68,82,83,84,85,86,87], and nanoflakes [88,89]. The effort to produce efficient hierarchical nanostructured photocatalysts materialized into a heterogeneous framework of synthesis strategies so far, all of them groupable into two macro-categories: templating strategies and template-free methods. Herein, we present an overview of the latest produced hierarchically nanostructured photocatalysts, focusing on the different synthesis strategies which have been implemented to develop them. Our sincere opinion is that this review can shed light on the most promising approaches that could be exploited to design cutting-edge photocatalysts with impressive performance in a huge range of energy and environmental applications, including H2 production, water remediation, and organic pollutant degradation.

Synthesis Approaches to Hierarchical Structures

The main synthesis strategies to hierarchical semiconductors are overviewed in Figure 3 and can be grouped into three wide categories: templating approaches, templated free methods, and post-synthesis approaches.
Templated approaches are the most employed methods to obtain hierarchical materials, offering good reproducibility, large scale application, as well as tailored structures, due to the huge range of available templates. These approaches can be further distinguished into hard templating, soft templating, and bio-templating strategies.
In hard templating methods, also known as nanocasting processes, either inorganic or organic structures with rigid shape are employed as sacrificial templates to build up hierarchical framework [90]. These are subsequently removed through either selective leaching or thermal decomposition to achieve the final porous structure.
As major drawbacks, the use of corrosive and harmful leaching agents or high energy demand hamper large scale applications of these processes.
Contrarily to hard templates, soft templates are highly deformable and more easily removable. They include block copolymers (e.g., Pluronics P123, F127, PEG), microemulsions (micelles and vesicles), ionic liquids [91] hydrogels, surfactants (e.g., sodium dedecyl sulfate (SDS) and cetyl trimethylammonium bromide (CTAB)), and even gas bubbles [92]. On the other hand, bioavailable moieties are emerging as a cheaper and more sustainable alternative.

3. Inorganic Templates

Inorganic template-assisted synthesis is widely used for generation of multi-modal and hierarchical pores in the desired products. Generally, during this synthesis strategy, the catalyst precursors are filled into preexisting templates with suitable pore size, followed by subsequent removal of sacrificial framework thus generating a stable porous structure with uniform pore distribution. The final porous architecture can be controlled tuning the pore size and the wall thickness of hard templates. Most used templates include silica and ice as hard pore-directing agents (Table 1). Worthy of note is a bi-templated method, proposed to produce both SrTiO3/TiO2/C (STC) (Figure 4) [93] and PbTiO3/TiO2/carbon (PTC) [94] heterostructures with a tri-modal micro-/meso-/macro-porosity. This was obtained by combining freeze casting process with the use of silica colloid as the hard template, then followed by an appropriate pyrolysis treatment. Particularly, ice and silica templates were used to obtain macro- and meso-pores, while pyrolysis introduced micropores. The proposed strategy was successful in tuning microstructure of both synthesized STC and PTC architectures leading to excellent photocatalytic performances in the degradation of methylene blue under UV light irradiation. In addition to the organic dye degradation, the photochemical reactivity of the STC samples was further investigated in the photocatalytic hydrogen production from water splitting resulting in a H2 production rate as high as 2.52 mmol h−1 g−1 under UV irradiation.
A similar procedure was also employed to fabricate porous N doped TiO2/C nanocomposites [95]. In this case, the ice/silica hard templates and urea pyrolysis were used to obtain flower-like hierarchical pores at micro-, meso-, and macro-scale and the resulting materials showed highly improved photocatalytic activity for both methylene blue degradation and hydrogen production under both UV and visible light irradiation. Furthermore, this approach was exploited to synthetize other metal oxide systems (TiO2 and ZrO2) [96], and in all cases the corresponding hierarchical carbon-based composites showed superior photocatalytic activity towards methylene blue degradation compared with the control samples. Silica, as mesopore structuring agent, was also exploited to produce hierarchical structure of porous core-shell homojunction constructed by crystalline and amorphous TiO2 [97]. Combining the amorphous TiO2 shell with the mesoporous rutile crystals resulted in a simultaneously enhancement of adsorption ability, removal rate, and mineralization efficiency, under UV irradiation, of Tetracycline hydrochloride (TCH) chosen as antibiotic model. The systems developed are shown in Table 1.

4. Organic Templates

A great number of different templating agents have been proposed for the synthesis of hierarchically porous materials and a widespread attention has been recently focused on the use of organic (Table 2) and/or bioavailable soft templates (Table 3). Biogenic-templated catalyst which owns intrinsic hierarchical, multi-dimensional, and multi-level structure holds promising potency to enhance light-harvesting and photocatalytic performance.
Hierarchically ordered macro–mesoporous anatase TiO2 was successfully produced by Zhao et al. [98] using triblock copolymer P123 and natural pearl oyster shell as dual templates. Inexpensive and environment-friendly pearl oyster shell is exploited as bio-template for the fabrication of macroporous structures and at the same time P123 is employed as a mesopore-directing agent to produce highly ordered mesopores. The authors investigated the obtained materials after calcination at different temperatures and demonstrated that the resultant crystalline anatase structure is made of both macropores and mesopores which are well-preserved after calcination at 350 °C or 450 °C. The photocatalytic activities of the produced samples were monitored through the photodegradation of an aqueous Rhodamine B (RhB) solution under UV irradiation. The highest photocatalytic activity was obtained in the case of TiO2 sample calcined at 450 °C, with a degradation of RhB of about 90%, because of the presence of hierarchical macro–mesoporous structures, high specific surface areas, and the anatase phase. Among bioavailable sources, bio-wastes are emerging as abundant and cheap templates to build up photoactive hierarchical structures, following a waste to wealth approach. Indeed, because of intrinsic lightness, some bio-wastes can drive formation of water floating photocatalysts, which exhibit remarkable advantages over powder suspensions, in terms of efficient recycling as well as high light harvesting efficiency due to their proximity to air-solution interface [99].
As a proof of concept, bio-waste poplar catkin was combined with triazine-based porous organic polymer to obtain a water-floatable photocatalyst for water remediation, exhibiting relevant sunlight activity towards Cr(VI) and methylene blue decontamination (Figure 5) [100].
Among biowastes, rice husk has attracted great research interest because of its unique and hierarchically porous structure with excellent mechanical property and extremely high light-harvesting and exploitation efficiency.
Rice husk was used by Chen et al. [101] as main raw material to form hybrid silica-carbon bio-template to fabricate porous TiO2 anatase.
A detailed study was reported on the influence of calcination temperature and impregnation times on the hierarchical porous structures and photocatalytic activity of the synthesized materials. Porous structures with 1D nanostructures network connection showed efficient light-harvesting and photocatalytic degradation of Rh B (120 mg L−1) which was completely removed in just 180 min. Furthermore, the observed high thermal stability opened new perspectives in the use of these materials in a wide range of applications including photocatalysis, catalysis, solar-cell, separation, and purification processes.
Biotemplated synthesis of hierarchical α-Fe2O3 fibers was proposed by Chen et al. [102], using cotton as biological template. The original structure of cotton was preserved in the final iron oxide fibers which also exhibit certain unexpected magnetic behavior; in fact, common α-Fe2O3 is not magnetic. In addition, superior photocatalytic decolorating performances were observed under visible light irradiation, up to 96.8% methylene blue degraded in only 200 min. Moreover, no loss of catalytic activity was appreciated after three uses, proving high stability. Indeed, this study evidenced the key role of hierarchical porosity in catalytic activity, with fibrous structures acting as effective catalytic centers for organic dye elimination. Recently, electrospinning fiber formation technology has emerged as a powerful and straightforward tool for tailoring material structures and assembling hierarchical architectures [103,104]. There is plenty of scope for improving photocatalytic properties if this methodology is combined with more conventional processes including solvothermal and microemulsion synthesis.
In this context, hierarchical titanium dioxide nanofibers with distinctive microstructures exhibiting mixed rutile/anatase crystalline phase were successfully fabricated via microemulsion electrospinning approach by Zhang et al. [105]. More in detail, in order to make the microemulsion, paraffin was used as oil phase, while tetrabutyl titanate was dissolved in an alcoholic acid solution (containing a mixture of CTAB and PVP) as continuous phase.
The structure of porous nanofibers was regulated by changing the ratio of continuous phase and oil phase. Different morphologies such as multi-channel, hollow, irregular hollow TiO2 nanofibers were obtained by varying the ratio between the TiO2 precursor, tetrabutyl titanate (TBT), and paraffin oil [105]. In addition, the authors demonstrated that brittle nanostructure with abundant mesopores were formed when the ratio of TBT/paraffin oil decreased to 1. Compared with solid TiO2 nanofibers, all samples prepared by microemulsion electrospinning had improved photocatalytic performances towards the MB degradation UV irradiation. These results can be attributed to a combination of multiple structure dependent effects: first, porosity plays a vital role in absorbing photoelectron transition from rutile to anatase. In addition, photocatalyst with higher porosity can absorb more oxygen moieties on the surface which react with photoelectron, thus avoiding recombination and resulting in more electron and holes available for photocatalytic process (Figure 6). Finally, mixed anatase and rutile crystalline structure are expected to enhance charge separation ultimately improving photocatalytic performance
A soft template approach was exploited by Zhou et al. [106] to produce three-dimensional (3D) flower-like β-Bi2O3/Bi2O2CO3 heterojunction photocatalyst. A composite soft template composed of DL-aspartic acid (DLAA) and Pluronic F123 was employed to fabricate a 3D flower-like Bi containing micro/nanomaterial, in which DLAA acted as both coordination- and structure-directing agent to control the hierarchical structure, while F127 acted as a capping agent.
For the first time, the β-Bi2O3/Bi2O2CO3 p–n heterojunction photocatalyst was applied in the simulated-sunlight-driven photodegradation of antibiotic agent tetracycline (TC). The photocatalyst obtained after calcination at 290 °C showed relevant photocatalytic performance with 98.79% TC degradation being achieved within 60 min of irradiation. The obtained results confirm that the combination between the narrow band gap, heterojunction structure, and 3D hierarchical structure results in excellent photocatalytic performances. In particular, •OH and h+ proved to be the main active species in TC photodegradation process. Finally, the β-Bi2O3/Bi2O2CO3 heterojunction catalyst was not photo-corroded after six consecutive cycles, suggesting very high photostability (Figure 7).
Among bioavailable polymers, cellulose appears to be a promising green bio-template for porous hierarchical structure synthesis, because of its intrinsic hierarchical structure [107,108]. It was successfully exploited to produce H3PW12O40/TiO2 nanocomposite with remarkable activity towards organic pollutant removal under sunlight [107].
A non-solvent induced phase separation (NIPS) technique was explored by Sun et al. [108]. One-pot route towards active TiO2 doped hierarchically porous cellulose provided for highly efficient photocatalysts for methylene blue degradation [108]. Ethyl acetate was chosen as the non-solvent with the aim to induce the phase separation of cellulose into a cellulose/LiCl/N,N-dimethylacetamide solution containing TiO2 nanoparticles. The subsequent solvent exchange/freeze-drying treatment allowed to obtain cellulose/TiO2 composite monoliths featuring large surface area and hierarchically porous structures with two kinds of interconnected macropores. The cellulose/TiO2 monoliths showed high efficiency of photocatalytic activity in the decomposition of methylene blue dye up to 99% within 60 min under UV light. This behavior is probably due to high adsorption properties of the material itself because of the hierarchical porous structure which improves the contact frequency between photocatalysts and MB. In addition, after 10 cycles, the monoliths retained 90% of the photodegradation efficiency.
Table 2. Hierarchical catalysts produced by using synthesis templates.
Table 2. Hierarchical catalysts produced by using synthesis templates.
CompositionSynthesis StrategyMorphologySSA (m2/g)ApplicationActivityIrradiationTemplate(s)Template NatureReference
Anatase TiO2Sol-gelHierarchical ordered macro-meso 125.91photodegradation of aqueous Rhodamine B (RhB)90% 365 nmcopolymer P12/oyster shellSoft/hard[98]
Dimethylglyoxime (DMG)/TiO2/polyacrylonitrile (PAN) nanofiberElectrospinningNanofibers50–60Photocatalytic MB degradation efficiency97%Visible LightPAN/PVPSoft [103]
Bi2WO6/WO3/PANElectrospinning and solvothermal process Nanofibrous membrane-Degradation of cationic pollutants 85% of RhB,
87.8% for BQ and 95.7% for IPA, 96% of MB,
77.4% of chlortetracycline hydrochloride and 90.37% of tetracycline hydrochloride
UV-VisTriton-XSoft[104]
Anatase/rutile TiO2microemulsion electrospinning (ME-ES)/pyrolysismulti-channel irregular mesoporous TiO2 nanofibers (hundreds of nm in diameter)/Photodegradation of aqueous methylene Blue100%365 nmCTABsoft[105]
β-Bi2O3/Bi2O2CO3Sol-gel with refluxMesoporous Micrometric flower-like structures27.78Degradation of tetracycline98.79%SunlightDL-aspartic acid (DLAA)/Pluronic F123soft[106]
Table 3. Hierarchical catalysts produced by using bioavailable templates (biowaste).
Table 3. Hierarchical catalysts produced by using bioavailable templates (biowaste).
CompositionSynthesis StrategyMorphologySSA (m2/g)ApplicationActivityIrradiationTemplate(s)Template NatureReference
PC/POPhydrothermalSelf-supporting film-like structures122.5Cr(IV) photoreduction/methylene blue photodegradation100% Cr (IV) reduction/100% methylen blue photodegradationVisible lightPoplar catckinsHard[100]
α-Fe2O3Sol-gelhierarchical porous and fibrous structures (15 μm diameter, hundreds μm length)51.3Photodegradation of aqueous methylene Blue96.8%Visible lightCottonhard[102]
H3PW12O40/TiO2Sol-gel & calcinationnanotubes80.7Methylene blue photodegradation95% methylene blue degradationUV lightCellulosesoft[107]
Cellulose-TiO2non-solvent induced phase separation (NIPS) techniqueMacro-Mesoporous composite monoliths16.96photodegradation of aqueous methylene Blue99%UV lampCellulosesoft[108]
--

5. Hybrid Templates

At the boundary between organic and inorganic structure directing agents, metal-organic frameworks (MOF) are hugely porous crystalline hybrid structures made of metal clusters and organic binders. They are emerging as powerful and versatile hard templates to obtain hierarchical photocatalysts with tailored structure as well as porosity and function [109] (Table 4).
Indeed, upon decomposition of organic moieties, obtained materials inherit MOF large mesoporous structure allowing high active site exposure and enabling reactant and product exchange as well as fast charge migration to exposed reactive positions, with limited recombination phenomena [109,110]. As a result, MOF-templated photocatalysts are more active than their bulk counterparts and can be exploited in a huge range of energy and environmental applications, including H2 production [111,112], CO2 reduction [110,113,114], as well as organic pollutant degradation [115,116,117,118,119]. The main explored technological uses of MOF derived hierarchical photocatalysts are reported in Table 4.
MOF templated approach is extremely versatile and can be exploited to produce different semiconductor compositions. Porous metal oxides are easily obtained by MOF thermal decomposition [109]. Particularly, a Ti-MOF upon calcination leads to TiO2, composed by both anatase and rutile nanostructured phases, whose intimate interaction suppresses electron-hole recombination and promote photocatalytic activity, which results even better than P25 [120].
Following a similar strategy, Co MOF with 1,4-naphtalenedicarboxylic acid as ligand has been decomposed to obtain Co3O4 nanosheets with uniform size distribution and significant CO2 reduction conversion to CO, with high selectivity (77%) [110]. Moreover, metal oxides produced by MOF decomposition can be easily converted into metal sulfides by sulphuration process [121,122]. This strategy was exploited to obtain hierarchical CdS structures, with higher photocatalytic activity than both nanostructured and bulk counterparts, towards H2 production by water splitting [109]. Following a similar procedure, ZnCoS solid solution was fabricated, which exhibited high stability and activity because of a wide light absorption range, a great number of catalytic sites, and fast electron migration [122] (Table 4).
However, MOF templated method is time consuming since it is built on two steps, synthesis and calcination, which are difficult to be carried out on a large scale. Furthermore, the criteria to adequately choose MOF precursors must be further investigated to create the desired composition exhibiting hierarchical structure [109]. Indeed, selection of MOFs based on the same metallic cation, but with different structures, can deeply affect the morphology, structure, and overall performance of obtained catalysts. As a general rule, thermal stability is strictly demanded to avoid structure collapse during thermal treatment. Among available composition, MIL(53) meets this requirement, thus it holds huge promise to build controlled multilevel porous structures [109].
A promising strategy to obtain unique hierarchical structures relies on the use of defect-rich MOF [111]. This approach was effectively exploited to obtain hollow C doped CuO structures through pyrolytic degradation of Cu-MOF. In this study, not only did MOF act as a template to tune the structure of the newborn catalytic phase, but if pyrolysis occurred in reducing atmosphere, they could also supply for carbon atoms to modify the catalyst lattice. Both C-doped and C-containing hybrid composites can be produced following this approach and exhibiting relevant photocatalytic activity towards H2 production as well as pollutants degradation, due to peculiar porous and electronic structure [123,124] (Table 4).
Indeed, due to the controlled high surface area, MOF derived hierarchical structures are suitable platforms to be loaded with dopants that act as co-catalysts and can further improve photocatalytic properties. Apart from doping with C atoms, modification options include metal nanoparticles, encompassing noble metals (Pd, Pt) [114], and also cheaper transition metals (Ni, W, Cu) [125,126,127] as well as metal sulfides [128].
Indeed, doping can be carried out through “extra-situ” approach; in that case, MOF thermal decomposition can provide for either the catalyst lattice (Figure 8) or doping moieties for the preformed catalytic network. This can be further decorated with a proper dopant through hydrothermal treatment (Figure 8) [122,129,130].
Alternatively, in-situ methodology can be also explored, wherein the formation of catalyst backbone from MOF decomposition and dopants growth in the lattice concurrently occur during solvothermal process [126]. Because of MOF huge surface area and controlled porosity, high dopant dispersion and tunable morphology are usually afforded, producing a significant improvement of both catalytic performance and stability [14] thanks to the presence of multifunctional sites and the improved charge carrier separation [127]. Furthermore, concurrent modification with several co-catalysts can be easily carried out to achieve high photocatalytic performance. In this regard, TiO2 was combined with conductive bimetallic NiCoS-porous carbon shell, which afforded fast charge transport and prompt reaction for hydrogen production [129].
Moreover, MOF templates offer a simple, straightforward, and largely effective strategy in creating heterojunctions within hierarchical structures.
This approach is also successful for sulfide compounds, which usually require multi-step preparation (Figure 8). The large number of heterojunctions promote charge separation and migration and consequently much higher photocatalytic activity than conventional catalysts [131] for a huge range of applications spanning from organic pollutant removal to hydrogen production through water splitting [112,132]. MOF tunable composition and porosity offer the chance to prepare a plethora of hybrid catalysts combining hierarchical structures with p-n heterojunctions for photocatalytic processes (Table 4) [133]. Following this approach, creative and unique morphologies and structures can be achieved, including hierarchical hollow heterostructural cages combined with 2d nanosheets (Figure 9), hollow caps [134], boxes, polyhedra [135], and capsules which can be further modified through hydrothermal approach, reaching outstanding catalytic performance even under visible light.
To exploit advantages of combining different semiconductors, interface must be accurately designed to match band gaps and enable fast charge migrations as well as their easy transfer to reactants. To this purpose an appropriate distribution of reactive sites must be achieved. All these requirements can be easily met through MOF templated approach [136]. As a proof of concept for the efficacy of this strategy, hepitaxially grown MOF-on-MOF heterostructure was exploited as a precursor to produce N-doped C encapsulated pagoda-like CuO–In2O3 (In2O3/CuO@N-C) micro-rods. They exhibited improved light-absorption efficacy due to peculiar lamellar structure, enhanced separation efficacy of charge carriers thanks to CuO–In2O3 p–n heterojunction, and separated reduction and oxidation sites to promote charge carrier transfer to reactants. These features confer outstanding activity for cross-dehydrogenative coupling (CDC) reaction [136].
Due to high surface and reactivity, combination of MOF obtained photocatalysts into heterojunctions can be easily carried out through solvothermal synthesis [137,138]. This strategy was exploited to produce p-n heterostructure NiFe2O4/CuInSe2 catalysts with high activity towards bisphenol-A and resorcinol degradation [139].
Indeed, solvothermal approach appears as a straightforward production route, since moderate temperature during the process get organic ligands degraded, thus straightly leading to the final desired p-n semiconductor [137]. Following this method, even Z-scheme heterojunctions can be easily obtained [138], such as CoSx@CdS polyhedrons, provoking a marked improvement of charge transfer and enabling a high sensitivity toward Hg2+ ions detection (Figure 10) [138]. As an alternative, p-n heterojunction can be obtained by thermal treatment of bi-metallic MOF [112].
Moreover, one of the highest strengths of MOF templated approach lies in the opportunity to design complex hierarchical architectures even with multi-level hierarchy. To this regard, MOF can be effectively used to obtain 2D hierarchical nanoarrays, by assembling 2D structures and conductive sheets, which exhibit outstanding catalytic activity because of high catalytic site exposure and fast charge transport [113,117]. As a proof of concept, in-situ MOF derived approach was successfully employed to produce hierarchical Co-Co layered double hydroxide/Ti3C2TX nanosheets (Co-Co LDH/TNS) nanoarray through solvothermal process and evidencing relevant activity towards CO2 photoreduction under visible light (Figure 11) [113].
Highly effective heterojunctions can be also obtained by combining MOF and C3N4 sheets via thermal treatment. Following this route, N doped carbon C3N4 composite was produced exhibiting relevant activity towards bisphenol A Degradation (Table 4) [117]. Finally, MOF derived multilevel composites exhibit multiple heterojunctions and consequently enhanced catalytic performance even in selective oxidation [116]. Similarly, one-pot calcination process of Zn-Fe mixed MOF produced ZnO/ZnFe2O4 hierarchical heterostructures with high photocatalytic activity towards dye degradation and which can be easily recovered from solution because of its ferromagnetic features [119].
Indeed, due to the high versatility of MOF templates, they provide the opportunity to engineer the photocatalyst structure, by combining heterostructure formation with co-catalyst surface modification [121,140]. This occurred in TiO2 based photocatalyst, which was modified by Co3O4 and Ni to promote oxidation and reduction, respectively. Obtained ternary TiO2/Co3O4/Ni photocatalyst disclosed an impressive hydrogen production 8.7 times higher than that of neat TiO2 under UV-visible irradiation [140]. Because of the high versatility of MOF templated approach, tuning of electronic properties, through heterojunction formation and co-catalyst introduction, is accompanied by fine morphology control. In this regard, hollow CdS/TiO2 nanohybrids modified with NiS cocatalyst exhibited very good photocatalytic performance towards H2 production under visible light conditions (Figure 12) [121].
Table 4. Hierarchical catalysts produced by using hybrid templates.
Table 4. Hierarchical catalysts produced by using hybrid templates.
CompositionSynthesis StrategyMorphologySSA (m2/g)ApplicationActivityIrradiationTemplate(s)Reference
Rutile/anatase TiO2Sol-gelSubmicrometric parallelepiped with rounded corners19Photocatalytic reduction of water1394 μmol
h−1 g−1 H2 production rate
UV-vis MIL-125[120]
Co3O4Oil bath method2D hierarchical nanosheets24.96Efficient photocatalytic conversion of CO239.70 μmol h−1 CO generationVisible lightCo/1,4-H2NDC[110]
NiS/CdS/TiO2Hydrolysis + sulfidationPorous/hollow structure Photocatalytic reduction of waterH2 production rate of 2149.15 µmol g−1 h−1Visible lightNH2-MIL-125[121]
CdSImpregnation + pyrolysisMicroporous nanoparticles 119Photocatalytic reduction of water634.0 µmol g−1 h−1Visible lightMIL-53 (Al)[109]
C-CuOpyrolysisHollow spheres72.8Photocatalytic reduction of water67.3 mmol/g/h H2 productionVisible lightCu/benzoic acid/1,4
dicarboxybenzene
[111]
C-ZnOPyrolysis under N2 atmosphereZnO crystals embedded within a porous carbonaceous matrix500Adsorption and photodegradation of RhB100% adsorption efficiency//MOF-5[123]
Pd-Anatase TiO2pyrolysisSubmicron TiO2 tablets Photocatalytic reduction of water979.7/112.7 mol h−1 H2 productionUV/solar lightNH2-MIL-125[114]
Ni/g-C3N4Pyrolysis under Ar atmospherelayered
platelets with curled edges
64.9Photocatalytic reduction of waterH2 production rate of 2989.5 μmol g−1 h−1Visible light2D Co-MOF[125]
W/Co3S4Hydrothermal treatment10 mm square sheet arrangement10.93Photocatalytic reduction of water85.7 μmol/h H2 production rateVisible lightCo-ZIF-9[126]
Cu/TiO2Photolysis under N2 atmosphereMesoporous core-shell Cu/TiO2 hybrid nanoplatforms-Photocatalytic reduction of water334 μmol g−1 h−1 H2 productionSimulated lightCu-MOF[127]
TiO2–NiCoS-PCAnnealing under Ar atmpsphereSpherical porous carbon shell Embedding TiO293.5Photocatalytic reduction of water1.29 mmol h−1 g−1Visible lightNiCo-ZIF[129]
MoS2−Zn0.5Co0.5SHydrothermalHollow rhombic dodecahedra57Photocatalytic reduction of water15.47 mmol h−1 g−1UV-VisZnxCo1−x-ZIF[122]
MnS/In2S3SolvothermalSub-micro rods /photocatalytic CO2 reduction58 μmol g−1 h−1 CO production rateXenon lamp 300 W (credo UV-vis)MIL-68 (In)[131]
NiO/CeO2Hydrothermal and calcinationPorous microsphere31.1Dye photodegradation/water splitting100% MO and MB degradation/71.5 μmol g−1 H2 production rateUVNi/Ce mixed-metal
MOF
[112]
Cu/CuO-TiO2Sol-gel + calcinationnanoparticles45.3Photocatalytic reduction of water286 mmol g−1 h−1 H2 production rateSolar (Sun)Cu-MOF[132]
CdS/
ZnxCo3-xO4
Hydrothermal methodHollow polyhedra29.6Photocatalytic reduction of water3978.6 μmol g−1 h−1 H2 production rateVisibleZnCo-ZIF[135]
In2O3/CuO@N-CcalcinationPagoda-like heteroepitaxial micro-rods147.1cross-dehydrogenative coupling (CDC)88–99% yield of reactionBlue lightMIL-68-In[136]
CuInSe2/NiFe2O4hydrothermalIrregular-shaped nanoparticles27.54Bisphenol-A and resorcinol degradation95–95% removal rateVisible lightNi–Fe MOF[137]
CoSx/CdSsolvothermalCdS NPs on the surface of CoSx polyhedrons/Hg2+ detection0.010 to 1000 nM Hg2+ detectionVisible lightZIF-67[138]
Co-Co LDH/Ti3C2TXsolvothermal3D nanosheets nanoarray/CO2 photoreduction1.25 × 104 µmol h−1 g−1 CO production rateVisible lightZIF-67[113]
ZIF-NC/g-C3N4 Thermal treatmentZIF-NC dodecahedra deposited onto g-C3N4 layers/bisphenol A degradation95% removalVisible lightZIF-8[117]
MoS2@Cu/Cu2O@CPyrolysisHierarchical rough polyhedra94Selective oxidation of cyclohexane to KA oil1.31% conversion rate; 98% selectivityVisible lightCu-MOF[116]
ZnO/ZnFe2O4calcinationSpherical assemble of flake-like nanosheets87.74RhB and MB degradation100% degradationUV-visZn(Fe)-MOF[119]

6. Template Free Approach

Using a templating agent proved to be effective in building up hierarchical structures, yet it is a labor consuming approach, because of the operations required to remove the backbone.
On the other hand, self-assembly of primary particles into larger structures by oriented aggregation occurring during solvothermal processes can be a more straightforward and greener strategy, since it can potentially avoid toxic chemicals [141]. In these processes, it appears that solvents play a key role in hierarchical morphology formation [142]; indeed, their molecules or, in the case of alcohols, polymers obtained by condensation reactions can adsorb on the particle surface and tune their growth [143]. Solvothermal route has been successfully exploited to obtain hierarchically structured conventional semiconductor oxides including ZnO [60,142] TiO2 [144,145,146], CeO2 [147], Cu2O [148], and more unusual compositions, such as WO3 [149] and mesoporous Bi2WO6 [150]. Furthermore, doping of the main lattice with various ions can be carried out concurrently with its formation [151]. Moreover, a relevant number of heterostructured composites has been easily produced so far through one pot solvothermal treatment; these include combinations of semiconductor oxides (TiO2, ZnO, CeO2) with other semiconductors [62] including metal sulfides (SnS2) [152], carbon nitride [133], or with carbonaceous species (CS). This class encompasses ZnO-graphene [141,153] and TIO2-CS particles [154] as well as TiO2 nanotubes wrapped by carbon coatings [155], among others. Furthermore, CoS/CdS heterojunctions [156] and even less conventional compositions, including BiVO4/Bi2WO6 [157] systems, can be easily fabricated. Indeed, the presence of different morphology nanostructures, such as g-C3N4 nanosheets and ZnO nanoparticles, in the same batch during the solvothermal treatment appears to be a good strategy promoting self-assembly into hierarchical structures to decrease surface energy [133]. Moreover, synthesis parameters and particularly reactant molar ratio, solvent and precursor nature, as well as temperature and treatment time [145] can tune morphological and microstructural features. Accordingly, a significant range of morphologies can be easily fabricated, including hollow microspheres [141,142,149,155,158,159], nanoplates [156], as well as more exotic arrangements, such as broom-like [151], urchin-like [160], and nest-like [149] structures (Figure 13).
As a matter of fact, photocatalytic activity of hierarchical ZnO obtained by solvothermal process decrease according to the following order of employed precursor: zinc sulfate > zinc nitrate > zinc acetate > zinc chloride [60]. Furthermore, the use of pure ethanol as solvent during solvothermal treatment of TiOSO4 precursor produces nanoparticle morphology, whereas the addition of ethyl ether and ethanol mixture has morphology changed to microrods with huge surface area (244.4 m2 g−1), promoting light scattering and reactivity [155].
Solvothermal hierarchical structures usually feature an interconnected porosity with proper pore size to enable mass transfer as well as improved light harvesting and even fast charge carrier mobility; these properties concur to achieve higher photocatalytic performance than conventional systems [150]. Obtained systems exhibit relevant activity as well as photoelectrochemical performances under UV [141,146] and even visible [145,148,156,157] and full solar irradiation [155] for energy and environmental photocatalysis. Applications range from environmental to energetic field, spanning from pollutant degradation in water and gas phase [144], including dye [62,133,145,146,157] and nitric oxide removal (Table 5) [150,161] to H2 production [156].
The use of small organic molecules as structure directing agents can aid self-assembly and significant enhance photocatalytic performance, due to the formation of peculiar morphologies, with high degree of crystallinity, high light-harvesting ability, as well as efficient separation of photogenerated carriers [159,161,162]. At the same time, their decomposition can provide a source of doping species for the catalyst including carbon atoms which can also give rise to organized porous frameworks hosting the photoactive phase and contributing to a significant improvement of its performance under UV-irradiation [163]. To this regard, citric acid has been exploited to synthetize urchin-like LaPO4 hollow spheres, exhibiting marked activity towards CO2 reduction under near UV [159].
Similarly, the use of N-Acetyl-d-Proline amino acid during hydrothermal treatment produced multi-shelled ZnO microspheres more active for NO oxidation under UV irradiation than other hierarchical ZnO samples [161]. Indeed, low molecular weight compounds can play multiple roles during the hydrothermal process, behaving not only as morphological agents but even as reducing compounds [164,165]. As a matter of fact, L-cysteine and glucose were employed during the hydrothermal treatment to obtain a TiO2-graphene-MoS2 composite as a high performance and stable catalyst for CO2 reduction (Figure 14). Notably, L-cysteine acted as a reducing agent for graphene oxide, whereas glucose played a key role to control morphology, since it formed amorphous carbon which inhibited the TiO2 and MoS2 growth (Table 5 and Figure 14) [165].
Template free solvothermal processes can be carried out sequentially [162] or straightforwardly combined with wet impregnation or deposition procedures to have surface of produced catalyst decorated with other active moieties [161,163]. These often include plasmonic metal nanoparticles [163,166] as well as other semiconductors including both inorganic and organic systems [162,167]. The former improve catalytic activity through surface plasmon resonance [163]. The latter enable formation of p-n [161] as well as Z scheme heterostructures (Figure 15) [156,162,167,168], which exhibit superior visible light photocatalytic activity coupled with relevant stability and reusability [168,169].
Among wet chemistry routes, precipitation represents a facile low temperature approach to make hierarchical structures [159,170]. This method was effectively employed to produce CdMoO4 hollow microspheres which exhibited higher photocatalytic activity towards Rhodamine-B aqueous solution under the UV irradiation than commercial TiO2 [170].
In some cases, self-assembly into a hierarchical architecture can be promoted by electrostatic aggregation [171]. To this regard, g-C3N4/ZnO microspheres were easily produced starting from a physical dispersion of the two components and exploiting their opposite surface charge. Z-scheme obtained by band matching between the two semiconductors as well as enhanced light absorption because of multiple scattering were responsible for a marked improvement of photocatalytic performance towards CO2 reduction under visible light (Figure 16) [171].
Other template free chemistry routes than solvothermal synthesis can be exploited to obtain hierarchical architectures. These include solid phase processes based on either solid or solid-gas reactions. Notably, solid-state reactions followed by hydrolysis at room temperature proved to be a simple route to make bismuth oxoiodide (BiO) with surface heterojunction between (0 0 1) and (1 1 0) surfaces promoting photocatalytic performance under visible light [172]. Solid-gas phase reactions usually involve pyrolysis [173], decomposition [174], or oxidation [175] and occur at moderate temperatures (higher than 150 °C). In these processes, temperature plays a key role in defining the final structure and overall photocatalytic performance, accordingly [174]. Moreover, either metal or non-metal heteroatom doping [175,176] as well as semiconductor heterostructures [171] can be easily achieved during the thermal process leading to superior photocatalytic performances [175,176]. Following this approach, nanosheet flower-like SnS2 nanostructures have been obtained and used for Cr(IV) photocatalytic reduction under visible light with high performance as stability [173], even though these methods have been more extensively applied to make C3N4 catalysts. As a matter of fact, a facile pyrolysis process with urea was exploited to produce Yttrium-doped graphitic carbon nitride (Y/g-C3N4) catalysts, which feature high surface area, poor recombination phenomena, as well as wide spectral absorption accounting for enhanced photocatalytic activity towards rhodamine B degradation under visible light [176]. Alternatively, thermal oxidation and exfoliation, followed by curling condensation were carried out on bulk C3N4. Obtained O-doped carbon nitride (g-C3N4) nanotubes exhibit outstanding photocatalytic activity towards CO2 reduction to synthetic fuels, sensitively higher than massive g-C3N4 [175].
Indeed, obtaining porous g-C3N4 is a challenging task, most of available processes to fabricate being time-consuming and energy intensive. A green and facile alternative consists in a fast quenching from high temperature, which produces an 3D interconnected porous assembly of g-C3N4 nano-scrolls, featuring high light absorption, poor recombination phenomena, and enhanced carrier mobility, accounting for improved catalytic performance toward CO2 reduction [177].
Hierarchical heterostructures between semiconductors can be straightforwardly obtained by simultaneous decomposition of their precursors [171]. Following this route, calcination of tetrabutyl titanate and melamine precursors produced Z-scheme TiO2/g-C3N4 heterojunction with unique features even due to concurrent C and N doping [171].
Template free processes also include combustion synthesis, a simple one-pot method exploiting fuel-oxidizer solutions, such as those containing sorbitol-nitrate mixtures, to fabricate mixed phase hierarchical structures, including bismuth rich-bismuth oxychlorides nanocatalysts, which exhibited high photocatalytic activity towards antibiotic ofloxacin degradation under solar light irradiation [178].
Moreover, non-conventional energy sources can be exploited to build self-templated hierarchical structures. To this regard, pulsed laser deposition (PLD) followed by thermal treatment had hierarchical nanostructured hematite coated onto glass surface and exhibiting enhanced and stable catalytic performance because of high availability of active sites, improved interaction with reactants as well as limited recombination phenomena [179].
Alternatively, ultrasound microwave and ultrasound irradiation can assist wet chemistry processes to promote hierarchical structure formation. Notably, microwave irradiation of copper acetate and thiourea aqueous solutions led to hierarchical CuS structures with tunable porosity and morphology, which evidenced higher activity than commercial CuS catalyst in dye degradation processes under visible light [180,181]. As an alternative, ultrasound irradiation is understood to be a facile and fast yet extremely efficient method to assist particles self-assembly processes into hierarchical architectures [160]. It proved effective in producing Urchin-shaped iron oxide nanoparticles, with huge surface area, up to 282.7 m2/g and which evidenced 25 times faster removal of As(V) and Cr(VI) than commercial Fe2O3 (Figure 17) [160].
Similarly, sonochemical treatment of Melamine in water and its further calcination produced hierarchical graphitic-C3N4, which was tested for rhodamine B degradation and H2 production and exhibited 16.7 and 8.7 higher visible light photo-catalytic activity than bulk homologous composition [182].
On the other hand, reactive magnetron sputtering deposition appears to be a facile methodology for large scale fabrication of hierarchical films [183]. This approach was exploited to fabricate two level porosity AgO nanorod arrays, which exhibited excellent photocatalytic properties under UV and visible-light conditions [183].
Table 5. Hierarchical catalysts produced without templates.
Table 5. Hierarchical catalysts produced without templates.
CompositionSynthesis StrategyMorphologySSA
(m2/g)
ApplicationActivityIrradiationRef
ZnCuCo layered double hydroxideHydrothermal3D flower-like hierarchical morphologies with in-
terlaced petal-like nanosheet
86–179photocatalytic H2 production and degradation of
SMZ
H2 production rate (3700 μmol g−1 h−1,
95% sulfamethazine (SMZ) degradation
Visible light[43]
Ti3C2/Bi2WO6hydrothermal2D/2D hetero junction i33.5–58.3Photocatalysis degradation of tetracycline hydrochloride (TC-HCl)0.430 min−1Visible light[44]
ZnFe2O4 modified Cu2Sin-situ self-assembly methodDendritic fractal structure similar to a snowflake-photocatalytic reduction of nitrobenzene with and degradation of methyl orange and methylene blue dyes proficiently 98% yield of degradation of nitrobenzene,
94.3 and 86% fordegradation of methyl orange and methylene blue, respectively
Visible light [45]
ZnO-graphene nanocompositeSolution route hollow microspheres 29.7–37.6Adsorption/photocatalytic activity towards degradation of water-soluble organic pollutants (such as Rhodamine B, methyl orange, phenol)90% adsorption capacityUV (254 nm)[141]
ZnOSolvothermal route Monodisperse microspheres 18Photocatalytic activity for degradation of methylene blue 100%UV-Vis [142]
ZnOHydrothermal and calcination Hierarchically porous microspheres composed of nanosheets 46–91Photocatalytic RhB degradation activityUp to 100%solar[60]
TiO2Hydrothermal route(i) long and well oriented macrochannels
(ii) surface macropores of 0.8–9.3 μm in size, and (iii) porous walls with pores mostly smaller than 0.5 μm.
82.9–216.1 Decomposition of methyl ethyl ketone (MEK) in a continuous flow photoreactorUp to 37%UV (254 nm)[144]
TiO2Hydrothermal route Partial spherical like structure30–43.7RhB photodegradation Up to 98%Visible light[145]
TiO2Hydrothermal routeNanoflowers with a spherical hierarchical structure20–80Aqueous methylene blue photo-oxidationUp to 50%solar[146]
CeO2Hydrothermal routeMesoporous nanosphere42.1–68.2photocatalytic activity of rhodamine B (RhB) dye degradation 97.8–92% RhB dye degradationUV-Vis and acidic condition[147]
Cu2OLow temperature routeSpherical, cuboctahedral or cubic nanoparticles7–13Photocatalytic degradation of the antibiotic trimethoprimUp to 48%Visible light[148]
WO3Hydrothermal routeRegular-shaped nanosheets with an average thickness of approximately 30–40 nm2–18Photocatalytic activity towards an aqueous solution of tetracycline (TC) and possess good stability and reusabilityUp to 94%Visible light[149]
Bi2WO6Hydrothermal method and calcining processMesoporous nanoplate multi-directional53.5Photocatalytic oxidation of NOUp to 90%Solar[150]
Sm, Y, La and Nd-doped CeO2hydrothermalbroom-like hierarchical structure-BPA degradation and on CO2 evolution from CH3CHO decompositionUp to 99%UV[151]
Yttria (Y2O3) nanosphere decorated ceria (CeO2)Hydrothermal routeYttrium nanoparticles are anchored on the surface of CeO2 nanorod with a particle size of 10 nm-Photocatalytic decomposition of aqueous Rhodamine BUp to 96%Solar light[62]
SnS2/TiO2Ultrasonic treatmentOrdered channels with a size about 1–3 μm were formed in the particles, and lots of holes appeared on the wall of the channels.28Photocatalytic degradation of Methyl Orange (MO)Up to 90.9%Solar light[152]
ZnO/g-C3N4SolvothermalPorous microsphere with a size of about 700nm. 9.9–32-7Photocatalytic degradation on rhodamine B and phenolUp to 100%Solar[133]
ZnO/grapheneSolvothermalCore-shell structure65–201Photocatalytic degradation of rhodamine BUp to 98.5%Visible light[153]
TiO2 Microspheres with Carbonaceous SpeciesSolvothermal Porous structure337Photodegradation of rhodamine BUp to 100%Visible light [154]
Carbon-coated TiO2solvothermalHierarchical nanotubesUp to 244.4Photocatalysts activity for water oxidationUp to 705 µmol h−1 g−1 Solar light[155]
S-deficient CoS/CdS SolvothermalHexagonal nanoplatesUp to 84.32Photocatalytic water-splittingUp to 14.5%Visible light[156]
BiVO4/Bi2WO6SolvothermalSelf-assembled hierarchical BiVO4/Bi2WO6 heterostructured composites5.16Photocatalytic activities for methylene blue (MB) degradation and photoelectrochemical performanceUp to 50%Visible light[157]
3D LaPO4Solution route using citric acid (CA)Urchin-like hollow sphere51–124Photocatalytic CO2-reduction performance6.8-fold enhancement of the AQYUV[159]
ZnOHydrothermal-calcinatio
n
Mesoporous multi-shelled ZnO
microspheres
3–20Photocatalysis for NO oxidation Up to 77%UV[161]
ZnO/CeO2Hydrothermal/calcinationSpherically hierarchical structure 41–56photocatalytic rhodamine B (RhB)Up to 96%Visible light [162]
TiO2/graphene/MoS2Hydrothermal2D rGO sheets assembled into macroporous 3D structures 124CO2 Reduction PhotocatalystUp to 97%UV-Vis[165]
Ag,Au/TiO2SolvothermalMesoporous spherical shape153–173Degradation of Textile DyesUp to 85%Solar and visible light[163]
Au-H-ZnOLow temperature aqueous reaction and heat treatmentNanosheets-Gas sensing and photocatalytic propertiesUp to 94.8%Visible light[166]
PDA-modified ZnSelf-assemblyRough microstructures32–38Photocatalytic CO2 reduction Up to 0.95 µmol h−1 g−1 of CH3OHUV[167]
ZnIn2S4 marigold flower/Bi2WO6 (ZIS/BW)Hydrothermal method followed by wet-impregnationHierarchical marigold flower and flower-like morphologies14–73Decomposition of metronidazole Up to 56%Visible light [168]
BiOCl/BiVO4Coprecipitation-hydrothermal methodMicro-nanosheet1.53–2.83Photo-
degradation rate of rhodamine B (RhB)
Up to 96%Visible light[169]
CuSMicrowave-assisted wet chemical processSpherical monodispersed submicron particles17.8–26.4 Decomposition of organic dyes including rhodamine B, methylene blue and malachite green Up to 100%Visible light[180]
CuSMicrowave irradiationAggregates of roughly spherical nanoparticles14.74Degradation of methylene blue, methyl orange, and 4- chlorophenolUp to 100%Solar light[181]
3D- Fe2O3Ultrasound irradiation methodLarge quantity of 3D sea urchin-like structures combining a 1D rod-like structure on spherical support129.4–282.7Adsorption for heavy metals and photocatalytic activities toward the dyes (methylene blue and phenol)Up to 100% Solar light[160]
g-C3N4Ultrasound-assisted molecular rearrangement strategyHierarchical Rodlike88.6RhB degradation and H2 evolutionUp to 23%Visible light[182]
CdMoO4Low temperature oil bath methodUniform and porous spheres-Photocatalytic removal of mixed dye aqueous solutionsUp to 100%UV-Vis[170]
g-C3N4/ZnOElectrostatic self-assembly methodnanosheets156Photocatalytic CO2 reduction activity0.64 µmol h−1 g−1Solar light[171]
BiOISolid-state reaction with
subsequent hydrolysis at room temperature.
Hierarchical microspheres assembled by nanoplates3.1–13.8Photocatalytic activity for phenol degradation.Up to 100%Visible light[172]
SnS2Heating the mixture of SnCl2·2H2O and thiourea in air at 170 °C for 2Porous flower-like hierarchical nanostructure36.15–82.4Adsorption and photocatalytic reduction of aqueous Cr(VI)Up to 79.4%Visible light [173]
ZnOAnnealing of zinc oxalate.Mesoporous nanostructured microlumps1.7–29−9Discoloration of the Methyl violet 2BUp to 100%UV[174]
O-Doped g-C3N4Successive thermal oxidation exfoliation and curling-condensation of bulk g-C3N4Uniform porous network36Photocatalytic CO2 Reduction ActivityUp to 0.88 μmol g−1 h−1 Visible light [175]
Yttrium-doped g-C3N4Pyrolysis methodSheet-like morphology with worm-like pores39–106Photocatalytic performance in rhodamine B degradationUp to 100%Visible light[176]
3D g-C3N4Cold quenchingSelf-assembled nanoscrolls76Photocatalytic CO2 reductionUp to 11.2 μmol g−1 h−1UV-Vis[177]
Bismuth oxychlorides
(BOC)
One-pot sorbitol-nitrate solution auto-combustion methodMesoporous-mixed-phase of grain-like93Photocatalytic application in treatment of antibiotic effluentsUp to 80%Visible light[178]
Fe2O3Pulsed Laser Deposition and thermal oxidationSpherical particulates to an urchin-like struc-
ture with evolution of nanowires
-Photocatalytic water purificationUp to 100%H2O2 and visible light [179]
AgOOxidizing solid Ag films in an environment of reactive magnetron sputtering deposition of NiOPorous AgO nanorod films.-Photocatalyst and all solid-state thin film batteryUp to 100%UV-Vis[183]

7. Conclusions

Hierarchical structures have raised great interest for different photocatalytic applications, because of improved light harvesting, charge separation, as well as mass transport and adsorption efficacy, resulting in superior catalytic properties. Even though different fabrication strategies have been explored, as discussed in this review, some major challenges must be still addressed to allow easy synthesis of hierarchical photocatalysts and help to scale up process from laboratory to industrial production.
Among available methods, templated free wet synthesis approaches based on self- aggregation processes offer a facile route to obtain a wide choice of compositions and structures. Although most of them feature lower average surface area than systems obtained using inorganic templates, they usually exhibit comparable photocatalytic performance especially towards pollutant degradation, which is complete under UV irradiation and, for some composition reaches even 95% under visible light. Yet, some issues must be addressed mainly related to the difficulty to carry out large scale and high yield synthesis. Furthermore, post-process modification is usually required to further improve photocatalytic properties. Among templating approaches, inorganic backbones lead to the largest surface area, as high as 400 m2/g, even though this feature does not necessarily result in the best photocatalytic performance, while organic templated systems exhibit comparable activity, allowing for organic pollutant complete degradation even under visible light. On the other hand, hybrid templating strategies, based on MOFs, are extremely versatile since they provide for a huge number of semiconductor compositions, with tunable morphology and structure. In addition, post-modification is usually not required, since catalyst functional modification including doping, hetero-junction formation, and loading with co-catalyst moieties can be easily carried out in-situ upon thermal treatments in dry or solvothermal conditions, concurrently with template decomposition. Carbonaceous residues resulting from this process can act themselves as doping species and sensitively improve photocatalytic properties. However, MOF templated method is expensive, time consuming, and difficult to be carried out on a large scale. Moreover, hierarchical structures obtained through this approach often show comparable performance to those obtained through self-templated methods, particularly for decontamination processes, thus the latter are usually preferred for these applications. On the other hand, more advanced technological routes involving hydrogen production are worth the effort of using MOF templated structures, which outperform with respect to conventional templates because of their extremely versatile features.
Nevertheless, the criteria to adequately choose MOF precursors must be further investigated to create the desired composition exhibiting hierarchical structure.
Bio-templates offer a sustainable and simple opportunity to design hierarchical photocatalytic nanomaterials. Among bioavailable compounds, bio-wastes are attracting great interest as abundant and cheap templates. However, their heterogeneity and poor knowledge of growth mechanisms make the process difficult and poor reproducible. Further investigation on building blocks growth and assembly is highly demanded to overcome these limitations and achieve rational control of the final hierarchical structure. More efforts should be taken towards the investigation of structure/function relationships and the mechanism underlying photocatalytic behavior.
Finally, a combination of the described strategies appears to be a promising way to combine benefits of each single approach. In this view, very recent studies evidence that integrating bio-templated design with MOF derived nanoparticles provide for a unique hierarchical structure, with efficient light absorption, superior charge separation efficacy, accounting for potent photocatalytic activity [184]. To this aim, this study could provide useful information to move towards more performant hierarchical nanostructures for advanced environmental and energy applications.

Author Contributions

G.L. drafted this review; V.V. and G.P. contributed equally to this work; B.S., A.C. and G.V. commented on the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mukherjee, A. Biomimetics Learning from Nature; IntechOpen: London, UK, 2012; ISBN 9789533070254. [Google Scholar]
  2. Bhushan, B. Biomimetics: Lessons from Nature—An overview. Philos. Trans. R. Soc. A 2009, 367, 1445–1486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wegst, U.G.K.; Bai, H.; Saiz, E.; Tomsia, A.P.; Ritchie, R.O. Bioinspired structural materials. Nat. Mater. 2015, 14, 23–36. [Google Scholar] [CrossRef] [PubMed]
  4. Guan, Q.F.; Yang, H.B.; Han, Z.M.; Ling, Z.C.; Yu, S.H. An all-natural bioinspired structural material for plastic replacement. Nat. Commun. 2020, 11, 5401. [Google Scholar] [CrossRef] [PubMed]
  5. Martin-Martinez, F.J.; Jin, K.; López Barreiro, D.; Buehler, M.J. The rise of hierarchical nanostructured materials from renewable sources: Learning from nature. ACS Nano 2018, 12, 7425–7433. [Google Scholar] [CrossRef]
  6. Fratzl, P.; Weinkamer, R. Nature’s hierarchical materials. Prog. Mater. Sci. 2007, 52, 1263–1334. [Google Scholar] [CrossRef] [Green Version]
  7. Dan, N. Synthesis of hierarchical materials. Trends Biotechnol. 2000, 18, 370–374. [Google Scholar] [CrossRef]
  8. Koch, K.; Bhushan, B.; Jung, Y.C.; Barthlott, W. Fabrication of artificial Lotus leaves and significance of hierarchical structure for superhydrophobicity and low adhesion. Soft Matter 2009, 5, 1386–1393. [Google Scholar] [CrossRef]
  9. Latthe, S.S.; Terashima, C.; Nakata, K.; Fujishima, A. Superhydrophobic surfaces developed by mimicking hierarchical surface morphology of lotus leaf. Molecules 2014, 19, 4256–4283. [Google Scholar] [CrossRef] [Green Version]
  10. Lingham-Soliar, T. Microstructural tissue-engineering in the rachis and barbs of bird feathers. Sci. Rep. 2017, 7, 45162. [Google Scholar] [CrossRef] [Green Version]
  11. Sullivan, T.N.; Wang, B.; Espinosa, H.D.; Meyers, M.A. Extreme lightweight structures: Avian feathers and bones. Mater. Today 2017, 20, 377–391. [Google Scholar] [CrossRef]
  12. Zhu, Y.; Zhang, W.; Zhang, D. Fabrication of Sensor Materials Inspired by Butterfly Wings. Adv. Mater. Technol. 2017, 2, 1600209. [Google Scholar] [CrossRef]
  13. Niu, S.; Li, B.; Mu, Z.; Yang, M.; Zhang, J.; Han, Z.; Ren, L. Excellent structure-based multifunction of morpho butterfly wings: A review. J. Bionic Eng. 2015, 12, 170–189. [Google Scholar] [CrossRef]
  14. Su, B.-L.; Sanchez, C.; Yang, X.-Y. Hierarchically Structured Porous Materials: From Nanoscience to Catalysis, Separation, Optics, Energy, and Life Science; John Wiley & Sons: Hoboken, NJ, USA, 2012; ISBN 3527639594. [Google Scholar]
  15. Li, X.; Yu, J.; Jaroniec, M. Hierarchical photocatalysts. Chem. Soc. Rev. 2016, 45, 2603–2636. [Google Scholar] [CrossRef] [PubMed]
  16. Trogadas, P.; Ramani, V.; Strasser, P.; Fuller, T.F.; Coppens, M. Hierarchically structured nanomaterials for electrochemical energy conversion. Angew. Chem. Int. Ed. 2016, 55, 122–148. [Google Scholar] [CrossRef] [Green Version]
  17. Shi, Y.; Ye, G.; Yang, C.; Tang, Y.; Peng, C.; Qian, G.; Yuan, W.; Duan, X.; Zhou, X. Pore engineering of hierarchically structured hydrodemetallization catalyst pellets in a fixed bed reactor. Chem. Eng. Sci. 2019, 202, 336–346. [Google Scholar] [CrossRef]
  18. Hsu, M.-H.; Chang, C.-J. S-doped ZnO nanorods on stainless-steel wire mesh as immobilized hierarchical photocatalysts for photocatalytic H2 production. Int. J. Hydrog. Energy 2014, 39, 16524–16533. [Google Scholar] [CrossRef]
  19. Miao, Y.-E.; Wang, R.; Chen, D.; Liu, Z.; Liu, T. Electrospun self-standing membrane of hierarchical SiO2@ γ-AlOOH (Boehmite) core/sheath fibers for water remediation. ACS Appl. Mater. Interfaces 2012, 4, 5353–5359. [Google Scholar] [CrossRef]
  20. Brady, R.; Woonton, B.; Gee, M.L.; O’Connor, A.J. Hierarchical mesoporous silica materials for separation of functional food ingredients—A review. Innov. Food Sci. Emerg. Technol. 2008, 9, 243–248. [Google Scholar] [CrossRef]
  21. Nayak, A.K.; Ghosh, R.; Santra, S.; Guha, P.K.; Pradhan, D. Hierarchical nanostructured WO3-SnO2 for selective sensing of volatile organic compounds. Nanoscale 2015, 7, 12460–12473. [Google Scholar] [CrossRef]
  22. Gao, X.; Wang, Y.; Ji, G.; Cui, R.; Liu, Z. One-pot synthesis of hierarchical-pore metal–organic frameworks for drug delivery and fluorescent imaging. CrystEngComm 2018, 20, 1087–1093. [Google Scholar] [CrossRef]
  23. Hartmann, M. Hierarchical zeolites: A proven strategy to combine shape selectivity with efficient mass transport. Angew. Chemie Int. Ed. 2004, 43, 5880–5882. [Google Scholar] [CrossRef] [PubMed]
  24. Su, B.-L.; Sanchez, C.; Yang, X.-Y. Zeolite Molecular Sieves Chemistry of Zeolites and Related Porous Materials Ordered Mesoporous Materials; John Wiley & Sons: Chichester, UK, 2012; ISBN 9780470577578. [Google Scholar]
  25. Holm, M.S.; Taarning, E.; Egeblad, K.; Christensen, C.H. Catalysis with hierarchical zeolites. Catal. Today 2011, 168, 3–16. [Google Scholar] [CrossRef]
  26. Parlett, C.M.A.; Wilson, K.; Lee, A.F. Hierarchical porous materials: Catalytic applications. Chem. Soc. Rev. 2013, 42, 3876–3893. [Google Scholar] [CrossRef] [PubMed]
  27. Costantini, A.; Venezia, V.; Pota, G.; Bifulco, A.; Califano, V.; Sannino, F. Adsorption of cellulase on wrinkled silica nanoparticles with enhanced inter-wrinkle distance. Nanomaterials 2020, 10, 1799. [Google Scholar] [CrossRef] [PubMed]
  28. Tang, Y.; Wee, P.; Lai, Y.; Wang, X.; Gong, D.; Kanhere, P.D.; Lim, T.T.; Dong, Z.; Chen, Z. Hierarchical TiO2 nanoflakes and nanoparticles hybrid structure for improved photocatalytic activity. J. Phys. Chem. C 2012, 116, 2772–2780. [Google Scholar] [CrossRef]
  29. Gao, C.; Wei, T.; Zhang, Y.; Song, X.; Huan, Y.; Liu, H.; Zhao, M.; Yu, J.; Chen, X. A Photoresponsive Rutile TiO2 Heterojunction with Enhanced Electron–Hole Separation for High-Performance Hydrogen Evolution. Adv. Mater. 2019, 31, 1806596. [Google Scholar] [CrossRef] [PubMed]
  30. Yahya, N.; Aziz, F.; Jamaludin, N.A.; Mutalib, M.A.; Ismail, A.F.; Salleh, W.N.W.; Jaafar, J.; Yusof, N.; Ludin, N.A. A review of integrated photocatalyst adsorbents for wastewater treatment. J. Environ. Chem. Eng. 2018, 6, 7411–7425. [Google Scholar] [CrossRef]
  31. Li, X.; Yu, J.; Low, J.; Fang, Y.; Xiao, J.; Chen, X. Engineering heterogeneous semiconductors for solar water splitting. J. Mater. Chem. A 2015, 3, 2485–2534. [Google Scholar] [CrossRef]
  32. Fang, M.; Dong, G.; Wei, R.; Ho, J.C. Hierarchical Nanostructures: Hierarchical Nanostructures: Design for Sustainable Water Splitting (Adv. Energy Mater. 23/2017). Adv. Energy Mater. 2017, 7, 1770135. [Google Scholar] [CrossRef] [Green Version]
  33. Stein, A.; Rudisill, S.G.; Petkovich, N.D. Perspective on the influence of interactions between hard and soft templates and precursors on morphology of hierarchically structured porous materials. Chem. Mater. 2014, 26, 259–276. [Google Scholar] [CrossRef]
  34. Wang, P.; Xuan, J.; Zhang, R.; Zhang, H.; Wang, Q.; Wang, H.; Liu, H.; Zhang, L. Hierarchically Structured Components: Design, Additive Manufacture, and Their Energy Applications. Adv. Mater. Technol. 2022, 7, 2100672. [Google Scholar] [CrossRef]
  35. Wang, S.; Wang, Y.; Zang, S.; Lou, X.W. Hierarchical hollow heterostructures for photocatalytic CO2 reduction and water splitting. Small Methods 2020, 4, 1900586. [Google Scholar] [CrossRef]
  36. Sun, C.; Yang, J.; Xu, M.; Cui, Y.; Ren, W.; Zhang, J.; Zhao, H.; Liang, B. Recent intensification strategies of SnO2-based photocatalysts: A review. Chem. Eng. J. 2022, 427, 131564. [Google Scholar] [CrossRef]
  37. Jiang, L.; Zhou, H.; Yang, H.; Sun, N.; Huang, Z.; Pang, H. Applications of hierarchical metal–organic frameworks and their derivatives in electrochemical energy storage and conversion. J. Energy Storage 2022, 55, 105354. [Google Scholar] [CrossRef]
  38. Janani, R.; Preethi, V.R.; Singh, S.; Rani, A.; Chang, C.-T. Hierarchical ternary sulfides as effective photocatalyst for hydrogen generation through water splitting: A review on the performance of ZnIn2S4. Catalysts 2021, 11, 277. [Google Scholar] [CrossRef]
  39. Meng, S.; Wu, H.; Cui, Y.; Zheng, X.; Wang, H.; Chen, S.; Wang, Y.; Fu, X. One-step synthesis of 2D/2D-3D NiS/Zn3In2S6 hierarchical structure toward solar-to-chemical energy transformation of biomass-relevant alcohols. Appl. Catal. B Environ. 2020, 266, 118617. [Google Scholar] [CrossRef]
  40. Tahir, M.B.; Asiri, A.M.; Nabi, G.; Rafique, M.; Sagir, M. Fabrication of heterogeneous photocatalysts for insight role of carbon nanofibre in hierarchical WO3/MoSe2 composite for enhanced photocatalytic hydrogen generation. Ceram. Int. 2019, 45, 5547–5552. [Google Scholar] [CrossRef]
  41. Tan, H.; Li, J.; He, M.; Li, J.; Zhi, D.; Qin, F.; Zhang, C. Global evolution of research on green energy and environmental technologies: A bibliometric study. J. Environ. Manage. 2021, 297, 113382. [Google Scholar] [CrossRef]
  42. Jiang, Y.; Liao, J.-F.; Xu, Y.-F.; Chen, H.-Y.; Wang, X.-D.; Kuang, D.-B. Hierarchical CsPbBr3 nanocrystal-decorated ZnO nanowire/macroporous graphene hybrids for enhancing charge separation and photocatalytic CO2 reduction. J. Mater. Chem. A 2019, 7, 13762–13769. [Google Scholar] [CrossRef]
  43. Gholami, P.; Khataee, A.; Ritala, M. Template-free hierarchical trimetallic oxide photocatalyst derived from organically modified ZnCuCo layered double hydroxide. J. Clean. Prod. 2022, 366, 132761. [Google Scholar] [CrossRef]
  44. Han, Z.; Zhang, X.; Zuo, Y.; Dong, H.; Ren, H. Decorating 2D Ti3C2 on flower-like hierarchical Bi2WO6 for the 2D/2D heterojunction construction towards photodegradation of tetracycline antibiotics. Sep. Purif. Technol. 2022, 299, 121715. [Google Scholar] [CrossRef]
  45. Kaushik, B.; Rana, P.; Solanki, K.; Rawat, D.; Yadav, S.; Naikwadi, D.R.; Biradar, A.V.; Sharma, R.K. In-situ synthesis of 3-D hierarchical ZnFe2O4 modified Cu2S snowflakes: Exploring their bifunctionality in selective photocatalytic reduction of nitroarenes and methyl orange degradation. J. Photochem. Photobiol. A Chem. 2022, 433, 114165. [Google Scholar] [CrossRef]
  46. Li, Q.; Zhou, J.; Fu, L.; Chen, C.; Mao, S.; Pu, Z.; Yang, J.; Shi, J.-W.; Wu, K. Fabrication of heterostructural Ru-SrTiO3 fibers through in-situ exsolution for visible-light-induced photocatalysis. J. Alloys Compd. 2022, 925, 166747. [Google Scholar] [CrossRef]
  47. Li, X.; Yu, J.; Jaroniec, M. Hierarchical porous photocatalysts. Interface Sci. Technol. 2020, 31, 63–102. [Google Scholar]
  48. Shi, J.W.; Zong, X.; Wu, X.; Cui, H.J.; Xu, B.; Wang, L.; Fu, M.L. Carbon-doped Titania Hollow Spheres with Tunable Hierarchical Macroporous Channels and Enhanced Visible Light-induced Photocatalytic Activity. ChemCatChem 2012, 4, 488–491. [Google Scholar] [CrossRef]
  49. Yu, J.; Zhang, L.; Cheng, B.; Su, Y. Hydrothermal preparation and photocatalytic activity of hierarchically sponge-like macro-/mesoporous Titania. J. Phys. Chem. C 2007, 111, 10582–10589. [Google Scholar] [CrossRef]
  50. Wang, X.; Yu, J.C.; Ho, C.; Hou, Y.; Fu, X. Photocatalytic activity of a hierarchically macro/mesoporous titania. Langmuir 2005, 21, 2552–2559. [Google Scholar] [CrossRef] [PubMed]
  51. Zheng, Z.; Huang, B.; Qin, X.; Zhang, X.; Dai, Y. Strategic synthesis of hierarchical TiO2 microspheres with enhanced photocatalytic activity. Chem. A Eur. J. 2010, 16, 11266–11270. [Google Scholar] [CrossRef]
  52. Yu, J.; Su, Y.; Cheng, B. Template-free fabrication and enhanced photocatalytic activity of hierarchical macro-/mesoporous titania. Adv. Funct. Mater. 2007, 17, 1984–1990. [Google Scholar] [CrossRef]
  53. Zhang, L.; Yu, J.C. A sonochemical approach to hierarchical porous titania spheres with enhanced photocatalytic activity. Chem. Commun. 2003, 3, 2078–2079. [Google Scholar] [CrossRef]
  54. Shao, G.S.; Wang, F.Y.; Ren, T.Z.; Liu, Y.; Yuan, Z.Y. Hierarchical mesoporous phosphorus and nitrogen doped titania materials: Synthesis, characterization and visible-light photocatalytic activity. Appl. Catal. B Environ. 2009, 92, 61–67. [Google Scholar] [CrossRef]
  55. Dong, G.; Wang, Y.; Lei, H.; Tian, G.; Qi, S.; Wu, D. Hierarchical mesoporous titania nanoshell encapsulated on polyimide nanofiber as flexible, highly reactive, energy saving and recyclable photocatalyst for water purification. J. Clean. Prod. 2020, 253, 120021. [Google Scholar] [CrossRef]
  56. Ta, Q.T.H.; Cho, E.; Sreedhar, A.; Noh, J.S. Mixed-dimensional, three-level hierarchical nanostructures of silver and zinc oxide for fast photocatalytic degradation of multiple dyes. J. Catal. 2019, 371, 1–9. [Google Scholar] [CrossRef]
  57. Yukhnovets, O.; Semenova, A.A.; Levkevich, E.A.; Maximov, A.I.; Moshnikov, V.A. Zinc oxide hierarchical nanostructures for photocatalysis. J. Phys. Conf. Ser. 2018, 993, 012009. [Google Scholar] [CrossRef]
  58. Kim, J.H.; Joshi, M.K.; Lee, J.; Park, C.H.; Kim, C.S. Polydopamine-assisted immobilization of hierarchical zinc oxide nanostructures on electrospun nanofibrous membrane for photocatalysis and antimicrobial activity. J. Colloid Interface Sci. 2018, 513, 566–574. [Google Scholar] [CrossRef]
  59. Adhyapak, P.V.; Meshram, S.P.; Tomar, V.; Amalnerkar, D.P.; Mulla, I.S. Effect of preparation parameters on the morphologically induced photocatalytic activities of hierarchical zinc oxide nanostructures. Ceram. Int. 2013, 39, 7367–7378. [Google Scholar] [CrossRef]
  60. Wang, S.; Kuang, P.; Cheng, B.; Yu, J.; Jiang, C. ZnO hierarchical microsphere for enhanced photocatalytic activity. J. Alloys Compd. 2018, 741, 622–632. [Google Scholar] [CrossRef]
  61. Yin, Q.; Qiao, R.; Li, Z.; Zhang, X.L.; Zhu, L. Hierarchical nanostructures of nickel-doped zinc oxide: Morphology controlled synthesis and enhanced visible-light photocatalytic activity. J. Alloys Compd. 2015, 618, 318–325. [Google Scholar] [CrossRef]
  62. Magdalane, C.M.; Kaviyarasu, K.; Priyadharsini, G.M.A.; Bashir, A.K.H.; Mayedwa, N.; Matinise, N.; Isaev, A.B.; Al-Dhabi, N.A.; Arasu, M.V.; Arokiyaraj, S. Improved photocatalytic decomposition of aqueous Rhodamine-B by solar light illuminated hierarchical yttria nanosphere decorated ceria nanorods. J. Mater. Res. Technol. 2019, 8, 2898–2909. [Google Scholar] [CrossRef]
  63. Qian, J.; Chen, Z.; Sun, H.; Chen, F.; Xu, X.; Wu, Z.; Li, P.; Ge, W. Enhanced photocatalytic H2 production on three-dimensional porous CeO2/carbon nanostructure. ACS Sustain. Chem. Eng. 2018, 6, 9691–9698. [Google Scholar] [CrossRef]
  64. Xiang, Q.; Cheng, F.; Lang, D. Hierarchical layered WS2/graphene-modified CdS nanorods for efficient photocatalytic hydrogen evolution. ChemSusChem 2016, 9, 996–1002. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, W.; Yu, J.; Xiang, Q.; Cheng, B. Enhanced photocatalytic activity of hierarchical macro/mesoporous TiO2–graphene composites for photodegradation of acetone in air. Appl. Catal. B Environ. 2012, 119, 109–116. [Google Scholar] [CrossRef]
  66. Luo, Q.-P.; Yu, X.-Y.; Lei, B.-X.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. Reduced graphene oxide-hierarchical ZnO hollow sphere composites with enhanced photocurrent and photocatalytic activity. J. Phys. Chem. C 2012, 116, 8111–8117. [Google Scholar] [CrossRef]
  67. Sin, J.-C.; Lam, S.-M.; Satoshi, I.; Lee, K.-T.; Mohamed, A.R. Sunlight photocatalytic activity enhancement and mechanism of novel europium-doped ZnO hierarchical micro/nanospheres for degradation of phenol. Appl. Catal. B Environ. 2014, 148, 258–268. [Google Scholar] [CrossRef]
  68. Chen, J.; Wang, H.; Huang, G.; Zhang, Z.; Han, L.; Song, W.; Li, M.; Zhang, Y. Facile synthesis of urchin-like hierarchical Nb2O5 nanospheres with enhanced visible light photocatalytic activity. J. Alloys Compd. 2017, 728, 19–28. [Google Scholar] [CrossRef]
  69. Sang, Y.; Cao, X.; Dai, G.; Wang, L.; Peng, Y.; Geng, B. Facile one-pot synthesis of novel hierarchical Bi2O3/Bi2S3 nanoflower photocatalyst with intrinsic p-n junction for efficient photocatalytic removals of RhB and Cr(VI). J. Hazard. Mater. 2020, 381, 120942. [Google Scholar] [CrossRef]
  70. Sharma, S.; Khare, N. Hierarchical Bi2S3 nanoflowers: A novel photocatalyst for enhanced photocatalytic degradation of binary mixture of Rhodamine B and Methylene blue dyes and degradation of mixture of p-nitrophenol and p-chlorophenol. Adv. Powder Technol. 2018, 29, 3336–3347. [Google Scholar] [CrossRef]
  71. Deas, R.; Pearce, S.; Goss, K.; Wang, Q.; Chen, W.T.; Waterhouse, G.I.N. Hierarchical Au/TiO2 nanoflower photocatalysts with outstanding performance for alcohol photoreforming under UV irradiation. Appl. Catal. A Gen. 2020, 602, 39–41. [Google Scholar] [CrossRef]
  72. Cao, F.; Shi, W.; Zhao, L.; Song, S.; Yang, J.; Lei, Y.; Zhang, H. Hydrothermal synthesis and high photocatalytic activity of 3D wurtzite ZnSe hierarchical nanostruetures. J. Phys. Chem. C 2008, 112, 17095–17101. [Google Scholar] [CrossRef]
  73. Zhang, H.; Hu, C. Effective solar absorption and radial microchannels of SnO2 hierarchical structure for high photocatalytic activity. Catal. Commun. 2011, 14, 32–36. [Google Scholar] [CrossRef]
  74. Li, D.; Fang, M.; Jiang, C.; Lin, H.; Luo, C.; Qi, R.; Huang, R.; Peng, H. Size-controlled synthesis of hierarchical bismuth selenide nanoflowers and their photocatalytic performance in the presence of H2O2. J. Nanoparticle Res. 2018, 20, 228. [Google Scholar] [CrossRef]
  75. Song, J.M.; Mao, C.J.; Niu, H.L.; Shen, Y.H.; Zhang, S.Y. Hierarchical structured bismuth oxychlorides: Self-assembly from nanoplates to nanoflowers via a solvothermal route and their photocatalytic properties. CrystEngComm 2010, 12, 3875–3881. [Google Scholar] [CrossRef]
  76. Liang, Y.; Ding, M.; Yang, Y.; Xu, K.; Luo, X.; Yu, T.; Zhang, W.; Liu, W.; Yuan, C. Highly dispersed Pt nanoparticles on hierarchical titania nanoflowers with {010} facets for gas sensing and photocatalysis. J. Mater. Sci. 2019, 54, 6826–6840. [Google Scholar] [CrossRef]
  77. Malik, R.; Tomer, V.K.; Rana, P.S.; Nehra, S.P.; Duhan, S. Surfactant assisted hydrothermal synthesis of porous 3-D hierarchical SnO2 nanoflowers for photocatalytic degradation of Rose Bengal. Mater. Lett. 2015, 154, 124–127. [Google Scholar] [CrossRef]
  78. Wu, Y.; Wang, H.; Tu, W.; Liu, Y.; Wu, S.; Tan, Y.Z.; Chew, J.W. Construction of hierarchical 2D-2D Zn3In2S6/fluorinated polymeric carbon nitride nanosheets photocatalyst for boosting photocatalytic degradation and hydrogen production performance. Appl. Catal. B Environ. 2018, 233, 58–69. [Google Scholar] [CrossRef]
  79. Jin, Z.; Dong, W.; Yang, M.; Wang, J.; Gao, H.; Wang, G. One-Pot Preparation of Hierarchical Nanosheet-Constructed Fe3O4/MIL-88B(Fe) Magnetic Microspheres with High Efficiency Photocatalytic Degradation of Dye. ChemCatChem 2016, 8, 3510–3517. [Google Scholar] [CrossRef]
  80. Jia, Y.; Ma, Y.; Tang, J.; Shi, W. Hierarchical nanosheet-based Bi2MoO6 microboxes for efficient photocatalytic performance. Dalt. Trans. 2018, 47, 5542–5547. [Google Scholar] [CrossRef]
  81. Tanveer, M.; Cao, C.; Ali, Z.; Aslam, I.; Idrees, F.; Khan, W.S.; But, F.K.; Tahir, M.; Mahmood, N. Template free synthesis of CuS nanosheet-based hierarchical microspheres: An efficient natural light driven photocatalyst. CrystEngComm 2014, 16, 5290–5300. [Google Scholar] [CrossRef] [Green Version]
  82. Taheri, M.; Abdizadeh, H.; Golobostanfard, M.R. Hierarchical ZnO nanoflowers and urchin-like shapes synthesized via sol-gel electrophoretic deposition with enhanced photocatalytic performance. Mater. Chem. Phys. 2018, 220, 118–127. [Google Scholar] [CrossRef]
  83. Li, H.; Fei, G.T.; Fang, M.; Cui, P.; Guo, X.; Yan, P.; Zhang, L. De Synthesis of urchin-like Co3O4 hierarchical micro/nanostructures and their photocatalytic activity. Appl. Surf. Sci. 2011, 257, 6527–6530. [Google Scholar] [CrossRef]
  84. Yang, Y.; Wu, J.; Xiao, T.; Tang, Z.; Shen, J.; Li, H.; Zhou, Y.; Zou, Z. Urchin-like hierarchical CoZnAl-LDH/RGO/g-C3N4 hybrid as a Z-scheme photocatalyst for efficient and selective CO2 reduction. Appl. Catal. B Environ. 2019, 255, 117771. [Google Scholar] [CrossRef]
  85. Liu, J.; Guo, Z.; Wang, W.; Huang, Q.; Zhu, K.; Chen, X. Heterogeneous ZnS hollow urchin-like hierarchical nanostructures and their structure-enhanced photocatalytic properties. Nanoscale 2011, 3, 1470–1473. [Google Scholar] [CrossRef]
  86. Xiang, L.; Zhao, X.; Yin, J.; Fan, B. Well-organized 3D urchin-like hierarchical TiO2 microspheres with high photocatalytic activity. J. Mater. Sci. 2012, 47, 1436–1445. [Google Scholar] [CrossRef]
  87. Cheng, P.; Wang, Y.; Xu, L.; Sun, P.; Su, Z.; Jin, F.; Liu, F.; Sun, Y.; Lu, G. High specific surface area urchin-like hierarchical ZnO-TiO2 architectures: Hydrothermal synthesis and photocatalytic properties. Mater. Lett. 2016, 175, 52–55. [Google Scholar] [CrossRef]
  88. Xiao, X.; Xing, C.; He, G.; Zuo, X.; Nan, J.; Wang, L. Solvothermal synthesis of novel hierarchical Bi4O5I2 nanoflakes with highly visible light photocatalytic performance for the degradation of 4-tert-butylphenol. Appl. Catal. B Environ. 2014, 148, 154–163. [Google Scholar] [CrossRef]
  89. Ong, W.L.; Natarajan, S.; Kloostra, B.; Ho, G.W. Metal nanoparticle-loaded hierarchically assembled ZnO nanoflakes for enhanced photocatalytic performance. Nanoscale 2013, 5, 5568–5575. [Google Scholar] [CrossRef]
  90. Sun, M.; Chen, C.; Chen, L.; Su, B. Hierarchically porous materials: Synthesis strategies and emerging applications. Front. Chem. Sci. Eng. 2016, 10, 301–347. [Google Scholar] [CrossRef]
  91. Yang, Q.; Hao, J. Synthesis of metal sulfides via ionic liquid-mediated assembly strategy and their photocatalytic degradation of dyes in water. Colloids Surf. A Physicochem. Eng. Asp. 2022, 633, 127848. [Google Scholar] [CrossRef]
  92. Sang, Y.; Ding, G.; Guo, Z.; Xue, Y.; Li, G.; Zhang, R. Facile synthesis of amorphous bimetallic hydroxide on Fe-doped Ni3S2 as an active electrocatalyst for oxygen evolution reaction. J. Alloys Compd. 2022, 919, 165855. [Google Scholar] [CrossRef]
  93. Xu, T.; Wang, S.; Li, L.; Liu, X. Dual templated synthesis of tri-modal porous SrTiO3/TiO2@ carbon composites with enhanced photocatalytic activity. Appl. Catal. A Gen. 2019, 575, 132–141. [Google Scholar] [CrossRef]
  94. Xu, T.; Liu, X.; Wang, S.; Li, L. Ferroelectric oxide nanocomposites with trimodal pore structure for high photocatalytic performance. Nano-micro Lett. 2019, 11, 37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Yao, L.; Wang, H.; Zhang, Y.; Wang, S.; Liu, X. Fabrication of N doped TiO2/C nanocomposites with hierarchical porous structure and high photocatalytic activity. Microporous Mesoporous Mater. 2019, 288, 109604. [Google Scholar] [CrossRef]
  96. Liu, H.; Liu, X.; Mu, S.; Wang, S.; Wang, S.; Li, L.; Giannelis, E.P. A novel fabrication approach for three-dimensional hierarchical porous metal oxide/carbon nanocomposites for enhanced solar photocatalytic performance. Catal. Sci. Technol. 2017, 7, 1965–1970. [Google Scholar] [CrossRef]
  97. Lyu, J.; Shao, J.; Wang, Y.; Qiu, Y.; Li, J.; Li, T.; Peng, Y.; Liu, F. Construction of a porous core-shell homojunction for the photocatalytic degradation of antibiotics. Chem. Eng. J. 2019, 358, 614–620. [Google Scholar] [CrossRef]
  98. Zhao, J.; Liao, C.; Chen, X.; Song, W. Hierarchically ordered macro–mesoporous anatase TiO2 prepared by pearl oyster shell and triblock copolymer dual templates for high photocatalytic activity. RSC Adv. 2018, 8, 38461–38469. [Google Scholar] [CrossRef] [Green Version]
  99. Rana, A.; Sudhaik, A.; Raizada, P.; Nguyen, V.-H.; Xia, C.; Khan, A.A.P.; Thakur, S.; Nguyen-Tri, P.; Nguyen, C.C.; Kim, S.Y. Graphitic carbon nitride based immobilized and non-immobilized floating photocatalysts for environmental remediation. Chemosphere 2022, 297, 134229. [Google Scholar] [CrossRef]
  100. Xu, M.; Wu, J.; Wang, J.; Mao, Y.; Liu, M.; Yang, Y.; Yang, C.; Sun, L.; Du, Y.; Li, Y. The integration of Triazine-based porous organic polymer with bio-waste poplar catkin as water-floatable photocatalyst. Appl. Surf. Sci. 2022, 581, 152409. [Google Scholar] [CrossRef]
  101. Chen, H.; Zhao, L.; He, X.; Huang, Z.; Wang, G.; Fang, W.; Li, W. Highly enhanced photocatalytic degradation of dye rhodamine B on the biogenic hierarchical porous TiO2/SiO2 with 1D/3D-chain. Clean Technol. Environ. Policy 2018, 20, 887–897. [Google Scholar] [CrossRef]
  102. CHEN, F.; CHEN, X.; CHEN, Z. Artificial Synthesis and Improved Photo Catalysis Performance of Bionic Hierarchical Fibers Alpha-Fe2O3; 2nd International Seminar on Applied Physics, Optoelectronics and Photonics (APOP): Suzhou, China, 2017; ISBN 978-1-60595-522-3. [Google Scholar]
  103. Lv, H.; Zhang, M.; Wang, P.; Xu, X.; Liu, Y.; Yu, D.-G. Ingenious construction of Ni(DMG)2/TiO2-decorated porous nanofibers for the highly efficient photodegradation of pollutants in water. Colloids Surf. A Physicochem. Eng. Asp. 2022, 650, 129561. [Google Scholar] [CrossRef]
  104. He, D.; Ma, Y.; Zhao, R.; Qiu, J.; Sun, B.; Wang, H.; Yang, M.; Jia, X.; Li, X.; Li, Y. Complete-lifecycle-available, lightweight and flexible hierarchical structured Bi2WO6/WO3/PAN nanofibrous membrane for X-Ray shielding and photocatalytic degradation. Adv. Mater. Interfaces 2021, 8, 2002131. [Google Scholar] [CrossRef]
  105. Zhang, J.; Hou, X.; Pang, Z.; Cai, Y.; Zhou, H.; Lv, P.; Wei, Q. Fabrication of hierarchical TiO2 nanofibers by microemulsion electrospinning for photocatalysis applications. Ceram. Int. 2017, 43, 15911–15917. [Google Scholar] [CrossRef]
  106. Zhou, H.; Zhong, S.; Shen, M.; Yao, Y. Composite soft template-assisted construction of a flower-like β-Bi2O3/Bi2O2CO3 heterojunction photocatalyst for the enhanced simulated sunlight photocatalytic degradation of tetracycline. Ceram. Int. 2019, 45, 15036–15047. [Google Scholar] [CrossRef]
  107. Lin, Z.; Huang, J. A hierarchical H3PW12O40/TiO2 nanocomposite with cellulose as scaffold for photocatalytic degradation of organic pollutants. Sep. Purif. Technol. 2021, 264, 118427. [Google Scholar] [CrossRef]
  108. Sun, X.; Wang, K.; Shu, Y.; Zou, F.; Zhang, B.; Sun, G.; Uyama, H.; Wang, X. One-pot route towards active TiO2 doped hierarchically porous cellulose: Highly efficient photocatalysts for methylene blue degradation. Materials 2017, 10, 373. [Google Scholar] [CrossRef] [Green Version]
  109. Xiao, J.; Jiang, H. Thermally Stable Metal-Organic Framework-Templated Synthesis of Hierarchically Porous Metal Sulfides: Enhanced Photocatalytic Hydrogen Production. Small 2017, 13, 1700632. [Google Scholar] [CrossRef]
  110. Ren, J.-T.; Zheng, Y.-L.; Yuan, K.; Zhou, L.; Wu, K.; Zhang, Y.-W. Self-templated synthesis of Co3O4 hierarchical nanosheets from a metal–organic framework for efficient visible-light photocatalytic CO2 reduction. Nanoscale 2020, 12, 755–762. [Google Scholar] [CrossRef]
  111. Wang, F.; Xiao, L.; Chen, J.; Chen, L.; Fang, R.; Li, Y. Regulating the Electronic Structure and Water Adsorption Capability by Constructing Carbon-Doped CuO Hollow Spheres for Efficient Photocatalytic Hydrogen Evolution. ChemSusChem 2020, 13, 5711–5721. [Google Scholar] [CrossRef]
  112. Li, P.; Zhang, M.; Li, X.; Wang, C.; Wang, R.; Wang, B.; Yan, H. MOF-derived NiO/CeO2 heterojunction: A photocatalyst for degrading pollutants and hydrogen evolution. J. Mater. Sci. 2020, 55, 15930–15944. [Google Scholar] [CrossRef]
  113. Chen, W.; Han, B.; Xie, Y.; Liang, S.; Deng, H.; Lin, Z. Ultrathin Co-Co LDHs nanosheets assembled vertically on MXene: 3D nanoarrays for boosted visible-light-driven CO2 reduction. Chem. Eng. J. 2020, 391, 123519. [Google Scholar] [CrossRef]
  114. Yan, B.; Zhang, L.; Tang, Z.; Al-Mamun, M.; Zhao, H.; Su, X. Palladium-decorated hierarchical titania constructed from the metal-organic frameworks NH2-MIL-125 (Ti) as a robust photocatalyst for hydrogen evolution. Appl. Catal. B Environ. 2017, 218, 743–750. [Google Scholar] [CrossRef]
  115. Zhao, F.; Yin, D.; Khaing, K.K.; Liu, B.; Chen, T.; Deng, L.; Li, L.; Guo, X.; Wang, J.; Xiao, S. Fabrication of hierarchical Co9S8@ ZnAgInS heterostructured cages for highly efficient photocatalytic hydrogen generation and pollutants degradation. Inorg. Chem. 2020, 59, 7027–7038. [Google Scholar] [CrossRef] [PubMed]
  116. Peng, D.; Zhang, Y.; Xu, G.; Tian, Y.; Ma, D.; Zhang, Y.; Qiu, P. Synthesis of multilevel structured MoS2@ Cu/Cu2O@ C visible-light-driven photocatalyst derived from MOF–guest polyhedra for cyclohexane oxidation. ACS Sustain. Chem. Eng. 2020, 8, 6622–6633. [Google Scholar] [CrossRef]
  117. Gong, Y.; Zhao, X.; Zhang, H.; Yang, B.; Xiao, K.; Guo, T.; Zhang, J.; Shao, H.; Wang, Y.; Yu, G. MOF-derived nitrogen doped carbon modified g-C3N4 heterostructure composite with enhanced photocatalytic activity for bisphenol A degradation with peroxymonosulfate under visible light irradiation. Appl. Catal. B Environ. 2018, 233, 35–45. [Google Scholar] [CrossRef]
  118. He, X.; Nguyen, V.; Jiang, Z.; Wang, D.; Zhu, Z.; Wang, W.-N. Highly-oriented one-dimensional MOF-semiconductor nanoarrays for efficient photodegradation of antibiotics. Catal. Sci. Technol. 2018, 8, 2117–2123. [Google Scholar] [CrossRef]
  119. Xu, Y.; Wu, S.; Li, X.; Huang, Y.; Wang, Z.; Han, Y.; Wu, J.; Meng, H.; Zhang, X. Synthesis, characterization, and photocatalytic degradation properties of ZnO/ZnFe2O4 magnetic heterostructures. New J. Chem. 2017, 41, 15433–15438. [Google Scholar] [CrossRef]
  120. Kampouri, S.; Ireland, C.P.; Valizadeh, B.; Oveisi, E.; Schouwink, P.A.; Mensi, M.; Stylianou, K.C. Mixed-phase MOF-derived titanium dioxide for photocatalytic hydrogen evolution: The impact of the templated morphology. ACS Appl. Energy Mater. 2018, 1, 6541–6548. [Google Scholar] [CrossRef]
  121. Li, N.; Huang, H.; Bibi, R.; Shen, Q.; Ngulube, R.; Zhou, J.; Liu, M. Noble-metal-free MOF derived hollow CdS/TiO2 decorated with NiS cocatalyst for efficient photocatalytic hydrogen evolution. Appl. Surf. Sci. 2019, 476, 378–386. [Google Scholar] [CrossRef]
  122. Mao, Q.; Chen, J.; Chen, H.; Chen, Z.; Chen, J.; Li, Y. Few-layered 1T-MoS2-modified ZnCoS solid-solution hollow dodecahedra for enhanced photocatalytic hydrogen evolution. J. Mater. Chem. A 2019, 7, 8472–8484. [Google Scholar] [CrossRef]
  123. Yang, S.J.; Im, J.H.; Kim, T.; Lee, K.; Park, C.R. MOF-derived ZnO and ZnO@C composites with high photocatalytic activity and adsorption capacity. J. Hazard. Mater. 2011, 186, 376–382. [Google Scholar] [CrossRef]
  124. Dolgopolova, E.A.; Shustova, N.B. Metal–organic framework photophysics: Optoelectronic devices, photoswitches, sensors, and photocatalysts. MRS Bull. 2016, 41, 890–896. [Google Scholar] [CrossRef]
  125. Li, M.; Song, S.; Su, C.; Li, L.; Yan, Z.; Cao, X. MOF-templated in situ fabrication of surface-modified Ni/graphitic carbon nitride with enhanced photocatalytic hydrogen evolution. Catal. Sci. Technol. 2019, 9, 3828–3835. [Google Scholar] [CrossRef]
  126. Wang, H. Rational design W-doped Co-ZIF-9 based Co3S4 composite photocatalyst for efficient visible-light-driven photocatalytic H 2 evolution. Sustain. Energy Fuels 2019, 3, 173–183. [Google Scholar] [CrossRef]
  127. Mondal, I.; Gonuguntla, S.; Pal, U. Photoinduced fabrication of Cu/TiO2 core–shell heterostructures derived from Cu-MOF for solar hydrogen generation: The size of the Cu nanoparticle matters. J. Phys. Chem. C 2019, 123, 26073–26081. [Google Scholar] [CrossRef]
  128. Liu, H.; Li, J.; Chen, Y.; Sun, X.; Xu, X.; Qiu, L.; Duo, S.; Li, P. Ternary photocatalysts based on MOF-derived TO2 co-decorated with ZnIn2S4 nanosheets and CdS nanoparticles for effective visible light degradation of organic pollutants. New J. Chem. 2022, 46, 7195–7201. [Google Scholar] [CrossRef]
  129. Gao, L.-M.; Zhao, J.-H.; Li, T.; Li, R.; Xie, H.-Q.; Zhu, P.-L.; Niu, X.-Y.; Li, K. High-performance TO2 photocatalyst produced by the versatile functions of the tiny bimetallic MOF-derived NiCoS-porous carbon cocatalyst. CrystEngComm 2019, 21, 3686–3693. [Google Scholar] [CrossRef]
  130. Zhang, C.; Liu, B.; Cheng, X.; Guo, Z.; Zhuang, T.; Lv, Z. Noble-Metal-Free CdS Decorated Porous NixCo1–xO Skeleton Derived from Metal–Organic Framework for Efficient Visible-Light H2 Production. ACS Appl. Energy Mater. 2019, 3, 852–860. [Google Scholar] [CrossRef] [Green Version]
  131. Tan, J.; Yu, M.; Cai, Z.; Lou, X.; Wang, J.; Li, Z. MOF-derived synthesis of MnS/In2S3 pn heterojunctions with hierarchical structures for efficient photocatalytic CO2 reduction. J. Colloid Interface Sci. 2021, 588, 547–556. [Google Scholar] [CrossRef]
  132. Mondal, I.; Pal, U. Synthesis of MOF templated Cu/CuO@ TO2 nanocomposites for synergistic hydrogen production. Phys. Chem. Chem. Phys. 2016, 18, 4780–4788. [Google Scholar] [CrossRef] [PubMed]
  133. Wu, S.; Zhao, H.-J.; Li, C.-F.; Liu, J.; Dong, W.; Zhao, H.; Wang, C.; Liu, Y.; Hu, Z.-Y.; Chen, L. Type II heterojunction in hierarchically porous zinc oxide/graphitic carbon nitride microspheres promoting photocatalytic activity. J. Colloid Interface Sci. 2019, 538, 99–107. [Google Scholar] [CrossRef]
  134. Chen, X.; Li, J.-J.; Chen, X.; Cai, S.-C.; Yu, E.-Q.; Chen, J.; Jia, H. MOF-Templated Approach for Hollow NiOx/Co3O4 Catalysts: Enhanced Light-Driven Thermocatalytic Degradation of Toluene. ACS Appl. Nano Mater. 2018, 1, 2971–2981. [Google Scholar] [CrossRef]
  135. Chen, W.; Fang, J.; Zhang, Y.; Chen, G.; Zhao, S.; Zhang, C.; Xu, R.; Bao, J.; Zhou, Y.; Xiang, X. CdS nanosphere-decorated hollow polyhedral ZCO derived from a metal–organic framework (MOF) for effective photocatalytic water evolution. Nanoscale 2018, 10, 4463–4474. [Google Scholar] [CrossRef]
  136. Yang, Y.; Sun, L.; Zhan, W.; Wang, X.; Han, X. Separated redox site strategies for engineering highly efficient photocatalysts: A pagoda-like In2O3/CuO heteroepitaxial structure coated with a N-doped C layer. J. Mater. Chem. A 2021, 9, 4310–4316. [Google Scholar] [CrossRef]
  137. Jiang, Z.; Feng, L.; Zhu, J.; Liu, B.; Li, X.; Chen, Y.; Khan, S. Construction of a hierarchical NiFe2O4/CuInSe2 (pn) heterojunction: Highly efficient visible-light-driven photocatalyst in the degradation of endocrine disruptors in an aqueous medium. Ceram. Int. 2021, 47, 8996–9007. [Google Scholar] [CrossRef]
  138. Zhang, L.; Feng, L.; Li, P.; Chen, X.; Jiang, J.; Zhang, S.; Zhang, C.; Zhang, A.; Chen, G.; Wang, H. Direct Z-scheme photocatalyst of hollow CoSx@ CdS polyhedron constructed by ZIF-67-templated one-pot solvothermal route: A signal-on photoelectrochemical sensor for mercury (II). Chem. Eng. J. 2020, 395, 125072. [Google Scholar] [CrossRef]
  139. Al-Hajji, L.A.; Ismail, A.A.; Bumajdad, A.; Alsaidi, M.; Ahmed, S.A.; Almutawa, F.; Al-Hazza, A. Construction of Au/TiO2 Heterojunction with high photocatalytic performances under UVA illumination. Ceram. Int. 2020, 46, 20155–20162. [Google Scholar] [CrossRef]
  140. Li, T.; Cui, J.-D.; Xu, M.-L.; Li, R.; Gao, L.-M.; Zhu, P.-L.; Xie, H.-Q.; Li, K. Engineering a hetero-MOF-derived TiO2–Co3O4 heterojunction decorated with nickel nanoparticles for enhanced photocatalytic activity even in pure water. CrystEngComm 2020, 22, 5620–5627. [Google Scholar] [CrossRef]
  141. Bera, S.; Pal, M.; Naskar, A.; Jana, S. Hierarchically structured ZnO-graphene hollow microspheres towards effective reusable adsorbent for organic pollutant via photodegradation process. J. Alloys Compd. 2016, 669, 177–186. [Google Scholar] [CrossRef]
  142. Zhou, L.; Han, Z.; Li, G.-D.; Zhao, Z. Template-free synthesis and photocatalytic activity of hierarchical hollow ZnO microspheres composed of radially aligned nanorods. J. Phys. Chem. Solids 2021, 148, 109719. [Google Scholar] [CrossRef]
  143. Bian, Z.; Zhu, J.; Li, H. Solvothermal alcoholysis synthesis of hierarchical TiO2 with enhanced activity in environmental and energy photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2016, 28, 72–86. [Google Scholar] [CrossRef]
  144. Mamaghani, A.H.; Haghighat, F.; Lee, C.-S. Photocatalytic oxidation of MEK over hierarchical TiO2 catalysts: Effect of photocatalyst features and operating conditions. Appl. Catal. B Environ. 2019, 251, 1–16. [Google Scholar] [CrossRef]
  145. Hiremath, V.; Deonikar, V.G.; Kim, H.; Seo, J.G. Hierarchically assembled porous TiO2 nanoparticles with enhanced photocatalytic activity towards Rhodamine-B degradation. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 586, 124199. [Google Scholar] [CrossRef]
  146. Harris, J.; Silk, R.; Smith, M.; Dong, Y.; Chen, W.-T.; Waterhouse, G.I.N. Hierarchical TiO2 nanoflower photocatalysts with remarkable activity for aqueous methylene blue photo-oxidation. ACS Omega 2020, 5, 18919–18934. [Google Scholar] [CrossRef] [PubMed]
  147. Singh, S.; Lo, S.-L. Single-phase cerium oxide nanospheres: An efficient photocatalyst for the abatement of rhodamine B dye. Environ. Sci. Pollut. Res. 2018, 25, 6532–6544. [Google Scholar] [CrossRef]
  148. Sekar, K.; Chuaicham, C.; Balijapalli, U.; Li, W.; Wilson, K.; Lee, A.F.; Sasaki, K. Surfactant-and template-free hydrothermal assembly of Cu2O visible light photocatalysts for trimethoprim degradation. Appl. Catal. B Environ. 2021, 284, 119741. [Google Scholar] [CrossRef]
  149. Rong, R.; Wang, L. Synthesis of hierarchical hollow nest-like WO3 micro/nanostructures with enhanced visible light-driven photocatalytic activity. J. Alloys Compd. 2021, 850, 156742. [Google Scholar] [CrossRef]
  150. Wan, J.; Du, X.; Wang, R.; Liu, E.; Jia, J.; Bai, X.; Hu, X.; Fan, J. Mesoporous nanoplate multi-directional assembled Bi2WO6 for high efficient photocatalytic oxidation of NO. Chemosphere 2018, 193, 737–744. [Google Scholar] [CrossRef]
  151. Xu, B.; Yang, H.; Zhang, Q.; Yuan, S.; Xie, A.; Zhang, M.; Ohno, T. Design and Synthesis of Sm, Y, La and Nd-doped CeO2 with a broom-like hierarchical structure: A photocatalyst with enhanced oxidation performance. ChemCatChem 2020, 12, 2638–2646. [Google Scholar] [CrossRef]
  152. Dai, G.; Qin, H.; Zhou, H.; Wang, W.; Luo, T. Template-free fabrication of hierarchical macro/mesoporpous SnS2/TiO2 composite with enhanced photocatalytic degradation of Methyl Orange (MO). Appl. Surf. Sci. 2018, 430, 488–495. [Google Scholar] [CrossRef]
  153. Zhu, L.; Liu, Z.; Xia, P.; Li, H.; Xie, Y. Synthesis of hierarchical ZnO&Graphene composites with enhanced photocatalytic activity. Ceram. Int. 2018, 44, 849–856. [Google Scholar]
  154. Liu, W.; Xu, Y.; Zhou, W.; Zhang, X.; Cheng, X.; Zhao, H.; Gao, S.; Huo, L. A facile synthesis of hierarchically porous TiO2 microspheres with carbonaceous species for visible-light photocatalysis. J. Mater. Sci. Technol. 2017, 33, 39–46. [Google Scholar] [CrossRef]
  155. Liang, Z.; Bai, X.; Hao, P.; Guo, Y.; Xue, Y.; Tian, J.; Cui, H. Full solar spectrum photocatalytic oxygen evolution by carbon-coated TiO2 hierarchical nanotubes. Appl. Catal. B Environ. 2019, 243, 711–720. [Google Scholar] [CrossRef]
  156. Li, Z.; Chen, H.; Li, Y.; Wang, H.; Liu, Y.; Li, X.; Lin, H.; Li, S.; Wang, L. Porous direct Z-scheme heterostructures of S-deficient CoS/CdS hexagonal nanoplates for robust photocatalytic H2 generation. CrystEngComm 2022, 24, 404–416. [Google Scholar] [CrossRef]
  157. Chen, L.; Meng, D.; Wu, X.; Wang, A.; Wang, J.; Yu, M.; Liang, Y. Enhanced visible light photocatalytic performances of self-assembled hierarchically structured BiVO4/Bi2WO6 heterojunction composites with different morphologies. RSC Adv. 2016, 6, 52300–52309. [Google Scholar] [CrossRef]
  158. Zhen, M.; Li, K.; Guo, S.-Q.; Li, H.; Shen, B. Template-free construction of hollow TiO2 microspheres for long-life and high-capacity lithium storage. J. Alloys Compd. 2021, 859, 157761. [Google Scholar] [CrossRef]
  159. Pan, B.; Zhou, Y.; Su, W.; Wang, X. Self-assembly synthesis of LaPO4 hierarchical hollow spheres with enhanced photocatalytic CO2-reduction performance. Nano Res. 2017, 10, 534–545. [Google Scholar] [CrossRef]
  160. Lee, S.C.; Jeong, Y.; Kim, Y.J.; Kim, H.; Lee, H.U.; Lee, Y.-C.; Lee, S.M.; Kim, H.J.; An, H.-R.; Ha, M.G. Hierarchically three-dimensional (3D) nanotubular sea urchin-shaped iron oxide and its application in heavy metal removal and solar-induced photocatalytic degradation. J. Hazard. Mater. 2018, 354, 283–292. [Google Scholar] [CrossRef]
  161. Chen, X.; Zhang, H.; Zhang, D.; Miao, Y.; Li, G. Controllable synthesis of mesoporous multi-shelled ZnO microspheres as efficient photocatalysts for NO oxidation. Appl. Surf. Sci. 2018, 435, 468–475. [Google Scholar] [CrossRef]
  162. Zhu, L.; Li, H.; Xia, P.; Liu, Z.; Xiong, D. Hierarchical ZnO decorated with CeO2 nanoparticles as the direct Z-scheme heterojunction for enhanced photocatalytic activity. ACS Appl. Mater. Interfaces 2018, 10, 39679–39687. [Google Scholar] [CrossRef] [PubMed]
  163. Anjugam Vandarkuzhali, S.A.; Pugazhenthiran, N.; Mangalaraja, R.V.; Sathishkumar, P.; Viswanathan, B.; Anandan, S. Ultrasmall plasmonic nanoparticles decorated hierarchical mesoporous TiO2 as an efficient photocatalyst for photocatalytic degradation of textile dyes. ACS Omega 2018, 3, 9834–9845. [Google Scholar] [CrossRef] [Green Version]
  164. Cai, Y.; Song, J.; Liu, X.; Yin, X.; Li, X.; Yu, J.; Ding, B. Soft BiOBr@ TiO2 nanofibrous membranes with hierarchical heterostructures as efficient and recyclable visible-light photocatalysts. Environ. Sci. Nano 2018, 5, 2631–2640. [Google Scholar] [CrossRef]
  165. Jung, H.; Cho, K.M.; Kim, K.H.; Yoo, H.-W.; Al-Saggaf, A.; Gereige, I.; Jung, H.-T. Highly efficient and stable CO2 reduction photocatalyst with a hierarchical structure of mesoporous TiO2 on 3D graphene with few-layered MoS2. ACS Sustain. Chem. Eng. 2018, 6, 5718–5724. [Google Scholar] [CrossRef]
  166. Qu, X.; Yang, R.; Tong, F.; Zhao, Y.; Wang, M.-H. Hierarchical ZnO microstructures decorated with Au nanoparticles for enhanced gas sensing and photocatalytic properties. Powder Technol. 2018, 330, 259–265. [Google Scholar] [CrossRef]
  167. Nie, N.; He, F.; Zhang, L.; Cheng, B. Direct Z-scheme PDA-modified ZnO hierarchical microspheres with enhanced photocatalytic CO2 reduction performance. Appl. Surf. Sci. 2018, 457, 1096–1102. [Google Scholar] [CrossRef]
  168. Jo, W.-K.; Lee, J.Y.; Natarajan, T.S. Fabrication of hierarchically structured novel redox-mediator-free ZnIn2S4 marigold flower/Bi2WO6 flower-like direct Z-scheme nanocomposite photocatalysts with superior visible light photocatalytic efficiency. Phys. Chem. Chem. Phys. 2016, 18, 1000–1016. [Google Scholar] [CrossRef] [PubMed]
  169. Song, L.; Pang, Y.; Zheng, Y.; Chen, C.; Ge, L. Design, preparation and enhanced photocatalytic activity of porous BiOCl/BiVO4 microspheres via a coprecipitation-hydrothermal method. J. Alloys Compd. 2017, 710, 375–382. [Google Scholar] [CrossRef]
  170. Madhusudan, P.; Zhang, J.; Yu, J.; Cheng, B.; Xu, D.; Zhang, J. One-pot template-free synthesis of porous CdMoO4 microspheres and their enhanced photocatalytic activity. Appl. Surf. Sci. 2016, 387, 202–213. [Google Scholar] [CrossRef]
  171. Nie, N.; Zhang, L.; Fu, J.; Cheng, B.; Yu, J. Self-assembled hierarchical direct Z-scheme g-C3N4/ZnO microspheres with enhanced photocatalytic CO2 reduction performance. Appl. Surf. Sci. 2018, 441, 12–22. [Google Scholar] [CrossRef]
  172. He, R.; Zhang, J.; Yu, J.; Cao, S. Room-temperature synthesis of BiOI with tailorable (0 0 1) facets and enhanced photocatalytic activity. J. Colloid Interface Sci. 2016, 478, 201–208. [Google Scholar] [CrossRef]
  173. Wei, H.; Hou, C.; Zhang, Y.; Nan, Z. Scalable low temperature in air solid phase synthesis of porous flower-like hierarchical nanostructure SnS2 with superior performance in the adsorption and photocatalytic reduction of aqueous Cr (VI). Sep. Purif. Technol. 2017, 189, 153–161. [Google Scholar] [CrossRef]
  174. Sedlak, J.; Kuritka, I.; Masar, M.; Machovsky, M.; Urbanek, P.; Bazant, P.; Janota, P.; Dvorackova, M. Contributions of morphological and structural parameters at different hierarchical morphology levels to photocatalytic activity of mesoporous nanostructured ZnO. Appl. Surf. Sci. 2020, 513, 145773. [Google Scholar] [CrossRef]
  175. Fu, J.; Zhu, B.; Jiang, C.; Cheng, B.; You, W.; Yu, J. Hierarchical porous O-doped g-C3N4 with enhanced photocatalytic CO2 reduction activity. Small 2017, 13, 1603938. [Google Scholar] [CrossRef]
  176. Wang, Y.; Li, Y.; Bai, X.; Cai, Q.; Liu, C.; Zuo, Y.; Kang, S.; Cui, L. Facile synthesis of Y-doped graphitic carbon nitride with enhanced photocatalytic performance. Catal. Commun. 2016, 84, 179–182. [Google Scholar] [CrossRef]
  177. Liu, M.; Wageh, S.; Al-Ghamdi, A.A.; Xia, P.; Cheng, B.; Zhang, L.; Yu, J. Quenching induced hierarchical 3D porous gC3N4 with enhanced photocatalytic CO2 reduction activity. Chem. Commun. 2019, 55, 14023–14026. [Google Scholar] [CrossRef]
  178. Shabani, M.; Haghighi, M.; Kahforoushan, D.; Haghighi, A. Mesoporous-mixed-phase of hierarchical bismuth oxychlorides nanophotocatalyst with enhanced photocatalytic application in treatment of antibiotic effluents. J. Clean. Prod. 2019, 207, 444–457. [Google Scholar] [CrossRef]
  179. Edla, R.; Tonezzer, A.; Orlandi, M.; Patel, N.; Fernandes, R.; Bazzanella, N.; Date, K.; Kothari, D.C.; Miotello, A. 3D hierarchical nanostructures of iron oxides coatings prepared by pulsed laser deposition for photocatalytic water purification. Appl. Catal. B Environ. 2017, 219, 401–411. [Google Scholar] [CrossRef]
  180. Hu, H.; Wang, J.; Deng, C.; Niu, C.; Le, H. Microwave-assisted controllable synthesis of hierarchical CuS nanospheres displaying fast and efficient photocatalytic activities. J. Mater. Sci. 2018, 53, 14250–14261. [Google Scholar] [CrossRef]
  181. Nethravathi, C.; Rajamathi, J.T.; Rajamathi, M. Microwave-assisted synthesis of porous aggregates of CuS nanoparticles for sunlight photocatalysis. ACS Omega 2019, 4, 4825–4831. [Google Scholar] [CrossRef] [Green Version]
  182. Huang, Z.; Yan, F.-W.; Yuan, G. Ultrasound-assisted fabrication of hierarchical rodlike graphitic carbon nitride with fewer defects and enhanced visible-light photocatalytic activity. ACS Sustain. Chem. Eng. 2018, 6, 3187–3195. [Google Scholar] [CrossRef]
  183. Xu, W.; Wang, S.Q.; Zhang, Q.Y.; Ma, C.Y.; Wang, Q.; Wen, D.H.; Li, X.N. Hierarchically structured AgO films with nano-porosity for photocatalyst and all solid-state thin film battery. J. Alloys Compd. 2019, 802, 210–216. [Google Scholar] [CrossRef]
  184. Zhong, G.; Liu, D. Biomorphic CdS/Cd Photocatalyst Derived from Butterfly Wing for Efficient Hydrogen Evolution Reaction. Adv. Mater. Interfaces 2022, 9, 2102423. [Google Scholar] [CrossRef]
Figure 1. Main advantages of hierarchical porous structures in photocatalysis.
Figure 1. Main advantages of hierarchical porous structures in photocatalysis.
Catalysts 12 01152 g001
Figure 2. Schematic illustration of the light-harvesting behavior of heterogeneous photocatalysts with different nanostructures [47].
Figure 2. Schematic illustration of the light-harvesting behavior of heterogeneous photocatalysts with different nanostructures [47].
Catalysts 12 01152 g002
Figure 3. Main Synthesis Methods to realize hierarchical nanostructures.
Figure 3. Main Synthesis Methods to realize hierarchical nanostructures.
Catalysts 12 01152 g003
Figure 4. Schematic of the experimental setup [93].
Figure 4. Schematic of the experimental setup [93].
Catalysts 12 01152 g004
Figure 5. Scheme of floating triazine--based porous organic polymer (POP) photocatalyst grown on poplar catkins and used for water multipurpose water decontamination [100].
Figure 5. Scheme of floating triazine--based porous organic polymer (POP) photocatalyst grown on poplar catkins and used for water multipurpose water decontamination [100].
Catalysts 12 01152 g005
Figure 6. The mechanism of improved photocatalytic performances for hierarchical TiO2 nanofibers [105].
Figure 6. The mechanism of improved photocatalytic performances for hierarchical TiO2 nanofibers [105].
Catalysts 12 01152 g006
Figure 7. Schematic representation of the band structure of the β-Bi2O3/Bi2O2CO3 heterojunction and the migration of photogenerated charges under simulated sunlight irradiation [106].
Figure 7. Schematic representation of the band structure of the β-Bi2O3/Bi2O2CO3 heterojunction and the migration of photogenerated charges under simulated sunlight irradiation [106].
Catalysts 12 01152 g007
Figure 8. Scheme for hierarchical heterostructured Co9S8@98ZnAgInS photocatalyst: fabrication through extra situ approach through (I) sulfidation reaction and thermal treatment in a nitrogen atmosphere and (II) deposition of ZnAgInS nanosheets [115].
Figure 8. Scheme for hierarchical heterostructured Co9S8@98ZnAgInS photocatalyst: fabrication through extra situ approach through (I) sulfidation reaction and thermal treatment in a nitrogen atmosphere and (II) deposition of ZnAgInS nanosheets [115].
Catalysts 12 01152 g008
Figure 9. SEM micrographs of some hierarchical photocatalysts obtained with hybrid templates [131,133,134].
Figure 9. SEM micrographs of some hierarchical photocatalysts obtained with hybrid templates [131,133,134].
Catalysts 12 01152 g009
Figure 10. Scheme of (A) the synthetic route of hollow CoSx polyhedrons and CoSx@CdS composites; (B) Z scheme charge migration mechanism CoSx@CdS composites and double CoSx@CdS/HgS heterojunction under the visible-light illumination [138].
Figure 10. Scheme of (A) the synthetic route of hollow CoSx polyhedrons and CoSx@CdS composites; (B) Z scheme charge migration mechanism CoSx@CdS composites and double CoSx@CdS/HgS heterojunction under the visible-light illumination [138].
Catalysts 12 01152 g010
Figure 11. Synthesis scheme of heterostructure Co-Co LDH/TNS nanosheets [113].
Figure 11. Synthesis scheme of heterostructure Co-Co LDH/TNS nanosheets [113].
Catalysts 12 01152 g011
Figure 12. Schematic illustration of the synthetic procedure of NiS/CdS/h-nanocomposites with schematic of the photocatalytic mechanism of H2 production on NiS/CdS/h-TiO2 photocatalysts [121].
Figure 12. Schematic illustration of the synthetic procedure of NiS/CdS/h-nanocomposites with schematic of the photocatalytic mechanism of H2 production on NiS/CdS/h-TiO2 photocatalysts [121].
Catalysts 12 01152 g012
Figure 13. SEM micrographs of some hierarchical photocatalysts obtained without templates [149,151,158,160].
Figure 13. SEM micrographs of some hierarchical photocatalysts obtained without templates [149,151,158,160].
Catalysts 12 01152 g013
Figure 14. Schematization of one-pot solvothermal synthesis of TiO2-graphene-MoS2 composite [165].
Figure 14. Schematization of one-pot solvothermal synthesis of TiO2-graphene-MoS2 composite [165].
Catalysts 12 01152 g014
Figure 15. Schematic representation of ZnO/CeO2 composites and charge transfer mechanism according to Z-scheme heterojunction [162].
Figure 15. Schematic representation of ZnO/CeO2 composites and charge transfer mechanism according to Z-scheme heterojunction [162].
Catalysts 12 01152 g015
Figure 16. Synthesis scheme (a) and Z-scheme heterojunction (b) for g-C3N4/ZnO hierarchical microspheres [171].
Figure 16. Synthesis scheme (a) and Z-scheme heterojunction (b) for g-C3N4/ZnO hierarchical microspheres [171].
Catalysts 12 01152 g016
Figure 17. Formation mechanism and TEM images of urchin-shaped iron oxide nanostructures (Fe2O3) produced through ultrasound irradiation [160].
Figure 17. Formation mechanism and TEM images of urchin-shaped iron oxide nanostructures (Fe2O3) produced through ultrasound irradiation [160].
Catalysts 12 01152 g017
Table 1. Hierarchical catalysts produced by using inorganic template(s).
Table 1. Hierarchical catalysts produced by using inorganic template(s).
CompositionSynthesis StrategyMorphologySSA (m2/g)ApplicationActivityIrradiationTemplate(s)Reference
SrTiO3/TiO2@carbon hydrothermalheterostructure with tri-modal (micro-, meso-, macro-) pores93–417photocatalytic hydrogen production from water splitting, methylene blue degradation2.52 mmol h−1 g−1
100% Degradation
UVKOH[93]
PbTiO3/TiO2/carbondual templatequasi-1D nanoneedle277–374photocatalytic and photoelectrochemical performancesMB degradation r 100% Under UV, 75% Under Visible+ ultrasound radiation the rates of hydrogen generation are 2360 and 9.6 μmol−1 g−1 UV-Visice/silica hard templates[94]
N–TiO2/Cdroppingflower-like in hierarchical porous structure217–407MB degradation and hydrogen production2.0 and 48 times of benchmark P25 for MB degradation: 95% under UV and 92% under Vis irradiation and the hydrogen production rates are as high as 2.832 and 0.038 mmol g−1 h−1 UV-Visice/silica hard templates[95]
Metal oxide/CDual-template method followed by heat treatmenthollow “dragon-bone” structure 14–375methylene blue degradation100% (UV)
90% (Vis)
UV-Visice/silica hard templates[96]
TiO2hydrothermal3D hierarchical porous core-shell13.5–42.3Adsorption and mineralization of tetracycline hydrochlorideup to 70% UV-Vissilica[97]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Venezia, V.; Pota, G.; Silvestri, B.; Costantini, A.; Vitiello, G.; Luciani, G. Tailoring Structure: Current Design Strategies and Emerging Trends to Hierarchical Catalysts. Catalysts 2022, 12, 1152. https://doi.org/10.3390/catal12101152

AMA Style

Venezia V, Pota G, Silvestri B, Costantini A, Vitiello G, Luciani G. Tailoring Structure: Current Design Strategies and Emerging Trends to Hierarchical Catalysts. Catalysts. 2022; 12(10):1152. https://doi.org/10.3390/catal12101152

Chicago/Turabian Style

Venezia, Virginia, Giulio Pota, Brigida Silvestri, Aniello Costantini, Giuseppe Vitiello, and Giuseppina Luciani. 2022. "Tailoring Structure: Current Design Strategies and Emerging Trends to Hierarchical Catalysts" Catalysts 12, no. 10: 1152. https://doi.org/10.3390/catal12101152

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop