Next Article in Journal
Hydroxypropyl Cellulose Hydrogel Containing Origanum vulgare ssp. hirtum Essential-Oil-Loaded Polymeric Micelles for Enhanced Treatment of Melanoma
Next Article in Special Issue
Improved Prediction of Elastic Modulus for Carbon-Based Aerogels Using Power-Scaling Model
Previous Article in Journal
Composite Hydrogel of Polyacrylamide/Starch/Gelatin as a Novel Amoxicillin Delivery System
Previous Article in Special Issue
Evolution Process of Fault Silica Aerogel under High Temperatures: A Molecular Dynamics Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress of Three-Dimensional Graphene-Based Composites for Photocatalysis

1
School of Materials Science and Engineering, Shenyang Ligong University, Shenyang 110159, China
2
School of Metallurgy, Northeastern University, Shenyang 110819, China
3
School of Equipment Engineering, Shenyang Ligong University, Shenyang 110159, China
*
Authors to whom correspondence should be addressed.
Gels 2024, 10(10), 626; https://doi.org/10.3390/gels10100626
Submission received: 22 August 2024 / Revised: 13 September 2024 / Accepted: 27 September 2024 / Published: 29 September 2024
(This article belongs to the Special Issue Recent Advances in Aerogels and Aerogel Composites)

Abstract

:
Converting solar energy into fuels/chemicals through photochemical approaches holds significant promise for addressing global energy demands. Currently, semiconductor photocatalysis combined with redox techniques has been intensively researched in pollutant degradation and secondary energy generation owing to its dual advantages of oxidizability and reducibility; however, challenges remain, particularly with improving conversion efficiency. Since graphene’s initial introduction in 2004, three-dimensional (3D) graphene-based photocatalysts have garnered considerable attention due to their exceptional properties, such as their large specific surface area, abundant pore structure, diverse surface chemistry, adjustable band gap, and high electrical conductivity. Herein, this review provides an in-depth analysis of the commonly used photocatalysts based on 3D graphene, outlining their construction strategies and recent applications in photocatalytic degradation of organic pollutants, H2 evolution, and CO2 reduction. Additionally, the paper explores the multifaceted roles that 3D graphene plays in enhancing photocatalytic performance. By offering a comprehensive overview, we hope to highlight the potential of 3D graphene as an environmentally beneficial material and to inspire the development of more efficient, versatile graphene-based aerogel photocatalysts for future applications.

1. Introduction

Environmental pollution and energy deficiency restrict the sustainable development of human society. Solar energy is an inexhaustible clean resource, and the photocatalytic technology based on solar energy has the unique advantages of low energy consumption, sustainability, and the complete mineralization of pollutants without causing secondary contamination. As such, it is considered one of the most viable approaches for both environmental remediation and chemical energy production [1,2,3,4]. Photocatalysts are the core of photocatalytic technology, and since Fujishima’s [5] groundbreaking discovery of water photoelectrolysis using a semiconductor TiO2 electrode in 1972, there has been a concerted effort to develop more efficient photocatalysts and innovative strategies to enhance their performance. These advances have significantly propelled the field of photocatalytic technology. To date, numerous semiconductor materials have been reported in research pertaining to photocatalysis, including metal oxides, metal sulfides, vanadates, carbonitrides, and metal-organic frameworks (MOFs), among others (see Table 1).
Photocatalytic reactions are complex physicochemical processes that typically unfold in several stages (Figure 1): (1) absorption of photons with energy greater than the semiconductor’s bandgap (Eg), leading to the generation of photoexcited electrons (e) and hole (h+) pairs; (2) charge separation, followed by migration and partial recombination in the semiconductor particles; (3) redox reactions occurring on the surface of the photocatalyst [6,7,8]. It is now widely recognized that the effectiveness of photocatalytic reactions relies not only on thermodynamic constraints but also on kinetic conditions. The intricate charge dynamics and surface reaction kinetics are considered to be the key factors in determining the quantum yield of these reactions [9]. Nevertheless, single-component photocatalysts often face limitations, such as restricted light absorption and high rates of electron-hole recombination, which severely hinder their potential for large-scale industrial applications. In response, various design strategies have been proposed to fabricate photocatalysts with excellent functionalities through morphological control, noble metal deposition, ion doping, semiconductor coupling, and surface sensitization [10,11,12,13,14,15,16]. Developing photocatalysts with optimized structural properties and a deeper understanding of the relationship between structure and activity is therefore essential to advancing their functionality and broader applicability.
Table 1. Classification of photocatalysts.
Table 1. Classification of photocatalysts.
ClassificationPhotocatalystSynthetic MethodMorphologyApplicationRef.
TiO2Anodic oxidationNanotube arraysDegradation RhB and BPA[17]
ZnOCalcination/SolvothermalNanorodsDegradation MB[18]
γ-Fe2O3Precipitation/CalcinationNanosheetsOxygen evolution[19]
Metal oxideCuOUltrasound/MicrowaveFlower-likeCr (VI) reduction[20]
WO3Precipitation/CalcinationNanoflakesDegradation CR and MB[21]
α-Bi2O3CalcinationNanoflowersDegradation CV[22]
CeO2HydrothermalNanorods/Nanowires Octahedrons/NanocubesNO oxidation and CO2 conversion[23]
CdSEtching/SulfurationDouble-shelled nanocagesCO2 reduction[24]
MoS2HydrothermalFlower-likeN2 fixation[25]
Metal sulfideFeS2HydrothermalNanorodsDegradation MB[26]
ZnSHydrothermal/CalcinationNanosheetsH2 evolution[27]
CuSSolvothermalMicrospheresDegradation RhB, MB, MO[28]
Bi2S3SolvothermalNanorodsN2 fixation[29]
BiVO4HydrothermalHollow nanotubesDegradation CR[30]
FeVO4Hydrothermal3D nanowall-likeCO2 reduction[31]
VanadatesAg3VO4HydrothermalMesoporousDesulfurization[32]
InVO4Microwave HydrothermalNanocrystals MicrospheresH2 evolution[33]
Carbonitrideg-C3N4Molten saltNanorodsH2O2 generation[34]
C3N5Molten saltNanosheetsH2 evolution[35]
Ti-MOFCondensationNeedle-likeCO2 reduction[36]
MOFsNi-MOFHydrothermalNanosheetsCO2 reduction[37]
Fe-MOFSolvothermalHexagonal bipyramidH2 evolution[38]
Co-MOFSolvothermalNanosheetsDegradation dyes[39]
Graphene is a two-dimensional (2D) planar carbon isomer with a honeycomb lattice structure formed by the sp2 hybridization of carbon atoms. Since its discovery by Geim in 2004, graphene has been recognized as a remarkable nanomaterial with exceptional optical transparency, electrical conductivity, mechanical strength, and chemical stability. Graphene provides a promising basis for photocatalytic applications due to its distinctive properties. There has been explosive interest in constructing various photocatalysts based on the versatile platform of graphene, ranging from simple semiconductor deposition to the precise control of multi-component growth, in the past few years [40,41,42,43,44,45,46]. Graphene-based nanocomposites with significantly enhanced photocatalytic performance have been widely employed in photochemical conversions, such as organic pollutants degradation, water splitting, CO2 reduction, N2 fixation, and photovoltaic application, etc. In photocatalytic reactions, the redox processes actually take place on the surface of the photocatalyst rather than in the liquid matrix. Effective photocatalysis depends not only on rapid charge separation and transfer but also on the ability of active sites on the photocatalyst surface to adsorb reactants [47,48]. Thus, designing photocatalysts with abundant active sites is critical for improving both the yield and selectivity of target products in target reactions. Graphene aerogel, as a derivative of graphene nanomaterials, is characterized by an orderly three-dimensional porous macroscopic structure combining mesopores and micropores, enabling the accommodation of plenty of nanoparticles on the framework or the skeleton due to its substantial surface area. Compared to its 2D counterpart, 3D graphene retains the beneficial properties of individual graphene sheets while avoiding the agglomeration caused by van der Waals interactions. Furthermore, its interconnected hierarchical porous structure provides easily accessible catalytic sites for reactive species, multidimensional electron transport pathways, and efficient channels for the mass transfer of electrolyte ions [49,50,51].
Over the past decade, some reviews have been conducted on 3D graphene-based materials, primarily focusing on the applications of graphene aerogels in energy storage (supercapacitors, batteries) [52,53,54,55], as well as their potential use as environmentally functional materials for pollutant removal [56,57,58]. However, comprehensive reviews specifically addressing the classification of 3D graphene-based photocatalysts and their diverse redox reactions remain relatively scarce. This review is dedicated to recent research breakthroughs in the synthesis and application of various types of 3D graphene-based photocatalysts and providing insights into the important role of 3D graphene in photocatalysis. We initially summarize the available synthetic methodologies for 3D graphene architectures. In the subsequent section, different types of 3D graphene-based photocatalysts are presented and their properties are highlighted. Subsequently, the application of these materials in photocatalysis is elaborated in detail, encompassing organic pollutant degradation, H2 evolution, and CO2 reduction. Finally, the challenge and prospective developments of 3D graphene in advanced multi-field photocatalytic applications are deliberated. It is expected that this paper may provide some support for the continuous advancement of the efficient design, manufacturing, and application of 3D graphene-based photocatalysts.

2. Basic Features

2.1. General Properties of Graphene and Its Derivatives

Among carbon nanomaterials, graphene has emerged as a revolutionary material in both scientific and engineering research. Composed of a monolayer of carbon atoms arranged in a honeycomb lattice, graphene exhibits remarkable electrical and optical properties, including high electron mobility (~200,000 cm2 V−1 s−1), ballistic electronic transport, and outstanding transparency with an optical transmittance of approximately 97.7%. Additionally, graphene has a theoretical specific surface area of 2630 m2 g−1, excellent thermal conductivity (~5000 W m−1 K−1), strong chemical stability, and high mechanical strength [59,60]. A schematic illustration of graphene and its derivatives can be seen in Figure 2 [61]. Reduced graphene oxide (rGO) and graphene oxide (GO) constitute two commonly recognized forms of graphene derivative. GO is an oxidized derivative of graphene that contains numerous functional groups on its basal planes and edges, allowing it to form stable suspensions in various solvents and functioning as nucleation sites for the growth of other nanoparticles. However, GO is inherently less conductive than graphene due to the structural disorder induced by sp3 C-O bonds. GO can be reduced through chemical/thermal reduction or other methods to form rGO, enhancing its chemical stability and elevating its electrical conductivity [62]. As a highly versatile 2D building block, graphene has been assembled into 0D fullerenes, 1D carbon nanotubes, and 3D aerogels, all of which possess the potential to substantively extend the field of graphene applications. Notably, 3D graphene structures not only retain the inherent properties of graphene but also acquire new physicochemical characteristics. The 3D architectures constituted from 2D graphene sheets (including GO and rGO) via van der Waals forces or functional groups are easier to construct compared to other carbon materials [53]. Given these unique properties, 3D graphene, whether as a supporting material or synergistic component, already meets many of the essential requirements for advanced photocatalysts.

2.2. Functionalities of 3D Graphene for Photocatalysis

Optical and electrical properties are key factors in the development of photocatalytic materials. As research on graphene-based photocatalysts advances, it is important to highlight the role of 3D graphene within these systems. 3D graphene can improve the photocatalytic activity of composite photocatalysts in multiple manners, as schematically illustrated in Figure 3. 3D Graphene as a Supporting Material; The large surface area and three-dimensional network of 3D graphene can support a substantial amount of semiconductor nanoparticles. Particularly, the negatively charged surface of reduced graphene oxide (rGO) facilitates the adsorption of nanoparticles, preventing agglomeration. This allows for controlled interfacial contact and the distribution of nanomaterials when precursors with varying compositions are added. Additionally, the diverse porous structure of 3D graphene (including micro-, meso-, and macropores) ensures better access for reactants and exposes more active sites, improving adsorption and surface photocatalytic reactions. 3D Graphene as a Photosensitizer; In photochemical reactions, photosensitizers are capable of transmitting light energy to semiconductors that are not inherently sensitive to visible light, thereby enhancing or amplifying their photosensitive properties [63]. Graphene can function as both a metal with a vanishing Fermi surface and a semiconductor with a vanishing bandgap that stems from graphene’s honeycomb lattice composed of two equivalent carbon sublattices. This peculiarity can be regulated through the incorporation of heteroatoms or oxygen functionalities. The introduction of oxygen functionalities leads to the formation of C-O covalent bonds, disrupting the symmetry of the carbon sublattices and thereby altering their electronic properties. The bandgap is contingent upon the extent of graphene oxidation [64,65]. In regard to this, 3D graphene acquired through chemical redox methods can act as a photosensitizer, generating electrons, enhancing its light harvesting ability via multi-reflections within its structure, and modifying the bandgap of the photoactive components to expand the range of photoabsorption [66]. 3D Graphene as a Photoelectron Acceptor; A major limitation of conventional photocatalytic materials is the rapid recombination of photogenerated electrons and holes, reducing the formation of energetic charge carriers. The efficient separation of these charges is critical for improving photocatalytic performance. Graphene’s electrical properties can be tuned by varying its oxidation level (C/O ratio), transforming it from a conductor to an insulator [67]. When electrically conductive 3D graphene is incorporated into photocatalysts, the formed three-dimensional conductive network offers rapid transport channels for photoelectrons, facilitating charge separation and promoting the overall photocatalytic efficiency. These multifunctional roles make 3D graphene a highly effective component in the design of advanced photocatalysts.

3. Synthetic Strategies of 3D Graphene Architectures

Among the myriad 3D nanostructured materials, 3D graphene has garnered considerable attention owing to its diverse structural properties, encompassing a large surface area, low density, high porosity, and compressibility. Furthermore, 3D graphene is derived from 2D graphene by cross-linking together to form a network structure while retaining the exceptional electrical conductivity and physicochemical stability of graphene. These characteristics make it an efficacious scaffold for accommodating active materials across a wide range of applications in energy storage and conversion [68,69,70], photocatalysis [71,72,73], oil/water separation [74,75,76], electromagnetic shielding [77,78,79,80], sensors [81,82,83], and other fields. The choice of synthesis method plays a crucial role in customizing the morphology, size, defect structure, and surface/interface properties of 3D graphene materials. Numerous methods have been reported for fabricating 3D graphene and graphene-based structures, such as spheres [84], foam [85], aerogel [86], film [87], and hollow fiber [88], as shown in Figure 4. This review is dedicated to exploring the utilization of 3D structured graphene in photocatalysis; hence, the controllable preparation of well-ordered graphene macroscopic composites is a pivotal step. Additionally, attention should be paid to parameters such as size, specific surface area, porosity, bulk density, and surface wettability, which are affected by the preparation method. This section provides a brief overview of the main synthesis strategies for constructing 3D graphene architectures, which are typically divided into two categories: template and non-template methods.

3.1. Template Method

Template carbonization pioneered the utilization of SiO2 and NaCl as templates to precisely control both the morphology and pore structure of carbon materials during the 1980s and further extended this methodology to the construction of 1D, 2D, and 3D nanostructured carbon materials [89]. Diversifying template options significantly widened the synthetic possibilities for creating well-ordered 3D graphene materials. Essentially, alternative synthesis methods are based either on hard templates or soft templates. Among the hard templates, commonly used materials include metal foams (e.g., Ni, Cu), solid nanostructured particles (e.g., polystyrene spheres, silica, zeolites), and solidified solvents such as ice. These hard templates are instrumental in fabricating 3D graphene-based composites with varying macro- and microscopic structures. During the synthesis process, van der Waals forces along with ionic bonds play a pivotal role in assembling graphene components around the template. Once the template is removed, typically through etching or dissolution, a porous 3D graphene structure is left behind.

3.1.1. Hard Template Method

Template-Directed CVD Method

Template-assisted chemical vapor deposition (CVD) is a highly effective method for the direct production of large-area and high-quality 3D graphene films by depositing carbon onto metal templates. The predominant architecture of the graphene is determined by the selection of different template forms and growth conditions, such as the flux and concentration of carbon sources. Graphene prepared using the CVD method closely resembles the physical structure of the original graphene sheet. Chen et al. [90] initially devised a methodology for the directional synthesis of 3D graphene through CVD by utilizing a nickel foam template. Graphene layers were formed on the surface of the template by the decomposition of CH4 at a temperature of 1000 °C. To maintain the integrity of the graphene network during the chemical etching process of the nickel template, a thin layer of polymethyl methacrylate was applied to support the graphene surface and subsequently removed by hot acetone. The resulting graphene retained the interconnected 3D scaffold structure of the nickel foam. Xia et al. [91] discovered that the stability of graphene samples is significantly affected by the removal of metal templates through solvent etching due to the introduction of defects and functional groups. To address this issue, they pioneered a solvent-free process using microporous copper as the template for creating a 3D graphene network. Following the growth of graphene via the CVD process using bubbling liquid ethanol as the carbon source, the copper template was evaporated at 1300 °C in a vacuum to obtain pure freestanding 3D graphene (Figure 5). Additionally, incorporating other precursors can lead to the formation of 3D graphene-based composites or doped structures. Jia et al. [92] fabricated 3D interconnected graphene foam/epoxy composites through a three-step process: (i) carbonization on a Ni foam template via CVD; (ii) impregnation of epoxy resin into the graphene-Ni foam followed by curing treatment; (iii) etching of the Ni template to create cellular-structured graphene/epoxy composites.

Microsphere Template-Assisted Method

3D interconnected graphene frameworks created through the template method feature an organized microstructure, with their hierarchically porous architecture adjustable by modifying the size and concentration of template particles. Spherical polystyrene (PS) and silica particles are commonly used to develop high-surface-area graphene materials with micro-, meso-, and microporosity. Choi et al. [93] synthesized 3D graphene films with microporous structures using 2 μm PS spheres as templates. They first chemically reduced a graphene oxide suspension with ammonia and hydrazine to create a negatively charged graphene solution, which facilitated the uniform distribution of positively charged PS particles through electrostatic interactions. The mixed solution was subsequently filtered through an anodic membrane and peeled off to obtain composite films, and the final free-standing 3D graphene was achieved by removing the PS particles with toluene. Additionally, the prepared 3D graphene served as a scaffold for depositing a thin layer of MnO2 onto macropores via a self-limiting reaction to construct MnO2/graphene films. Xu et al. [94] utilized a PS microsphere monolayer colloid crystal as a template to fabricate ordered macroporous graphene-based films with tightly arranged pores on a curved surface (Figure 6). This fabrication process involved three steps: ① immersing the monolayer PS colloidal crystal, prepared by spin coating, in a graphene oxide solution; ② retrieving the colloidal crystals containing graphene oxide with a ceramic tube; and ③ heating the ceramic tube to 340 °C using a Ni/Cd alloy resistance wire to burn off the PS microspheres and reduce graphene oxide, resulting in macroporous graphene films. For further fabrication of graphene-incorporating (SnO2, Fe2O3, and NiO) composite film, the synthetic procedure remained consistent with that described above, except for directly dissolving specific materials into the graphene oxide solution. Salazar Aguilar et al. [95] employed a layer-by-layer method to construct 3D graphene by integrating spherical silica particles as templates. They first functionalized silica dispersions with a positively charged polyelectrolyte PEI to enhance interaction with the graphene oxide surface. The PEI-modified silica suspension was then added dropwise to the graphene oxide solution. After centrifugal washing of the silica particles covered with graphene oxide, the process was repeated with additional PEI modification and graphene oxide coating. Finally, thermal treatment and HF etching were employed for graphene oxide reduction and template removal to achieve the 3D graphene architecture.

Ice Template-Assisted Method

The ice template method is widely utilized for creating oriented microstructure materials with homogeneity due to its simplicity and ease of operation. This method relies on the segregation of the solidifying solvent template from the target phase through ice sublimation under extremely low temperature and vacuum conditions, resulting in porous materials with a templated pore structure. Common solvents used in ice templates include water, camphene, and tert-butanol, with water being the most frequently used dispersing medium due to its abundance, low cost, and non-toxic nature [96]. Zong et al. [97] introduced a bidirectional freeze-casting method to fabricate graphene aerogels with three distinct microstructures by employing a dual temperature gradient approach (Figure 7). They began with a graphene oxide aqueous solution prepared using the Hummers method, which was mixed with ascorbic acid and heated at 70 °C for varying durations to produce graphene hydrogels. Subsequently, these hydrogels were respectively immersed in −35 °C ethanol for 5 h, followed by thawing at room temperature. The final graphene aerogels were obtained after lyophilization and hydrazine reduction. The study demonstrated that the oriented microstructure of the graphene aerogels could be precisely controlled by adjusting the hydrothermal duration and the dual temperature gradient, which influenced ice crystal growth. Qiu et al. [98] fabricated 3D graphene nanoplates with vertical channels extending along the radius, driven by the growth of crystals in a radial pattern (Figure 8). They dispersed commercial graphene nanoplates into a solution of itaconic acid and chitosan using sonication. The mixture was then rapidly cryogenically frozen using liquid nitrogen at −60 °C for 30 min to promote the orientation of ice crystals. Subsequently, the frozen sample underwent freeze-drying to sublimate water and obtain the graphene/chitosan composite aerogel. Kota et al. [99] fabricated 3D nitrogen-doped graphene monoliths with a mesoporous structure using ice-templated assembly, incorporating melamine as the nitrogen source. After homogeneously blending the graphene oxide solution with melamine, the mixture was subjected to cryogenic freezing using liquid nitrogen, followed by thawing. The graphene oxide was then significantly reduced by annealing at 900 °C for 30 min, leading to the formation of a more robust porous structure.

3.1.2. Soft Template Method

Soft templates, typically organic molecules that self-assemble into micelles or other nanoscale structures in a liquid phase, provide a simpler and milder alternative to hard templates for controlling the shape and growth of rigid particles [100]. Huang et al. [101] initially proposed a strategy for fabricating 3D graphene foams with tunable pore structures using microemulsions and micelles as soft templates (Figure 9A). The process began with the introduction of TMB (or n-hexadecane) into 2M HCl and sonication to create a cloudy suspension. This suspension was then combined with an aqueous graphene oxide solution, allowing the graphene oxide sheets to spontaneously arrange around the templates due to hydrophobic interactions. The resulting composites were separated via vacuum filtration and subjected to dual-step calcination under an inert atmosphere to obtain the final graphene material. This approach enabled the production of graphene with adjustable pore diameters ranging from a few micrometers to tens of micrometers, as illustrated in Figure 9B,C. Li et al. [102] employed a soft template-assisted approach combined with heat-pyrolysis to synthesize N and S co-doped 3D multi-porous graphene. Their method involved using melamine as a crosslinking agent and nitrogen source, and benzyl disulfide as the sulfur source. Initially, graphene oxide dispersion was mixed with melamine and formaldehyde aqueous solution to form a homogeneous mixture. Subsequently, the mixture was subjected to hydrothermal treatment at a temperature of 180 °C for 12 h to form a hybrid hydrogel, which was dried overnight at 80 °C to produce xerogel. Finally, the resulting xerogel underwent pyrolysis in an argon atmosphere to form N and S co-doped graphene.

3.2. Non-Template Method

3.2.1. Self-Assembly Method

Self-assembly has been recognized as a highly effective strategy in the realm of ‘bottom-up’ nanotechnology. It is also a prevalent method for fabricating graphene and graphene-based composites featuring three-dimensional structures, capable of establishing intermolecular cross-links through hydrogen bonding, π-π stacking interactions, electrostatic forces, etc. [103]. This method offers advantages in terms of suitability for large-scale production due to its simple process. This section discusses the synthetic strategies for creating 3D graphene macrostructures from graphene oxide through self-assembly under hydro/solvothermal synchronous thermal reduction and chemical reduction processes.

Hydrothermal/Solvothermal Reduction

The hydrothermal method has become a leading approach for synthesizing interconnected 3D porous graphene hydrogels, notably advanced by Shi et al. [104]. Their work demonstrated that treating a homogeneous graphene oxide aqueous solution at 180 °C for 12 h yields a three-dimensional graphene hydrogel with a porous network. Graphene oxide’s hydrophilicity and electrostatic repulsion effect enable stable dispersion in water. Hydrothermal reduction then removes most oxygen-containing functional groups, enhancing the material’s hydrophobicity. This leads to the random stacking of flexible graphene sheets through π-π interactions, forming a porous framework with pore sizes ranging from sub-micrometer to several micrometers (Figure 10). Building on this foundation, they further prepared the 3D macrostructure graphene organogel using solvothermal reduction of graphene oxide dispersed in propylene carbonate [105]. Both hydrothermal and solvothermal reduction are the most direct methods for preparing graphene gels without requiring additional purification treatment. Moreover, these processes are compatible with synthesizing many functional materials, facilitating the incorporation of multiple active components into graphene frameworks, greatly expanding the diversity and functionality of the self-assembled graphene macrostructures. Numerous researchers have reported the construction of three-dimensional graphene/metal composites [106,107,108], graphene/metallic compounds [109,110,111,112], and graphene/polymers [113,114,115] hybrids through hydrothermal or solvothermal methodologies.

Chemical Reduction

In contrast to the high-temperature and high-pressure conditions required by the hydro/solvothermal method, the chemical reduction approach necessitates slightly milder experimental parameters. It is one of the conventional and economical methods for preparing graphene in large quantities through the reduction of graphene oxide. For chemical reduction, by altering the properties and mass of reducing agents, the surface chemistry can be readily adjusted to achieve the self-assembly of 3D graphene. Wang et al. [116] developed a method for the fabrication of macroscopic porous graphene aerogels using a mild in situ self-assembly process combined with thermal annealing. Sodium bisulfite (NaHSO3) was employed as the reducing agent in this approach. The chemical reduction took place at 80 °C under atmospheric pressure, which enhanced the surface hydrophobicity of graphene oxide. This change in hydrophobicity strengthened the π-π interactions between adjacent graphene sheets, promoting self-assembly into a 3D network. By varying the mass ratio of graphene oxide to NaHSO3 during similar processes and conditions, the self-assembly behavior and the resulting graphene aerogel’s density can be regulated. Higher reducing agent concentrations led to lower-density aerogels due to excessive accumulation of reduced graphene sheets. Chen et al. [117] carried out research on the effects of various reducing agents including NaHSO3, Na2S, Vitamin C, HI, and hydroquinone on the reduction of graphene oxide. Their study highlighted that different reducing agents influenced the reduction times required for forming 3D structures. Additionally, altering the reactor type also enabled precise manipulation of the shape of the hydrogel, revealing the isotropic contraction of self-assembled graphene oxide sheets. The flexibility of chemical reduction in manipulating reducing agents and reaction conditions offers a valuable strategy for tailoring the properties and structures of 3D graphene materials, rendering it a versatile approach for large-scale and economical graphene production.

3.2.2. 3D Printing Method

3D printing is an additive manufacturing technique that builds objects layer by layer based on digital models [118]. This method is particularly promising when combined with hybrid graphene materials, as it allows for the rapid creation of complex 3D structures with specific performance and functional requirements. Currently, several 3D printing methods are used to manufacture graphene-reinforced composites, including inkjet printing [119,120], direct ink writing [121,122,123,124,125], fused deposition modeling [126,127,128], and the light-curing molding method [129,130,131]. Of these techniques, the direct ink writing method is the most versatile strategy for assembling satisfying three-dimensional hybrids due to its flexibility in material selection and facile printing process. Compared to conventional bulk graphene aerogels, the graphene materials obtained through 3D printing showcase unique micro- and macroscale architectures, resulting in exceptional mechanical properties (stiffness and stretchability) at a low density. For example, Peng et al. [132] employed an inkjet 3D printing methodology to fabricate ultralight graphene structures. This 3D printing process was to continuously squeeze the highly viscous paste onto a glass wafer at room temperature in the air by precisely controlling nozzle movement. They optimized partially reduced graphene oxide ink and combined it with freeze-casting and thremal reduction processes to achieve a hierarchical structure with both a macroscopically hollow scaffold and microscopically cellular features (Figure 11). Hu et al. [133] functionalized graphene oxide with amphiphilic polyethylene glycol to produce inks with stable dispersion properties in various organic and aqueous solutions for printing self-supporting 3D structures. Tang et al. [134] developed a versatile 3D printing approach for the manufacturing of graphene-based aerogels with complex architectures. In their printing process, urea was employed as a precursor to crosslink graphene oxide sheets and incorporate negatively charged materials (0D Ag nanoparticles, 1D multiwalled carbon nanotubes, and 2D MoS2 nanosheets) into the graphene oxide matrix. This approach allowed the creation of homogeneous ink systems and produced hybrid aerogels with intricate internal structures through 3D printing, followed by freeze-drying and chemical reduction.

4. Synthesis of 3D Graphene-Based Composite Photocatalysts

The conversion of solar energy into chemicals within photocatalytic systems relies on crucial processes: light absorption, charge separation, and surface reactions. Three-dimensional structured graphenes have emerged as highly effective materials for creating efficient photocatalysts due to their unique macrostructure and porous properties. 3D graphene enhances photocatalytic activity in multiple ways. Its outstanding electrical conductivity and extensive electron transport pathways make it an excellent mediator and acceptor for photoelectrons, thus promoting the effective separation of electron-hole pairs generated by light. Additionally, the large surface area and complex porous structures of 3D graphene improve adsorption capacity and prevent the aggregation of semiconductor particles. This structure not only provides more active sites for surface catalytic reactions but also supports better performance. Moreover, 3D graphene can act as a photosensitizer to intensify light harvesting and extend the light adsorption range of composite photocatalysts [135,136]. In this section, we review the synthesis strategies of various 3D graphene-based photocatalysts and explore how to fully use the diversiform roles of graphene in photocatalytic systems by adjusting active components and optimizing structure.

4.1. 3D Graphene/Metal Oxides

Metal oxides are currently the most widely employed photocatalysts due to their cost-effectiveness, chemical stability, as well as low toxicity. Of these, titanium (TiO2) stands as the most extensively studied, abundantly available, and eco-friendly n-type semiconductor. However, its relatively large band gap of about 3.2 eV restricts its light absorption to the ultraviolet region, and the rapid recombination of photogenerated charge carriers results in low quantum yield. Recently, numerous studies have been conducted to enhance photocatalytic activity through the integration of nano TiO2 with 3D graphene. Liu et al. [137] synthesized TiO2/graphene aerogels with uniformly distributed TiO2 nanorods using a hydrothermal and freeze-drying method for the photocatalytic reduction. The total yield of carbon generated by TiO2/graphene was 15.7 times higher than that of pure P25, which is attributed to the expansion of the light response range and the enhancement of interface charge transfer to mitigate electron-hole pair recombination. Huang et al. [138] employed chemical vapor deposition to synthesize high-quality graphene, which was then combined with TiO2 nanoparticles through the sol-gel method for photocatalytic hydrogen evolution. Benefiting from the low defect density and excellent electrical conductivity of few-layer graphene, the separation and lifetime of photoexcited charge carriers in the graphene/TiO2 photocatalyst were enhanced, resulting in high sustained stability with a hydrogen evolution rate 26.2 times higher than that of blank TiO2. Nawaz et al. [139] employed an in situ hydrothermal method to create TiO2/graphene aerogels for the photodegradation of recalcitrant carbamazepine. The composites exhibited enhanced adsorption and nearly doubled photodegradation capability compared to bare TiO2, attributed to the macroporous structure, effective charge separation, and efficient mass transportation of carbamazepine to the photocatalyst surface. Additionally, 3D graphene/TiO2 heterojunction photocatalysts have garnered significant attention for their ability to prolong charge carrier recombination times through built-in electric fields. Liu et al. [140] prepared TiO2/C/BiOBr graphene aerogel with an indirect Z-scheme heterojunction through the combination of electrostatic spinning and solvothermal methods. Many BiOBr nanosheets were grown on the surface of TiO2/C nanofibers to form a 3D hierarchical nanostructure in conjunction with graphene. The results of photocatalytic experiments revealed that the composite photocatalyst achieved a degradation rate of 97.5% for RhB, approximately 8.7 times higher than that of TiO2/C. This enhancement is attributed to the synergistic effect of the heterojunction formed between TiO2 and BiOBr as well as the improved conductivity and physisorption resulting from graphene. By optimizing the relative proportions of each constituent and the configuration of photocatalysts, Jung et al. [141] developed a composite material featuring mesoporous TiO2 and layered MoS2 cocatalyst on 3D graphene for CO2 photoreduction. The electron transfer from TiO2 through graphene to MoS2 effectively reduced the charge recombination rate, resulting in a higher CO photoconversion rate of 97% and a production yield of 92.33 μmol/g·h, outperforming other combinations such as bare TiO2, TiO2/graphene, and TiO2-MoS2 under simulated sunlight irradiation.
Many other metal oxides incorporating 3D graphene materials have also been prepared and applied in photocatalysis. Zinc oxide (ZnO) is an n-type wide gap semiconductor similar to TiO2, with high photosensitivity, relatively low cost, and environmental compatibility. Thus, ZnO has the same limitation as TiO2 in terms of rapid electron/hole recombination. Men et al. [142] created a 3D self-supporting graphene structure using Ni foam as a template, and subsequently grew ZnO nanorods on the graphene surface through a hydrothermal process, as illustrated in Figure 12. The ZnO/graphene composite foam features a micro-nano hierarchical structure; the microporous graphene framework not only enhances the light harvesting but also functions as an electron storage reservoir to facilitate the separation of electrons and holes, leading to the higher photocurrent and photocatalytic degradation activity of RhB compared to pure ZnO. Iron oxide (α-Fe2O3) is an earth-abundant, environmentally friendly n-type semiconductor with a relatively narrow band gap (∼2.1 eV), which makes it responsive to visible light (absorbance edge ∼600 nm). Despite these advantages, the utilization of α-Fe2O3 has been constrained by the high electron/hole recombination effect and low diffusion length [143]. Wang et al. [144] utilized a microgel template-assisted solvothermal method to synthesize uniform hollow α-Fe2O3 microspheres, and subsequently employed direct writing 3D printing to fabricate α-Fe2O3/graphene aerogel microreactors. The resulting macroscopic porous and highly interconnected networks exhibit a high specific surface area and structural stability, thereby enhancing multi-dimensional mass transfer channels and improving reusability. Additionally, the conducting graphene adjacent to the α-Fe2O3 facilitates rapid electron transport, leading to the efficient formation of photoinduced holes in the α-Fe2O3 microspheres. Under 120 min of solar irradiation, the degradation rate of RhB reached 97.8%, while maintaining 96.2% catalytic stability after four cycles. Tungsten trioxide (WO3) is another n-type semiconductor known for its non-toxicity, photochemical stability, and visible light absorption (up to ~480 nm) with a bandgap of 2.6–3.0 eV. Pure WO3 often suffers from low charge carrier mobility and rapid recombination of photogenerated carriers [145,146,147]. Enhancements in the performance of WO3 can be achieved through structural and morphological adjustments as well as the incorporation of high-quality conductive graphene. Azimirad et al. [148] prepared graphene foam through chemical vapor deposition and utilized it as a conductive substrate for the nucleation and growth of WO3 nanoparticles. The visible light photocatalytic activity of the 3D WO3/graphene composite in RhB degradation was found to be strengthened compared with pure WO3 due to the formation of W-C and W-O-C bonds between graphene and WO3, which facilitates faster photoexcited electron transfer to graphene. Li et al. [149] used a one-step hydrothermal method to synthesize 3D graphene-wrapped WO3 microspheres assembled with nanowires. The hierarchical structure of WO3 and the intimate wrapping of graphene on the nanowire surface synergistically improved the photocatalytic activity of WO3/graphene in degrading phenol and RhB. Additionally, the content of graphene played a crucial role in improving photocatalytic performance, with optimal performance reaching about 2.5 times that of bare WO3. Cuprous oxide (Cu2O) is a typical p-type semiconductor characterized by its low cost, non-toxicity, and bandgap of approximately 2.0 eV, enabling a specific response to visible light. However, its broad applicability is constrained by the rapid recombination of photogenerated carriers. Additionally, the stability of Cu2O poses a significant issue in redox reactions. The efficient coalescence between Cu2O and graphene not only enhances electron transport but also modulates the stability of the anchored Cu2O crystals [150,151,152,153].

4.2. 3D Graphene/Metal Sulfides

Compared to metal oxides, metal sulfides offer several advantages for photocatalysis. The valence band of metal sulfides, which is occupied by S 3p orbitals, is more negative than that of metal oxides, where the valence band is occupied by O 2p orbitals. This difference allows metal sulfides to harvest a broader range of light and achieve higher carrier concentrations. The versatility in selecting and adjusting various metal elements further enhances the electronic structure and surface physicochemical properties of metal sulfides, making them highly tunable for photocatalytic applications [154,155]. Thus, MoS2, CdS, WS2, CuS, and other binary metal sulfides with narrow bandgaps are widely recognized as efficacious visible light photocatalysts. Nevertheless, the photocatalytic efficiency of metal sulfides remains unsatisfactory primarily due to the sluggish kinetics of carrier separation and migration. And the relatively low chemical stability caused by the photo corrosion of surface sulfide ions severely impedes highly efficient solar energy conversion. Recent studies have shown that combining metal sulfides with graphene can significantly improve their optoelectronic properties [156]. For example, Das et al. [157] synthesized hierarchically porous and macroscale MoS2/graphene photocatalysts using a one-pot hydrothermal method. This approach assembled rosette-like MoS2 nanoflowers into the interconnected networks of graphene aerogels, resulting in composites with a narrow bandgap (~1.3 eV) and enhanced photon absorption across the visible light spectrum up to 700 nm. The extensive planar interface between MoS2 and graphene facilitates efficient spatial separation of photoinduced charge carriers, while the micro/mesoporosity of the structure improves interactions with reactants. The optimal MoS2/graphene aerogel demonstrated a photocatalytic degradation efficiency of 91% for tetracycline, about 1.7 times higher than that of pure MoS2, and retained its crystallinity and integrity after three repeated cycles. Wei et al. [158] used an in situ hydrothermal and freeze-drying method to assemble spherical CdS nanoparticles (~10 nm) on graphene aerogel. The resultant hybrids feature a hierarchical porous structure and a robust electronic interaction between CdS and graphene, thereby leading to the significantly enhanced adsorption capacity of reactants and efficient transfer of photo-generated carriers. This composite exhibited significantly improved photocatalytic degradation of organic contaminants, with efficiencies 15.6, 6.6, 4.4, 2.8, and 2.2 times higher for MO, MB, CIP, RhB, and AcbK, respectively, compared to pure CdS. Bano et al. [159] synthesized a CuS/graphene hybrid aerogel through a chemical reduction process to achieve the synergistic photocatalytic reduction of Cr (VI) and degradation of cationic dyes (MB and RhB). In comparison with powdered CuS nanospheres, this 3D porous framework facilitated better spatial separation and transport of photoinduced charge carriers, resulting in enhanced visible light catalytic activity and improved recyclability.
By leveraging the exceptional electron mobility of graphene as an electron trap to stimulate charge separation in semiconductors, the photocatalytic activity can be further augmented through doping with heteroatoms (such as N) to customize the electronic and structural properties of graphene [160]. For example, Shafi et al. [161] proposed a novel photocatalyst by in situ growing few-layer WS2 nanosheets on N-doped graphene aerogel through hydrothermal processing. The coexistence of pyridinic N species and WS2 clusters synergistically provides abundant active sites for catalytic reactions, while the interconnected conductive networks effectively regulate charge separation and migration. This combination allowed the WS2/graphene composite to achieve a degradation of up to 93% of the psychoactive drug caffeine within 180 min under visible light irradiation. Zhang et al. [162] employed a bottom-up approach to fabricate CoS2/MoS2/N-graphene aerogel photocatalysts. The heterostructure formed by CoS2 nanoparticles and ultra-thin MoS2 nanosheets connected via S-atoms chemical bonding can effectively prevent the aggregation of nanoscale MoS2, thereby exposing more edge active sites and accelerating charge separation at the heterojunction interface. Moreover, N-doped graphene served as the substrate for CoS2/MoS2 heterostructures to enhance adsorption, facilitate the migration of photogenerated electrons, and promote photocatalyst recovery. As indicated by the results, the composite exhibited a photocatalytic efficiency of 97.1% for RhB degradation, surpassing pure MoS2 (51.7%) by a significant margin.

4.3. Other 3D Graphene-Based Photocatalysts

In addition to metal oxides and metal sulfides semiconductor materials, graphitic carbon nitride, bismuth-based materials, and metal-organic frameworks are also considered as promising photocatalysts. Non-metallic g-C3N4 is a polymer semiconductor with a bandgap of 2.7 eV, structured as a 2D graphitic lattice formed by tri-s-triazine units linked with amino groups. It possesses several beneficial properties, including chemical stability, low cost, environmental friendliness, and facile preparation through the thermal polycondensation of diverse nitrogen-containing precursors [163]. However, g-C3N4 poses great challenges for practical applications due to the high recombination rate of photogenerated carriers, low electronic conductivity, and limited absorbance of visible light below 460 nm [164]. To address these issues, constructing heterojunctions between g-C3N4 and graphene can enhance photocatalytic performance by improving charge carrier separation and extending light absorption. To optimize its charge carrier separation and broaden the light absorption region, the construction of heterojunctions between g-C3N4 and graphene is anticipated to deliver excellent photocatalytic properties. Zhang et al. [165] prepared a porous g-C3N4/graphene aerogel photocatalyst using a hydrothermal co-assembly method. The formation of heterojunctions effectively mitigated electron-hole recombination and intensified visible light utilization through multireflection across the 3D interconnected porous frameworks. Furthermore, the large planar interface between g-C3N4 and graphene sheets increased the active sites, resulting in an 83.0% purification efficiency for dye MB within 3 h under visible light. To further suppress the high carrier recombination rate of g-C3N4, Zhang et al. [166] developed an efficient aerogel photocatalyst embedding palladium (Pd) within g-C3N4/graphene for the conversion of CO2 to CH4. In this photocatalytic system, Pd nanoparticles acted as electron traps to enhance electron-hole separation, while graphene provided a macroscopic substrate that facilitated contact CO2 and light energy utilization. The integration of Pd with g-C3N4 and graphene established a 2D-2D electron pathway, resulting in a maximum CH4 evolution rate of 6.4 μmol/g·h, a 12.8-fold increase compared to pure g-C3N4.
Bismuth-based semiconductors (such as Bi2O3, Bi2S3, BiVO4, BiOBr, Bi2WO6, etc.) are promising candidates for photocatalysis due to their unique electronic band structures, adjustable and expandable spectral response ranges from 400 nm to 700 nm, and low toxicity [167]. Despite these advantages, they suffer from rapid recombination of photogenerated carriers and high susceptibility to photo corrosion. Crystal facet engineering, ion doping, and coupling with noble metals or graphene materials are effective strategies to improve the characteristics of Bi-based photocatalysts [168]. Yu et al. [169] prepared a series of BiOBr/graphene aerogels using a two-step hydrothermal method, where flower-like BiOBr self-assembled on the graphene surface to form heterostructures with dopamine as a cross-linker. An increase in redshift and absorption intensity in the visible light range was observed with the increasing graphene content. The optimal ratio of BiOBr/graphene showed higher selective adsorption and degradation of anionic MO compared to BiOBr because of strong π-π interactions with dye through the conjugate aromatic structure. Furthermore, these aerogels serving as bulk catalysts can be easily retrieved from photocatalytic systems for recycling. Yang et al. [170] fabricated a composite aerogel by means of a two-step hydrothermal procedure, combining BiVO4 quantum tubes as the light-absorbing constituent with graphene as a structural support and photoelectron transport pathway. The improved interfacial charge carrier transfer leads to more efficient spatial charge separation. The BiVO4/graphene composite photocatalyst exhibited favorable catalytic performance, capable of degrading gaseous formaldehyde from 1.0 ppm to 0.4 ppm within 15 min, approximately three times faster than pristine BiVO4.
Metal-organic frameworks (MOFs) are crystalline materials characterized by their periodic network structure units, composed of metal ions/metal clusters and organic ligands connected through coordination covalent bonds. MOFs possess the advantages of abundant catalytic active sites, diverse pore structures, high surface area, and distinctive optical properties, rendering them an extremely appealing material in photocatalysis [171,172,173,174]. However, pristine MOFs still encounter challenges such as limited stability, weak conductivity, and inefficient separation of electron-hole pairs. To address these issues, integrating MOFs with graphene has proven to be an effective strategy. The hydrophobic nature of graphene enhances the stability of MOFs in aqueous environments, while π-π stacking interactions between graphene sheets and MOFs improve their overall performance [175]. For example, Zhao et al. [176] synthesized a series of DUT-67 (Zr-MOF)/graphene photocatalysts with 3D morphology through hydrothermal processing, enabling the selective catalytic conversion of CO2 by harnessing the synergistic effect between the macrostructure of graphene and DUT-67. The incorporation of graphene and aerogel structure enhanced the photo charge separation efficiency of DUT-67, thus promoting more charges to participate in the photocatalytic reaction. The optimized ratio of DUT-67/graphene composites achieved near 99.6% selectivity in converting CO2 to CO. Similarly, Shah et al. [177] developed a hybrid photocatalyst by linking 3D graphene with NH2-MIL-125 (Ti-MOF) through a solvothermal method with NH2-IL serving as bridging. The designed hybrid photocatalyst exhibited a higher degradation efficiency (97%) for acetaldehyde compared to pure NH2-MIL-125 (57%), attributed to the intimate interfacial contact, high dispersion, excellent surface area, and improved electrical conductivity.

5. Applications of 3D Graphene-Based Photocatalysts

In the preceding sections, we explored the structural advantages of three-dimensional graphene and various synthesis methods, along with the preparation and photocatalytic performance of graphene-based photocatalysts. This section will focus on their applications in environmental photocatalysis, specifically their roles in oxidizing organic pollutants, hydrogen evolution, and CO2 reduction. A summary of the photocatalytic activity and reaction conditions of these systems is presented in Table 2. Due to space limitations, this section will concentrate on the impact of 3D graphene on photocatalytic reactions.

5.1. Photocatalytic Degradation of Organic Pollutants

The removal of contaminants from water bodies is crucial due to their detrimental effects on aquatic ecosystems and potential risks to human health. Photocatalysis stands out as a promising approach for wastewater treatment, capable of the degradation of various toxic and organic pollutants. In general, understanding the mechanism of photocatalytic reactions is key to designing effective semiconductor photocatalysts by predicting charge carrier pathways. Yang et al. [178] synthesized a composite photocatalyst by integrating flower-like Bi2WO6 with 3D porous graphene through a hydrothermal approach for the degradation of MB and 2,4-dichlorophenol, as shown in Figure 13. In this photocatalytic degradation process, graphene acted as a supporter for Bi2WO6, enabling rapid adsorption and enrichment of pollutants onto the composite’s surface. Upon exposure to visible light, Bi2WO6 was excited to generate electrons and holes. The holes in the valence band partially reacted with H2O to generate •OH radicals, both of which possess strong oxidation capabilities, resulting in the complete oxidation of pollutants into H2O and CO2. Furthermore, electrons in the conduction band of Bi2WO6 efficiently transferred to the 3D structural graphene due to its excellent conductivity, significantly improving electron-hole separation efficiency and enhancing overall photocatalytic performance. Jin et al. [179] prepared TiO2 nanowires/graphene 3D framework material for the photocatalytic degradation of micro-organic contaminant ethenzamide. The interconnected structure provides convenient migration pathways and abundant pores to enhance reactant adsorption and capture light through the refraction-reflection effect. Furthermore, the superior electron transport capability of graphene and the chemical bonding between TiO2 facilitate the rapid transition of photoelectrons, resulting in more e reacting with dissolved oxygen to form •O2. Free radical trapping experiments confirmed that h+, •O2, and •OH all contributed to the photocatalytic degradation process. Dong et al. [180] fabricated a macroscopic monolithic ZnSnO3/graphene aerogel for the adsorption and visible light photocatalytic degradation of ciprofloxacin wastewater. The incorporation of 3D graphene optimized the interfacial and electronic band structure, mitigating photogenerated electron-hole pair recombination and generating more •O2 and •OH active species. Liu et al. [181] reported the fabrication of a Z-scheme-type CeVO4/3D graphene/BiVO4 ternary photocatalyst for visible-light-driven tetracycline degradation. Graphene plays a crucial role as an effective solid-state electron mediator in promoting electron-hole separation and improving light utilization. Das et al. [110] prepared 3D composites comprising MoS2 nanoflakes grown on graphene sheets via hydrothermal processing, followed by chemical activation with KOH solution under a high-temperature atmosphere to achieve more orderly mesopores and high-quality graphene layers for the photocatalytic degradation of tetracycline, as shown in Figure 14. The induction of mesoporosity augments light harvesting by facilitating both light penetration and scattering throughout the entire volume. In this photocatalytic system, the electrons photogenerated by MoS2 can be transferred to graphene on account of their work function differences, the interconnected graphene networks are capable of effectively inhibiting the recombination of electron/hole pairs by providing multidimensional transport channels. Reactive oxygen species scavenging experiments revealed that 1O2 and h+ primarily drive tetracycline dissociation, whereas •O2 and •OH play an auxiliary role in the photocatalytic process. Additionally, the degradation pathways are hypothesized based on intermediate compounds identification, with all these intermediates eventually mineralizing to CO2, H2O, NH4+, and NO3.

5.2. Photocatalytic Hydrogen Evolution

The conversion of abundant solar energy into chemical energy through photocatalysis has garnered significant attention in recent years. Specifically, water photolysis for hydrogen evolution is considered an effective strategy to alleviate energy and environmental issues. To satisfy the thermodynamic prerequisites for redox reactions, an active photocatalyst must possess an appropriate bandgap and more negative conduction band edge than the reduction potential of H+/H2, as well as a more positive valence band edge than the oxidation potential of H2O to O2 [182,183]. The development of robust photocatalysts to improve the H2 production rate is a recent research hotspot. Lu et al. [66] combined the widely used semiconductor TiO2 with 3D graphene to fabricate a composite photocatalyst for efficient hydrogen production, where TiO2 particles (~10 nm) are uniformly distributed over the surfaces of graphene sheets (Figure 15A). Multiple pathways for hydrogen evolution may exist in the photocatalytic system (Figure 15B): graphene can act as an independent photocatalyst, utilizing electrons generated under light to reduce water to hydrogen (path 1); photogenerated electrons may transfer between graphene and TiO2, boosting photocatalytic activity (paths 2a and 2b); or graphene can reduce the TiO2 bandgap, thus promoting electron transfer from the Fermi level of graphene to the conduction band of TiO2 (path 3) and suppressing the recombination of charge carriers. Samajdar et al. [184] fabricated a Na0.5Bi0.5TiO3/graphene heterogeneous photocatalyst with a 2D/3D interface for catalytic hydrogen evolution under visible light irradiation. In this composite, graphene exhibits higher conduction band and valence band potentials than Na0.5Bi0.5TiO3, resulting in the formation of a Z-scheme heterojunction at the interface when they are in close contact. The effective separation of photogenerated charge carriers and enhanced electron reduction ability led to a significantly increased rate of H2 generation. Zhang et al. [185] developed a composite graphene hydrogel incorporating CdS nanoparticles. The interfacial contact and energy level matching between them favor the directional transfer of photoelectrons from CdS to graphene for hydrogen generation. The unique 3D network structure of graphene not only exposes more active sites for catalysis but also suppresses photo corrosion and enhances the charge migration of CdS nanoparticles. Liu et al. [186] synthesized a ternary CdS/g-C3N4/graphene aerogel to enhance the photocatalytic H2 production activity. CdS nanoparticles and g-C3N4 sheets were uniformly distributed within the 3D hierarchical networks of graphene, creating an S-scheme heterojunction to broaden the optical absorption range and promote the separation of photogenerated carriers. Seo et al. [187] constructed edge-rich 3D structured metal chalcogenide/graphene for photocatalytic hydrogen generation by combining laser irradiation with metal-organic chemical vapor deposition for the growth of MoS2 and WS2 nanosheets on porous graphene, as depicted in Figure 16. In this photocatalytic system, photogenerated electrons partake in the hydrogen evolution reaction at specified active sites, while holes are seized by sacrificial agents. The high electrical conductivity of graphene operates synergistically with the heterojunction effect between Mo(W)S2 and graphene to facilitate photocarriers participation in the catalytic reaction. Moreover, the configuration of graphene improved the density of accessible active sites, thereby escalating photocatalytic performance. The hydrogen generation rate of MoS2/graphene and WS2/graphene was comparable to that of previously reported composite materials.

5.3. Photocatalytic CO2 Reduction

In recent decades, the excessive consumption of fossil fuels has led to a persistent rise in atmospheric CO2, resulting in global warming and posing substantial ecological and environmental challenges. Against the background of carbon neutrality, developing efficient CO2 capture strategies and converting them into high-value-added chemical products has become one of the most environmentally friendly and sustainable approaches [188,189]. Since the pioneering work of Inoue et al. [190] in 1979 on the photocatalytic reduction of CO2 using semiconductors, significant progress has been made in the study of CO2 conversion through photocatalysis. Nevertheless, the practical application of photocatalytic CO2 reduction is greatly constrained by factors such as absorption capability, light utilization efficiency, electron-hole recombination rate, etc. Especially, the activation of CO2 is exceedingly challenging due to the high dissociation energy of the C=O bond in its molecule [191,192,193]. Regarding this matter, a further enhancement of CO2 reduction efficiency catalyzed by graphene-based photocatalysts has been proposed. For example, Park et al. [194] reported the synthesis of a nanocomposite photocatalyst through the covalent attachment of TiO2 nanoparticles to pristine 3D graphene. The TiO2/graphene exhibited enhanced photocatalytic activity for CO2 reduction to CO compared to bare TiO2, which can be attributed to the high specific surface area, strong interactions, and excellent dispersion of TiO2 nanoparticles on graphene. Zhang et al. [195] promoted the activation and conversion of CO2 by in situ growing ZnO nanowire arrays on the surface of 3D N-doped graphene. The efficiency of ZnO/N-graphene photocatalytic CO2 reduction to CH3OH was increased by 2.3 times. In this composite, N-doped graphene not only facilitated the uniform growth of ZnO but also improved the separation of electron-hole pairs and functioned as active sites for the adsorption and reduction of CO2. Song et al. [196] developed MOF-808/graphene aerogel materials for the photocatalytic reduction of CO2 to CO. The improved photothermal and photoelectric conversion efficiencies were associated with the three-dimensional macroscopic structure, which boasted a large specific surface area, abundant internal pore structure, and increased active sites. These factors collectively contribute to the enhancement of light energy utilization and the acceleration of the electron transfer rate. Xia et al. [197] reported a hierarchical composite photocatalyst for solar-driven CO2 reduction, created by assembling vertically aligned ZnIn2S4 nanowall arrays on N-doped graphene foam via hydrothermal synthesis, as illustrated in Figure 17A. This composite maintained the integrity of the 3D network structure while anchoring the ZnIn2S4 nanowalls securely onto the N-doped graphene surface (Figure 17B). ISI-XPS combined with flat-band potential measurements was employed to elucidate the energy band structure and photocatalytic mechanism (Figure 17C). Analysis using ISI-XPS and flat-band potential measurements revealed that electrons transfer from ZnIn2S4 to N-doped graphene when their Fermi levels align. Under simulated sunlight irradiation, photoelectrons migrate from ZnIn2S4 to N-doped graphene through the heterojunction interface. The N dopants possess a relatively strong electron affinity, functioning as polar sites and electron collectors to attract CO2 molecules and facilitate the reduction reaction of CO2 with photoelectrons on the graphene surface.
Table 2. Summary of photocatalytic applications over some 3D graphene-based photocatalysts.
Table 2. Summary of photocatalytic applications over some 3D graphene-based photocatalysts.
PhotocatalystLight SourceReaction TypeReaction ConditionsPerformanceRef
Bi2WO6/grapheneXenon lamp (500 W)
>420 nm
Degradation40 ppm MB, 50 mL
5 ppm 2,4-CDP, 50 mL
Static system: 50.6% Dynamic system: 28.1% (MB), 17% (2,4-CDP)[178]
TiO2/grapheneUV lamp (24 W)
VUV lamp (1.2 W)
Degradation500 ppb ethenzamide, 150 mL98.5%, 60 min
99.8%, 3 min
[179]
ZnSnO3/grapheneXenon lampDegradation100 mg/L CIP, 100 mLalmost 100%, 120 min[180]
CeVO4/graphene/BiVO4Xenon lamp (500 W)Degradation20 mg/L TC, 100 mL100%, 60 min[181]
MoS2/grapheneMercury lamp (250 W, with UV cutoff filter)Degradation5 mg/L TC, 100 mL97%, 120 min[110]
TiO2/grapheneXenon lamp (200 W) 320~780 nmH2 evolutionmethanol (10 vol%)1205 μmol·h−1·g−1[66]
Na0.5Bi0.5TiO3/grapheneXenon lamp (250 W)H2 evolutionmethanol (25 vol%)100 mmol·h−1·g−1[184]
CdS/grapheneXenon lamp (500 W) 320~780 nmH2 evolution0.1 M Na2S·9H2O,
0.5 M Na2SO3
213.358 μmol·h−1·g−1[185]
CdS/g-C3N4/grapheneXenon lamp (300 W)
>420 nm
H2 evolutionTEOA (5 vol%)86.38 μmol·h−1·g−1[186]
MoS2/graphene WS2/grapheneXenon lamp (150 W)H2 evolution0.35 M Na2S,
0.25 M Na2SO3
6.51 mmol·h−1·g−1
7.26 mmol·h−1·g−1
[187]
TiO2/grapheneMercury lamp (200 W)CO2 reductionCO2,
triethylamine vapor
CO conversion:
1.26 μmol·mg−1
[194]
ZnO/N-grapheneXenon lamp (300 W)CO2 reduction84 mg NaHCO3,
2M H2SO4
CH3OH conversion:
1.51 μmol·h−1·g−1
[195]
MOF-808/grapheneXenon lampCO2 reductionCO2, H2OCO conversion:
14.35 μmol·g−1
[196]
ZnIn2S4/N-grapheneXenon lamp (300 W)CO2 reduction84 mg NaHCO3,
2M H2SO4
CH4: 1.01 μmol·h−1·g−1
CO: 2.45 μmol·h−1·g−1 CH3OH: 1.37 μmol·h−1·g−1
[197]

6. Summary and Perspectives

Over the past decade, extensive research efforts have been dedicated to the systematic investigation of three-dimensional graphene materials owing to their distinctive structural and morphological characteristics. Incorporating photocatalytic active components into the graphene networks has created new possibilities for synergistically improving the performance of various photocatalytic processes. In this review, we commence by outlining the primary synthetic strategies for 3D graphene and graphene-based architectures, encompassing hard template methods (CVD, microsphere, and ice template methods), soft template methods, and non-template methods (self-assembly, chemical reduction, and 3D printing). It is evident that the selection of appropriate synthesis methods and optimization conditions is crucial for customizing the macroscopic structure, porosity, micro-morphology, defects, and surface/interface properties of 3D materials. Furthermore, recent advancements in the synthesis of 3D graphene-based composite photocatalysts are summarized herein; these include metal oxides, metal sulfides, g-C3N4, bismuth-based compounds, and MOFs. And we also attempt to comprehensively understand the multifaceted roles of 3D graphene in enhancing photoredox performance as well as the key factors influencing their catalytic activity. Finally, we elaborate on the applications of 3D graphene-based photocatalysts in organic pollutant degradation, hydrogen evolution, and CO2 reduction.
Despite the ongoing advancements in the design and synthesis of 3D graphene-based photocatalysts and understanding their reaction mechanisms, several challenges remain that require further attention. Firstly, with regard to practical applications, there is a pressing need to explore and optimize innovative approaches for the large-scale synthesis of highly active yet cost-effective photocatalysts in future research. Substantial efforts are also necessary for reactor design, stability, and the recyclability of photocatalysts to enable repeated use. Additionally, synergies between the photoactive components and 3D graphene skeleton are commonly cited to elucidate the enhanced photocatalytic properties. With the rapid advancement in experimental techniques, approaches such as in situ characterization and theoretical simulation can provide valuable insights into the growth mechanisms of photoactive materials on graphene. These approaches can also shed light on charge transfer dynamics and the real-time transformation pathways of reactant molecules and intermediates.
The brief history of 3D graphene-based photocatalysis showcases impressive progress, and we sincerely hope that this review may give some inspiration for further research in this field. From the initial laboratory investigations to practical application, the development of novel photocatalysts based on 3D graphene will be a lengthy process requiring constant efforts and substantial breakthroughs. However, it is certain that the integration of multidisciplinary knowledge will continue to open up the potential applications of 3D graphene-based photocatalysts in environmental and energy-related fields in the near future.

Author Contributions

Writing-original draft preparation, F.Z.; review and editing, J.L. and L.H.; project administration and funding acquisition, C.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Postdoctoral Scientific Research Foundation of Shenyang Ligong University (No.1010148001402), the Key Laboratory of Weapon Science & Technology Research (LJ232410144071), and the Light-Selection Team Plan of Shenyang Ligong University (SYLUGXTD5).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Mei, W.; Lu, H.; Dong, R.; Tang, S.; Xu, J. Ce-Doped Brookite TiO2 Quasi-Nanocubes for Efficient Visible Light Catalytic Degradation of Tetracycline. ACS Appl. Nano Mater. 2024, 7, 1825–1834. [Google Scholar] [CrossRef]
  2. He, J.; Hu, L.; Shao, C.; Jiang, S.; Sun, C.; Song, S. Photocatalytic H2O Overall Splitting into H2 Bubbles by Single Atomic Sulfur Vacancy CdS with Spin Polarization Electric Field. ACS Nano 2021, 15, 18006–18013. [Google Scholar] [CrossRef]
  3. Wang, Y.; Jia, K.; Pan, Q.; Xu, Y.; Liu, Q.; Cui, G.; Guo, X.; Sun, X. Boron-Doped TiO2 for Efficient Electrocatalytic N2 Fixation to NH3 at Ambient Conditions. ACS Sustain. Chem. Eng. 2018, 7, 117–122. [Google Scholar] [CrossRef]
  4. Basyach, P.; Prajapati, P.K.; Rohman, S.S.; Sonowal, K.; Kalita, L.; Malik, A.; Guha, A.K.; Jain, S.L.; Saikia, L. Visible Light-Active Ternary Heterojunction Photocatalyst for Efficient CO2 Reduction with Simultaneous Amine Oxidation and Sustainable H2O2 Production. ACS Appl. Mater. Interfaces 2023, 15, 914–931. [Google Scholar] [CrossRef]
  5. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  6. Lin, Z.; Jiang, X.; Xu, W.; Li, F.; Chen, X.; Wang, H.; Liu, S.; Lu, X. The effects of water, substrate, and intermediate adsorption on the photocatalytic decomposition of air pollutants over nano-TiO2 photocatalysts. Phys. Chem. Chem. Phys. 2024, 26, 662–678. [Google Scholar] [CrossRef]
  7. Ding, Y.; Yang, G.; Zheng, S.; Gao, X.; Xiang, Z.; Gao, M.; Wang, C.; Liu, M.; Zhong, J. Advanced photocatalytic disinfection mechanisms and their challenges. J. Environ. Manag. 2024, 366, 121875. [Google Scholar] [CrossRef]
  8. Batool, S.S.; Saleem, R.; Khan, R.R.M.; Saeed, Z.; Pervaiz, M.; Summer, M. Enhancing photocatalytic performance of zirconia-based nanoparticles: A comprehensive review of factors, doping strategies, and mechanisms. Mater. Sci. Semicond. Process. 2024, 178, 108419. [Google Scholar] [CrossRef]
  9. Han, H.; Khan, M.R.; Ahmad, I.; Al-Qattan, A.; Ali, I.; Karim, M.R.; Bayahia, H.; Khan, F.S.; Ahmad, Z.; Ullah, S. Hierarchical S-scheme heterojunction systems: A comprehensive review on outstanding photocatalysts. J. Water Process Eng. 2024, 61, 105346. [Google Scholar] [CrossRef]
  10. Wang, S.; Teramura, K.; Hisatomi, T.; Domen, K.; Asakura, H.; Hosokawa, S.; Tanaka, T. Effective Driving of Ag-Loaded and Al-Doped SrTiO3 under Irradiation at λ > 300 nm for the Photocatalytic Conversion of CO2 by H2O. ACS Appl. Energy Mater. 2020, 3, 1468–1475. [Google Scholar] [CrossRef]
  11. Zhang, J.; Deng, P.; Deng, M.; Shen, H.; Feng, Z.; Li, H. Hybrid Density Functional Theory Study of Native Defects and Nonmetal (C, N, S, and P) Doping in a Bi2WO6 Photocatalyst. ACS Omega 2020, 5, 29081–29091. [Google Scholar] [CrossRef] [PubMed]
  12. Li, X.; Dai, K.; Pan, C.; Zhang, J. Diethylenetriamine-Functionalized CdS Nanoparticles Decorated on Cu2S Snowflake Microparticles for Photocatalytic Hydrogen Production. ACS Appl. Nano Mater. 2020, 3, 11517–11526. [Google Scholar] [CrossRef]
  13. Liu, B.; Bie, C.; Zhang, Y.; Wang, L.; Li, Y.; Yu, J. Hierarchically Porous ZnO/g-C3N4 S-Scheme Heterojunction Photocatalyst for Efficient H(2)O(2) Production. Langmuir 2021, 37, 14114–14124. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, P.; Wang, J.; Ming, T.; Wang, X.; Yu, H.; Yu, J.; Wang, Y.; Lei, M. Dye-sensitization-induced visible-light reduction of graphene oxide for the enhanced TiO2 photocatalytic performance. ACS Appl. Mater. Interfaces 2013, 5, 2924–2929. [Google Scholar] [CrossRef] [PubMed]
  15. Chibac-Scutaru, A.L.; Podasca, V.-E.; Dascalu, I.A.; Rusu, D.; Melinte, V. ZnO nanostructures with controlled morphological and optical properties for applications as efficient photocatalyst for malachite green degradation. Ceram. Int. 2024, 50, 34291–34303. [Google Scholar] [CrossRef]
  16. Dufour, F.; Pigeot-Remy, S.; Durupthy, O.; Cassaignon, S.; Ruaux, V.; Torelli, S.; Mariey, L.; Maugé, F.; Chanéac, C. Morphological control of TiO2 anatase nanoparticles: What is the good surface property to obtain efficient photocatalysts? Appl. Catal. B Environ. 2015, 174–175, 350–360. [Google Scholar] [CrossRef]
  17. Petrisková, P.; Monfort, O.; Satrapinskyy, L.; Dobročka, E.; Plecenik, T.; Plesch, G.; Papšík, R.; Bermejo, R.; Lenčéš, Z. Preparation and photocatalytic activity of TiO2 nanotube arrays prepared on transparent spinel substrate. Ceram. Int. 2021, 47, 12970–12980. [Google Scholar] [CrossRef]
  18. Kapusuz Yavuz, D.; El Accen, M.; Bedir, M. Unlocking the potential of ZnO nanorods: Structural insights for enhanced photocatalytic activity. J. Phys. Chem. Solids 2024, 193, 112120. [Google Scholar] [CrossRef]
  19. Wang, Z.; Yang, B.; Zhao, X.; Chen, Y.; Wei, D.; Zhang, L.; Su, X. Facile synthesis of ultrathin γ-Fe2O3 magnetic nanosheets rich in oxygen vacancies and their photocatalytic activity for water oxidation. Appl. Surf. Sci. 2022, 578, 151999. [Google Scholar] [CrossRef]
  20. Xu, Z.; Yu, Y.; Fang, D.; Liang, J.; Zhou, L. Simulated solarlight catalytic reduction of Cr(VI) on microwave–ultrasonication synthesized flower-like CuO in the presence of tartaric acid. Mater. Chem. Phys. 2016, 171, 386–393. [Google Scholar] [CrossRef]
  21. Hkiri, K.; Mohamed, H.E.A.; Harrisankar, N.; Gibaud, A.; van Steen, E.; Maaza, M. Environmental water treatment with green synthesized WO3 nanoflakes for cationic and anionic dyes removal: Photocatalytic studies. Catal. Commun. 2024, 187, 106851. [Google Scholar] [CrossRef]
  22. Ravi, A.; Annamalai, P.; Sankar, V.; Achutharaman, K.R.; Valdes, H.; SaravanaVadivu, A.; MuthaiahPillai, V.; Alharbi, S.A. Sustainable synthesis, superior performance: Nanoflower-like α-Bi2O3 from solvent-free solid state for photocatalytic crystal violet degradation. J. Taiwan Inst. Chem. E 2024, 157, 105413. [Google Scholar] [CrossRef]
  23. Tran, D.P.H.; You, S.J.; Wang, Y.F. Morphological optimization of CeO2 via low-temperature hydrothermal synthesis for enhanced photocatalytic performance in NO oxidation and CO2 conversion. J. Environ. Chem. Eng. 2024, 12, 112667. [Google Scholar] [CrossRef]
  24. Liu, H.; Zhou, H.; Yang, L.; Pan, Y.; Zhao, X.; Wang, F.; Fang, R.; Li, Y. Photocatalytic CO2 reduction coupled with biomass-based amines oxidation over double-shelled CdS nanocages. Catal. Commun. 2024, 187, 106884. [Google Scholar] [CrossRef]
  25. Hu, B.; Wang, B.H.; Bai, Z.J.; Chen, L.; Guo, J.K.; Shen, S.; Xie, T.L.; Au, C.T.; Jiang, L.L.; Yin, S.F. Regulating MoS2 edge site for photocatalytic nitrogen fixation: A theoretical and experimental study. Chem. Eng. J. 2022, 442, 136211. [Google Scholar] [CrossRef]
  26. Morales Gallardo, M.V.; Ayala, A.M.; Pal, M.; Cortes Jacome, M.A.; Toledo Antonio, J.A.; Mathews, N.R. Synthesis of pyrite FeS2 nanorods by simple hydrothermal method and its photocatalytic activity. Chem. Phys. Lett. 2016, 660, 93–98. [Google Scholar] [CrossRef]
  27. Yang, Y.; Chen, X.; Pan, Y.; Song, H.; Zhu, B.; Wu, Y. Two-dimensional ZnS (propylamine) photocatalyst for efficient visible light photocatalytic H2 production. Catal. Today 2021, 374, 4–11. [Google Scholar] [CrossRef]
  28. Yoo, J.-H.; Ji, M.; Kim, J.H.; Ryu, C.-H.; Lee, Y.-I. Facile synthesis of hierarchical CuS microspheres with high visible-light-driven photocatalytic activity. JPPA 2020, 401, 112782. [Google Scholar] [CrossRef]
  29. Lan, M.; Wang, Y.; Dong, X.; Yang, F.; Zheng, N.; Wang, Y.; Ma, H.; Zhang, X. Controllable fabrication of sulfur-vacancy-rich Bi2S3 nanorods with efficient near-infrared light photocatalytic for nitrogen fixation. Appl. Surf. Sci. 2022, 591, 153205. [Google Scholar] [CrossRef]
  30. Liu, K.; Shen, X.; Shen, H.; Jiang, K.; Sun, T.; Wang, M. One-step fabrication of novel BiVO4 hollow nanotubes with improved photocatalytic activity. Mater. Lett. 2023, 349, 134792. [Google Scholar] [CrossRef]
  31. Pawar, R.C.; Khan, H.; Charles, H.; Lee, C.S. Growth of 3D nanowall-like structures of FeVO4 by controlling reaction rate for effective CO2 reduction using UV-visible light. J. Environ. Chem. Eng. 2023, 11, 110236. [Google Scholar] [CrossRef]
  32. Mahboob, I.; Shafiq, I.; Shafique, S.; Akhter, P.; Amjad, U.-e.-S.; Hussain, M.; Park, Y.-K. Effect of active species scavengers in photocatalytic desulfurization of hydrocracker diesel using mesoporous Ag3VO4. Chem. Eng. J. 2022, 441, 136063. [Google Scholar] [CrossRef]
  33. Yan, Y.; Cai, F.; Song, Y.; Shi, W. InVO4 nanocrystal photocatalysts: Microwave-assisted synthesis and size-dependent activities of hydrogen production from water splitting under visible light. Chem. Eng. J. 2013, 233, 1–7. [Google Scholar] [CrossRef]
  34. Yang, T.; Tang, Y.; Yang, F.; Qu, J.; Yang, X.; Cai, Y.; Du, F.; Li, C.M.; Hu, J. Molten salt-assisted anti-defect engineering to tailor ordered, highly crystalline g-C3N4 nanorods for efficient photocatalytic H2O2 production. Chem. Eng. J. 2023, 475, 146497. [Google Scholar] [CrossRef]
  35. Shen, Y.; Du, X.; Shi, Y.; Nguetsa Kuate, L.J.; Chen, Z.; Zhu, C.; Tan, L.; Guo, F.; Li, S.; Shi, W. Bound-state electrons synergy over photochromic high-crystalline C3N5 nanosheets in enhancing charge separation for photocatalytic H2 production. APM 2024, 3, 100202. [Google Scholar] [CrossRef]
  36. Wei, Z.; Xu, W.; Peng, P.; Sun, Q.; Li, Y.; Ding, N.; Zhao, C.; Li, S.; Pang, S. Covalent synthesis of Ti-MOF for enhanced photocatalytic CO2 reduction. Mol. Catal. 2024, 558, 114042. [Google Scholar] [CrossRef]
  37. Song, K.; Liang, S.; Zhong, X.; Wang, M.; Mo, X.; Lei, X.; Lin, Z. Tailoring the crystal forms of the Ni-MOF catalysts for enhanced photocatalytic CO2-to-CO performance. Appl. Catal. B Environ. 2022, 309, 121232. [Google Scholar] [CrossRef]
  38. Li, S.; Wu, F.; Lin, R.; Wang, J.; Li, C.; Li, Z.; Jiang, J.; Xiong, Y. Enabling photocatalytic hydrogen production over Fe-based MOFs by refining band structure with dye sensitization. Chem. Eng. J. 2022, 429, 132217. [Google Scholar] [CrossRef]
  39. Lu, X.Y.; Zhang, M.L.; Ren, Y.X.; Wang, J.J.; Yang, X.G. Design of Co-MOF nanosheets for efficient adsorption and photocatalytic degradation of organic dyes. J. Mol. Struct. 2023, 1288, 135796. [Google Scholar] [CrossRef]
  40. Xu, T.; Zhang, L.; Cheng, H.; Zhu, Y. Significantly enhanced photocatalytic performance of ZnO via graphene hybridization and the mechanism study. Appl. Catal. B Environ. 2011, 101, 382–387. [Google Scholar] [CrossRef]
  41. Yue, G.; Lin, J.Y.; Tai, S.Y.; Xiao, Y.; Wu, J. A catalytic composite film of MoS2/graphene flake as a counter electrode for Pt-free dye-sensitized solar cells. Electrochim. Acta 2012, 85, 162–168. [Google Scholar] [CrossRef]
  42. Zhang, Y.; Zhou, Z.; Chen, T.; Wang, H.; Lu, W. Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol. J. Environ. Sci. 2014, 26, 2114–2122. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, P.; Zhan, S.; Xia, Y.; Ma, S.; Zhou, Q.; Li, Y. The fundamental role and mechanism of reduced graphene oxide in rGO/Pt-TiO2 nanocomposite for high-performance photocatalytic water splitting. Appl. Catal. B Environ. 2017, 207, 335–346. [Google Scholar] [CrossRef]
  44. Lu, H.; Zhao, Y.M.; Saji, S.E.; Yin, X.; Wibowo, A.; Tang, C.S.; Xi, S.; Cao, P.; Tebyetekerwa, M.; Liu, B.; et al. All room-temperature synthesis, N2 photofixation and reactivation over 2D cobalt oxides. Appl. Catal. B Environ. 2022, 304, 121001. [Google Scholar] [CrossRef]
  45. Otgonbayar, Z.; Liu, Y.; Oh, W.C. Design of type-II heterojunction MoS2/gC3N4/graphene ternary nanocomposite for photocatalytic CO2 reduction to hydrocarbon fuels: In aqueous solvents with a sacrificial donor. J. Environ. Chem. Eng. 2023, 11, 109884. [Google Scholar] [CrossRef]
  46. Wang, Y.; Xu, F.; Zhou, L.; Li, H.; Meng, Q.; Jing, L.; Tian, Z.; Hou, C. 2D N-doped graphene/CoFe MOFs heterostructure functionalized CNF aerogels impart highly efficient photocatalytic oxidation of gaseous VOCs. J. Environ. Chem. Eng. 2024, 12, 112225. [Google Scholar] [CrossRef]
  47. Zhang, X.; Shi, W.; Wang, T.; Wang, X.; Li, C.; He, W. Investigation on the dual roles of pollutants adsorption on the catalyst surface during photocatalytic process. J. Water Process Eng. 2024, 63, 105558. [Google Scholar] [CrossRef]
  48. He, Y.; Yin, L.; Yuan, N.; Zhang, G. Adsorption and activation, active site and reaction pathway of photocatalytic CO2 reduction: A review. Chem. Eng. J. 2024, 481, 148754. [Google Scholar] [CrossRef]
  49. Xiong, T.; Ye, Y.; Luo, B.; Shen, L.; Wang, D.; Fan, M.; Gong, Z. Facile fabrication of 3D TiO2-graphene aerogel composite with enhanced adsorption and solar light-driven photocatalytic activity. Ceram. Int. 2021, 47, 14290–14300. [Google Scholar] [CrossRef]
  50. Li, Y.; Yang, J.; Zheng, S.; Zeng, W.; Zhao, N.; Shen, M. One-pot synthesis of 3D TiO2-reduced graphene oxide aerogels with superior adsorption capacity and enhanced visible-light photocatalytic performance. Ceram. Int. 2016, 42, 19091–19096. [Google Scholar] [CrossRef]
  51. Li, Q.; Sun, Z.; Yin, C.; Chen, Y.; Pan, D.; Yu, B.; Zhang, Y.; He, T.; Chen, S. Template-assisted synthesis of ultrathin graphene aerogels as bifunctional oxygen electrocatalysts for water splitting and alkaline/neutral zinc-air batteries. Chem. Eng. J. 2023, 458, 141492. [Google Scholar] [CrossRef]
  52. Elsehsah, K.A.A.A.; Noorden, Z.A.; Saman, N.M. Graphene aerogel electrodes: A review of synthesis methods for high-performance supercapacitors. J. Energy Storage 2024, 97, 112788. [Google Scholar] [CrossRef]
  53. Jiang, W.; Wang, H.; Xu, Z.; Li, N.; Chen, C.; Li, C.; Li, J.; Lv, H.; Kuang, L.; Tian, X. A review on manifold synthetic and reprocessing methods of 3D porous graphene-based architecture for Li-ion anode. Chem. Eng. J. 2018, 335, 954–969. [Google Scholar] [CrossRef]
  54. Wu, Y.; Zhu, J.; Huang, L. A review of three-dimensional graphene-based materials: Synthesis and applications to energy conversion/storage and environment. Carbon 2019, 143, 610–640. [Google Scholar] [CrossRef]
  55. Thakur, A. Graphene aerogel based energy storage materials—A review. Mater. Today 2022, 65, 3369–3376. [Google Scholar] [CrossRef]
  56. Sharma, R.; Jaiswal, S.; Chauhan, R.; Bhardwaj, M.; Verma, K.; Dwivedi, J.; Sharma, S. Applications of graphene-based photocatalysts for efficient functionalized degradation of some common antibiotics. Inorg. Chem. Commun. 2024, 168, 112941. [Google Scholar] [CrossRef]
  57. Zhang, H.; Yang, X.; Zhang, X.; Wang, L.; Wang, B.; Zeng, X.; Ren, B. Graphene-based aerogel for efficient oil sorption and water pollution remediation. Colloids Surf. A 2024, 699, 134693. [Google Scholar] [CrossRef]
  58. Gao, B.; Feng, X.; Zhang, Y.; Zhou, Z.; Wei, J.; Qiao, R.; Bi, F.; Liu, N.; Zhang, X. Graphene-based aerogels in water and air treatment: A review. Chem. Eng. J. 2024, 484, 149604. [Google Scholar] [CrossRef]
  59. Ahmed, M.A.; Mohamed, A.A. Recent progress in semiconductor/graphene photocatalysts: Synthesis, photocatalytic applications, and challenges. RSC Adv. 2022, 13, 421–439. [Google Scholar] [CrossRef] [PubMed]
  60. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.W.; Potts, J.R.; Ruoff, R.S. Graphene and graphene oxide: Synthesis, properties, and applications. Adv. Mater. 2010, 22, 3906–3924. [Google Scholar] [CrossRef] [PubMed]
  61. Hashmi, A.; Nayak, V.; Singh, K.R.; Jain, B.; Baid, M.; Alexis, F.; Singh, A.K. Potentialities of graphene and its allied derivatives to combat against SARS-CoV-2 infection. Mater. Today Adv. 2022, 13, 100208. [Google Scholar] [CrossRef] [PubMed]
  62. Samriti; Manisha; Chen, Z.; Sun, S.; Prakash, J. Design and engineering of graphene nanostructures as independent solar-driven photocatalysts for emerging applications in the field of energy and environment. Mol. Syst. Des. Eng. 2022, 7, 213–238. [Google Scholar] [CrossRef]
  63. Gao, W.; Zhang, S.; Wang, G.; Cui, J.; Lu, Y.; Rong, X.; Gao, C. A review on mechanism, applications and influencing factors of carbon quantum dots based photocatalysis. Ceram. Int. 2022, 48, 35986–35999. [Google Scholar] [CrossRef]
  64. Shyamala, R.; Gomathi Devi, L. Reduced graphene oxide/SnO2 nanocomposites for the photocatalytic degradation of rhodamine B: Preparation, characterization, photosensitization, vectorial charge transfer mechanism and identification of reaction intermediates. Chem. Phys. Lett. 2020, 748, 137385. [Google Scholar] [CrossRef]
  65. Bramhaiah, K.; Bhattacharyya, S. Challenges and future prospects of graphene-based hybrids for solar fuel generation: Moving towards next generation photocatalysts. Mater. Adv. 2022, 3, 142–172. [Google Scholar] [CrossRef]
  66. Lu, Y.; Ma, B.; Yang, Y.; Huang, E.; Ge, Z.; Zhang, T.; Zhang, S.; Li, L.; Guan, N.; Ma, Y.; et al. High activity of hot electrons from bulk 3D graphene materials for efficient photocatalytic hydrogen production. Nano Res. 2017, 10, 1662–1672. [Google Scholar] [CrossRef]
  67. Patnaik, S.; Behera, A.; Parida, K. A review on g-C3N4/graphene nanocomposites: Multifunctional roles of graphene in the nanohybrid photocatalyst toward photocatalytic applications. Catal. Sci. Technol. 2021, 11, 6018–6040. [Google Scholar] [CrossRef]
  68. Cheng, G.; Wang, X.; He, Y. 3D graphene paraffin composites based on sponge skeleton for photo thermal conversion and energy storage. Appl. Therm. Eng. 2020, 178, 115560. [Google Scholar] [CrossRef]
  69. Zhu, Z.; Wang, Z.; Ba, Z.; Li, X.; Dong, J.; Fang, Y.; Zhang, Q.; Zhao, X. 3D MXene-holey graphene hydrogel for supercapacitor with superior energy storage. J. Energy Storage 2022, 47, 103911. [Google Scholar] [CrossRef]
  70. Singh, M.; Dawsey, T.; Gupta, R.K. Recent advancement in 3D graphene for metal-sulfur batteries. J. Energy Storage 2023, 73, 109059. [Google Scholar] [CrossRef]
  71. Liu, H.; Sun, F.; Shao, H.; Yin, D.; Li, X.; Ma, Q.; Liu, G.; Yu, H.; Yu, W.; Dong, X. One-pot synthesis of ZnIn2S4 graphene aerogels with S vacancies for efficient photocatalytic reduction of Cr(VI) and hydrogen evolution. J. Environ. Chem. Eng. 2023, 11, 110181. [Google Scholar] [CrossRef]
  72. Zhao, Y.; Xu, Z.; Li, M.; Zhou, L.; Liu, M.; Yang, D.; Zeng, J.; Xie, R.; Hu, W.; Dong, F. S defect-rich MoS2 aerogel with hierarchical porous structure: Efficient photocatalysis and convenient reuse for removal of organic dyes. Chemosphere 2024, 354, 141649. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, Z.; Liu, H.; Lei, Z.; Huang, L.; Wu, T.; Liu, S.; Ye, G.; Lu, Y.; Wang, X. Graphene aerogel for photocatalysis-assist uranium elimination under visible light and air atmosphere. Chem. Eng. J. 2020, 402, 126256. [Google Scholar] [CrossRef]
  74. Noer Addina, S.; Idrus Alhamid, M.; Usman, A.; Adi Prasetyanto, E.; Sufyan, M.; Kusrini, E. Economic analysis of graphene aerogel polyurethane sponge production as adsorbent for oil-water separation. Mater. Today 2022, 51, 1896–1907. [Google Scholar] [CrossRef]
  75. Chen, L.; Liu, S.; Guo, X.; Wang, S.; He, Z.; Xu, Q.; Zhang, Q.; Zhu, J.; Zhao, P.; Yang, S.; et al. Ultralight, superhydrophobic and fire-resistant nitrogen-doped graphene aerogel for oil/water separation. Sep. Purif. Technol. 2024, 330, 125192. [Google Scholar] [CrossRef]
  76. Liu, Z.; Gao, B.; Zhao, P.; Fu, H.; Kamali, A.R. Magnetic luffa/graphene/CuFe2O4 sponge for efficient oil/water separation. Sep. Purif. Technol. 2024, 337, 126347. [Google Scholar] [CrossRef]
  77. Carvalho, A.S.; Santos, A.R.; Cabral, D.C.O.; Oliveira, D.M.; Assis, L.K.C.S.; França, E.L.T.; Quirino, F.R.S.; Castro-Lopes, S.; da Costa, O.M.M.M.; Padrón-Hernández, E. Binder-free ultrathin pellets of nanocomposites based on Fe3O4@nitrogen-doped reduced graphene oxide aerogel for electromagnetic interference shielding. J. Alloys Compd. 2024, 978, 173329. [Google Scholar] [CrossRef]
  78. Liu, H.; Xu, Y.; Yang, K.; Han, D.; Yong, H.; Yang, Q. Alternating current assisted preparation of three-dimensional graphene aerogels and the application in electromagnetic interference shielding. Mater. Today Phys. 2024, 40, 101308. [Google Scholar] [CrossRef]
  79. Cao, Y.; Cheng, Z.; Wang, R.; Liu, X.; Zhang, T.; Fan, F.; Huang, Y. Multifunctional graphene/carbon fiber aerogels toward compatible electromagnetic wave absorption and shielding in gigahertz and terahertz bands with optimized radar cross section. Carbon 2022, 199, 333–346. [Google Scholar] [CrossRef]
  80. González, M.; Baselga, J.; Pozuelo, J. Modulating the electromagnetic shielding mechanisms by thermal treatment of high porosity graphene aerogels. Carbon 2019, 147, 27–34. [Google Scholar] [CrossRef]
  81. Mao, R.; Yao, W.; Qadir, A.; Chen, W.; Gao, W.; Xu, Y.; Hu, H. 3-D graphene aerogel sphere-based flexible sensors for healthcare applications. Sens. Actuat. A Phys. 2020, 312, 112144. [Google Scholar] [CrossRef]
  82. Xie, M.; Qian, G.; Yu, Y.; Chen, C.; Li, H.; Li, D. High-performance flexible reduced graphene oxide/polyimide nanocomposite aerogels fabricated by double crosslinking strategy for piezoresistive sensor application. Chem. Eng. J. 2024, 480, 148203. [Google Scholar] [CrossRef]
  83. Deng, Z.; Gao, C.; Feng, S.; Zhang, H.; Liu, Y.; Zhu, Y.; Wang, J.; Xiang, X.; Xie, H. Highly Compressible, Light-Weight and robust Nitrogen-Doped graphene composite aerogel for sensitive pressure sensors. Chem. Eng. J. 2023, 471, 144790. [Google Scholar] [CrossRef]
  84. Yao, W.; Mao, R.; Gao, W.; Chen, W.; Xu, Z.; Gao, C. Piezoresistive effect of superelastic graphene aerogel spheres. Carbon 2020, 158, 418–425. [Google Scholar] [CrossRef]
  85. Chen, Y.; Pang, K.; Liu, X.; Li, K.; Lu, J.; Cai, S.; Liu, Y.; Xu, Z.; Gao, C. Environment-adaptive, anti-fatigue thermal interface graphene foam. Carbon 2023, 212, 118142. [Google Scholar] [CrossRef]
  86. Lu, Z.; Zhao, Z.; Gong, W.; Zhang, H.; Pang, M.; Tang, X.-Z. The impact of reduction mode on the microwave absorption characteristics of reduced graphene oxide aerogels. J. Alloys Compd. 2024, 991, 174451. [Google Scholar] [CrossRef]
  87. Xi, J.; Li, Y.; Zhou, E.; Liu, Y.; Gao, W.; Guo, Y.; Ying, J.; Chen, Z.; Chen, G.; Gao, C. Graphene aerogel films with expansion enhancement effect of high-performance electromagnetic interference shielding. Carbon 2018, 135, 44–51. [Google Scholar] [CrossRef]
  88. Li, G.; Fang, D.; Hong, G.; Eychmüller, A.; Zhang, X.; Song, W. Assembling graphene aerogel hollow fibres for solar steam generation. Compos. Commun. 2022, 35, 101302. [Google Scholar] [CrossRef]
  89. Sunahiro, S.; Pirabul, K.; Pan, Z.; Yoshii, T.; Hayasaka, Y.; Zhao, Q.; Crespo-Otero, R.; Di Tommaso, D.; Kyotani, T.; Nishihara, H. Synthesis of graphene mesosponge using CaO nanoparticles formed from CaCO3. Catal. Today 2024, 437, 114763. [Google Scholar] [CrossRef]
  90. Chen, Z.; Ren, W.; Gao, L.; Liu, B.; Pei, S.; Cheng, H.M. Three-dimensional flexible and conductive interconnected graphene networks grown by chemical vapour deposition. Nat. Mater. 2011, 10, 424–428. [Google Scholar] [CrossRef]
  91. Xia, D.; Yi, K.; Zheng, B.; Li, M.; Qi, G.; Cai, Z.; Cao, M.; Liu, D.; Peng, L.; Wei, D.; et al. Solvent-Free Process to Produce Three Dimensional Graphene Network with High Electrochemical Stability. J. Phys. Chem. C 2017, 121, 3062–3069. [Google Scholar] [CrossRef]
  92. Jia, J.; Sun, X.; Lin, X.; Shen, X.; Mai, Y.; Kim, J. Exceptional Electrical Conductivity and Fracture Resistance of 3D Interconnected Graphene Foam/Epoxy Composites. ACS Nano 2014, 8, 5774–5783. [Google Scholar] [CrossRef]
  93. Choi, B.; Yang, M.; Hong, W.; Choi, J.; Huh, Y. 3D Macroporous Graphene Frameworks for Supercapacitors with High Energy and Power Densities. ACS Nano 2012, 6, 4020–4028. [Google Scholar] [CrossRef]
  94. Xu, S.; Sun, F.; Pan, Z.; Huang, C.; Yang, S.; Long, J.; Chen, Y. Reduced Graphene Oxide-Based Ordered Macroporous Films on a Curved Surface: General Fabrication and Application in Gas Sensors. ACS Appl. Mater. Interfaces 2016, 8, 3428–3437. [Google Scholar] [CrossRef]
  95. Salazar-Aguilar, A.D.; Rodriguez Rodriguez, J.I.; Piñeiro-García, A.; Tristan, F.; Labrada Delgado, G.J.; Meneses-Rodríguez, D.; Vega Díaz, S.M. Layer-by-Layer Method to Prepare Three-Dimensional Reduced Graphene Materials with Controlled Architectures Using SiO2 as a Sacrificial Template. Ind. Eng. Chem. Res. 2021, 60, 11063–11069. [Google Scholar] [CrossRef]
  96. Lai, K.C.; Lee, L.Y.; Hiew, B.Y.Z.; Thangalazhy-Gopakumar, S.; Gan, S. Environmental application of three-dimensional graphene materials as adsorbents for dyes and heavy metals: Review on ice-templating method and adsorption mechanisms. J. Environ. Sci. 2019, 79, 174–199. [Google Scholar] [CrossRef]
  97. Zong, Y.; Fan, Y.; Chen, Z.; Ouyang, X.; Liu, Y.; Zhou, S.; Mei, J.; Niu, K. Control of microstructure to prepare compressible graphene aerogel via ice template. Carbon 2024, 227, 119269. [Google Scholar] [CrossRef]
  98. Qiu, L.; Han, W.; Yu, Q.; Yin, Y.; Yi, L.; Ji, X.; Yu, Y. Radial assembly of graphene nanoplates via ice-template method to construct 3D thermally conductive phase change materials. Fuel 2024, 369, 131738. [Google Scholar] [CrossRef]
  99. Kota, M.; Yu, X.; Yeon, S.H.; Cheong, H.W.; Park, H.S. Ice-templated three dimensional nitrogen doped graphene for enhanced supercapacitor performance. J. Power Sources 2016, 303, 372–378. [Google Scholar] [CrossRef]
  100. Peng, L.; Domenech Carbo, A.; Primo, A.; Garcia, H. 3D defective graphenes with subnanometric porosity obtained by soft-templating following zeolite procedures. Nanoscale Adv. 2019, 1, 4827–4833. [Google Scholar] [CrossRef]
  101. Huang, X.; Sun, B.; Su, D.; Zhao, D.; Wang, G. Soft-template synthesis of 3D porous graphene foams with tunable architectures for lithium-O2 batteries and oil adsorption applications. J. Mater. Chem. A 2014, 2, 7973–7979. [Google Scholar] [CrossRef]
  102. Li, Y.; Yang, J.; Huang, J.; Zhou, Y.; Xu, K.; Zhao, N.; Cheng, X. Soft template-assisted method for synthesis of nitrogen and sulfur co-doped three-dimensional reduced graphene oxide as an efficient metal free catalyst for oxygen reduction reaction. Carbon 2017, 122, 237–246. [Google Scholar] [CrossRef]
  103. Xu, Y.; Shi, G.; Duan, X. Self-Assembled Three-Dimensional Graphene Macrostructures: Synthesis and Applications in Supercapacitors. Acc. Chem. Res. 2015, 48, 1666–1675. [Google Scholar] [CrossRef]
  104. Xu, Y.; Sheng, K.; Li, C.; Shi, G. Self-Assembled Graphene Hydrogel via a One-Step Hydrothermal Process. ACS Nano 2010, 4, 4324–4330. [Google Scholar] [CrossRef] [PubMed]
  105. Sun, Y.; Wu, Q.; Shi, G. Supercapacitors based on self-assembled graphene organogel. Phys. Chem. Chem. Phys. 2011, 13, 17249–17254. [Google Scholar] [CrossRef] [PubMed]
  106. Zhao, L.; Wang, Z.-B.; Li, J.-L.; Zhang, J.-J.; Sui, X.-L.; Zhang, L.-M. Hybrid of carbon-supported Pt nanoparticles and three dimensional graphene aerogel as high stable electrocatalyst for methanol electrooxidation. Electrochim. Acta 2016, 189, 175–183. [Google Scholar] [CrossRef]
  107. Zhao, Z.; Zhang, J.; Wang, Y.; Guo, W.; Lv, L.; Li, B.; Zheng, K.; Bo, T.; Peng, D.; Zhou, Z.; et al. Thermally stable Pd/reduced graphene oxide aerogel catalysts for solvent-free oxidation of benzyl alcohol. Phys. Chem. Chem. Phys. 2011, 13, 17249–17254. [Google Scholar] [CrossRef]
  108. Kwok, Y.H.; Tsang, A.C.H.; Wang, Y.; Leung, D.Y.C. Ultra-fine Pt nanoparticles on graphene aerogel as a porous electrode with high stability for microfluidic methanol fuel cell. J. Power Sources 2017, 349, 75–83. [Google Scholar] [CrossRef]
  109. Liu, W.; Cai, J.; Li, Z. Self-Assembly of Semiconductor Nanoparticles/Reduced Graphene Oxide (RGO) Composite Aerogels for Enhanced Photocatalytic Performance and Facile Recycling in Aqueous Photocatalysis. ACS Sustain. Chem. Eng. 2015, 3, 277–282. [Google Scholar] [CrossRef]
  110. Das, C.; Shafi, T.; Thulluru, L.P.; Naushad, M.; Dubey, B.K.; Chowdhury, S. Three-Dimensional Highly Porous MoS2/Graphene Aerogel for Visible-Light-Driven Photocatalytic Degradation of Tetracycline. ACS EST Water 2024, 4, 2144–2158. [Google Scholar] [CrossRef]
  111. Zhu, K.; Yan, W.; Liu, S.; Wu, X.; Cui, S.; Shen, X. One-step hydrothermal synthesis of MnOx-CeO2/reduced graphene oxide composite aerogels for low temperature selective catalytic reduction of NOx. Appl. Surf. Sci. 2020, 508, 145024. [Google Scholar] [CrossRef]
  112. Liu, Y.; He, D.; Wu, H.; Duan, J.; Zhang, Y. Hydrothermal Self-assembly of Manganese Dioxide/Manganese Carbonate/Reduced Graphene Oxide Aerogel for Asymmetric Supercapacitors. Electrochim. Acta 2015, 164, 154–162. [Google Scholar] [CrossRef]
  113. Kim, M.Y.; Seo, K.D.; Park, H.; Mahmudunnabi, R.G.; Hwan Lee, K.; Shim, Y.B. Graphene-anchored conductive polymer aerogel composite for the electrocatalytic detection of hydrogen peroxide and bisphenol A. Appl. Surf. Sci. 2022, 604, 154430. [Google Scholar] [CrossRef]
  114. Li, C.; Yang, Z.; Tang, Z.; Guo, B.; Tian, M.; Zhang, L. A scalable strategy for constructing three-dimensional segregated graphene network in polymer via hydrothermal self-assembly. Chem. Eng. J. 2019, 363, 300–308. [Google Scholar] [CrossRef]
  115. Narukulla, R.; Ojha, U.; Sharma, T. Facile one pot green synthesis of -NH2 surface functionalized graphene-polymer nanocomposite: Subsequent utilization as stabilizer in pickering emulsions. Colloids Surf. A 2022, 641, 128594. [Google Scholar] [CrossRef]
  116. Wang, Z.; Wei, R.; Gu, J.; Liu, H.; Liu, C.; Luo, C.; Kong, J.; Shao, Q.; Wang, N.; Guo, Z.; et al. Ultralight, highly compressible and fire-retardant graphene aerogel with self-adjustable electromagnetic wave absorption. Carbon 2018, 139, 1126–1135. [Google Scholar] [CrossRef]
  117. Chen, W.; Yan, L. In situ self-assembly of mild chemical reduction graphene for three-dimensional architectures. Nanoscale 2011, 3, 3132–3137. [Google Scholar] [CrossRef]
  118. Wu, X.; Mu, F.; Lin, Z. Three-dimensional printing of graphene-based materials and the application in energy storage. Mater. Today Adv. 2021, 11, 100157. [Google Scholar] [CrossRef]
  119. Li, D.; Zhang, H.; Zhang, L.; Wang, P.; Xu, H.; Xuan, J. Rapid Synthesis of Porous Graphene Microspheres through a Three-Dimensionally Printed Inkjet Nozzle for Selective Pollutant Removal from Water. ACS Omega 2019, 4, 20509–20518. [Google Scholar] [CrossRef] [PubMed]
  120. Sun, J.; Sun, Y.; Jia, H.; Bi, H.; Chen, L.; Que, M.; Xiong, Y.; Han, L.; Sun, L. A novel pre-deposition assisted strategy for inkjet printing graphene-based flexible pressure sensor with enhanced performance. Carbon 2022, 196, 85–91. [Google Scholar] [CrossRef]
  121. You, X.; Yang, J.; Feng, Q.; Huang, K.; Zhou, H.; Hu, J.; Dong, S. Three-dimensional graphene-based materials by direct ink writing method for lightweight application. IJLLM 2018, 1, 96–101. [Google Scholar] [CrossRef]
  122. Li, Q.; Dong, Q.; Wang, J.; Xue, Z.; Li, J.; Yu, M.; Zhang, T.; Wan, Y.; Sun, H. Direct ink writing (DIW) of graphene aerogel composite electrode for vanadium redox flow battery. J. Power Sources 2022, 542, 231810. [Google Scholar] [CrossRef]
  123. van Hazendonk, L.S.; Vonk, C.F.; van Grondelle, W.; Vonk, N.H.; Friedrich, H. Towards a predictive understanding of direct ink writing of graphene-based inks. Appl. Mater. Today 2024, 36, 102014. [Google Scholar] [CrossRef]
  124. Wajahat, M.; Lee, S.; Kim, J.H.; Ahn, J.; Sim, H.H.; Kim, J.H.; Bae, J.; Kim, S.H.; Pyo, J.; Seol, S.K. Three-dimensional printing of silver nanoparticle-decorated graphene microarchitectures. Addit. Manuf. 2022, 60, 103249. [Google Scholar] [CrossRef]
  125. Chi, H.; Lin, Z.; Chen, Y.; Zheng, R.; Qiu, H.; Hu, X.; Bai, H. Three-Dimensional Printing and Recycling of Multifunctional Composite Material Based on Commercial Epoxy Resin and Graphene Nanoplatelet. ACS Appl. Mater. Interfaces 2022, 14, 13758–13767. [Google Scholar] [CrossRef] [PubMed]
  126. Li, L.; Chao, B.; Wu, H.; Liu, S.; Fu, W.; Jiang, J.; Deng, K.; Li, Y. Preparation of GR-TiN/PLA/TPU composite materials and study on their microwave absorption properties. Mater. Sci. Eng. B 2024, 299, 116964. [Google Scholar] [CrossRef]
  127. Jing, J.; Xiong, Y.; Shi, S.; Pei, H.; Chen, Y.; Lambin, P. Facile fabrication of lightweight porous FDM-Printed polyethylene/graphene nanocomposites with enhanced interfacial strength for electromagnetic interference shielding. Compos. Sci. Technol. 2021, 207, 108732. [Google Scholar] [CrossRef]
  128. Jing, J.; Chen, Y.; Shi, S.; Yang, L.; Lambin, P. Facile and scalable fabrication of highly thermal conductive polyethylene/graphene nanocomposites by combining solid-state shear milling and FDM 3D-printing aligning methods. Chem. Eng. J. 2020, 402, 126218. [Google Scholar] [CrossRef]
  129. Markandan, K.; Lai, C.Q. Enhanced mechanical properties of 3D printed graphene-polymer composite lattices at very low graphene concentrations. Compos. Part A-Appl. Sci. Manuf. 2020, 129, 105726. [Google Scholar] [CrossRef]
  130. Korhonen, H.; Sinh, L.H.; Luong, N.D.; Lehtinen, P.; Verho, T.; Partanen, J.; Seppälä, J. Fabrication of graphene-based 3D structures by stereolithography. Phys. Status Solidi (A) 2016, 213, 982–985. [Google Scholar] [CrossRef]
  131. Hanon, M.M.; Ghaly, A.; Zsidai, L.; Klébert, S. Tribological characteristics of digital light processing (DLP) 3D printed graphene/resin composite: Influence of graphene presence and process settings. Mater. Design 2022, 218, 110718. [Google Scholar] [CrossRef]
  132. Peng, M.; Wen, Z.; Xie, L.; Cheng, J.; Jia, Z.; Shi, D.; Zeng, H.; Zhao, B.; Liang, Z.; Li, T.; et al. 3D Printing of Ultralight Biomimetic Hierarchical Graphene Materials with Exceptional Stiffness and Resilience. Adv. Mater. 2019, 31, 1902930. [Google Scholar] [CrossRef]
  133. Hu, C.; Chen, Z.; Tang, L.; Liu, J.; Yang, J.; Lai, W.-F.; Wu, T.; Liao, S.; Zhang, X.; Pan, H.; et al. A universally dispersible graphene-based ink modifier facilitates 3D printing of multi-functional tissue-engineered scaffolds. Mater. Design 2022, 216, 110551. [Google Scholar] [CrossRef]
  134. Tang, X.; Zhou, H.; Cai, Z.; Cheng, D.; He, P.; Xie, P.; Zhang, D.; Fan, T. Generalized 3D Printing of Graphene-Based Mixed-Dimensional Hybrid Aerogels. ACS Nano 2018, 12, 3502–3511. [Google Scholar] [CrossRef]
  135. Li, Y.H.; Tang, Z.R.; Xu, Y.J. Multifunctional graphene-based composite photocatalysts oriented by multifaced roles of graphene in photocatalysis. Chin. J. Catal. 2022, 43, 708–730. [Google Scholar] [CrossRef]
  136. Lu, K.Q.; Xin, X.; Zhang, N.; Tang, Z.R.; Xu, Y.J. Photoredox catalysis over graphene aerogel-supported composites. J. Mater. Chem. A 2018, 6, 4590–4604. [Google Scholar] [CrossRef]
  137. Liu, S.; Jiang, T.; Fan, M.; Tan, G.; Cui, S.; Shen, X. Nanostructure rod-like TiO2-reduced graphene oxide composite aerogels for highly-efficient visible-light photocatalytic CO2 reduction. J. Alloys Compd. 2021, 861, 158598. [Google Scholar] [CrossRef]
  138. Huang, Z.; Wang, J.; Lu, S.; Xue, H.; Chen, Q.; Yang, M.Q.; Qian, Q. Insight into the Real Efficacy of Graphene for Enhancing Photocatalytic Efficiency: A Case Study on CVD Graphene-TiO2 Composites. ACS Appl. Energy Mater. 2021, 4, 8755–8764. [Google Scholar] [CrossRef]
  139. Nawaz, M.; Miran, W.; Jang, J.; Lee, D.S. One-step hydrothermal synthesis of porous 3D reduced graphene oxide/TiO2 aerogel for carbamazepine photodegradation in aqueous solution. Appl. Catal. B Environ. 2017, 203, 85–95. [Google Scholar] [CrossRef]
  140. Liu, H.; Li, X.; Ma, L.; Sun, F.; Yue, B.; Ma, Q.; Wang, J.; Liu, G.; Yu, H.; Yu, W.; et al. A three-dimensional TiO2/C/BiOBr graphene aerogel for enhancing photocatalysis and bidirectional sulfur conversion reactions in lithium-sulfur batteries. J. Environ. Chem. Eng. 2022, 10, 108606. [Google Scholar] [CrossRef]
  141. 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]
  142. Men, X.; Chen, H.; Chang, K.; Fang, X.; Wu, C.; Qin, W.; Yin, S. Three-dimensional free-standing ZnO/graphene composite foam for photocurrent generation and photocatalytic activity. Appl. Catal. B Environ. 2016, 187, 367–374. [Google Scholar] [CrossRef]
  143. Mishra, M.; Chun, D.M. α-Fe2O3 as a photocatalytic material: A review. Appl. Catal. A Gen. 2015, 498, 126–141. [Google Scholar] [CrossRef]
  144. Wang, Y.; Zhang, Y.; Wang, S.; Guan, Y.; Zhang, Y. A 3D printed synergistic aerogel microreactor toward stable and high-efficiency photocatalytic degradation. Mater. Today Chem. 2021, 22, 100566. [Google Scholar] [CrossRef]
  145. Singha, A.; Kaishyop, J.; Khan, T.S.; Chowdhury, B. Visible-Light-Driven Toluene Oxidation to Benzaldehyde over WO3 Nanostructures. ACS Appl. Nano Mater. 2023, 6, 21818–21828. [Google Scholar] [CrossRef]
  146. Lei, Y.; Deng, Y.; Luo, L.; Bao, S.; Tang, Z.; Chen, J.; Wang, Y.; Li, C.; Du, B. In Situ Synthesis of Mixed-Phase WO3 on Nitrogen-Doped Graphene with Enhanced Photoelectric Properties. ACS Appl. Nano Mater. 2023, 6, 16805–16814. [Google Scholar] [CrossRef]
  147. Wang, Z.; Zhu, C.; Ni, Z.; Hojo, H.; Einaga, H. Enhanced Photocatalytic Benzene Oxidation to Phenol over Monoclinic WO3 Nanorods under Visible Light. ACS Catal. 2022, 12, 14976–14989. [Google Scholar] [CrossRef]
  148. Azimirad, R.; Safa, S. Preparation of three dimensional graphene foam-WO3 nanocomposite with enhanced visible light photocatalytic activity. Mater. Chem. Phys. 2015, 162, 686–691. [Google Scholar] [CrossRef]
  149. Li, Y.; Tang, Z.; Zhang, J.; Zhang, Z. Reduced graphene oxide three-dimensionally wrapped WO3 hierarchical nanostructures as high-performance solar photocatalytic materials. Appl. Catal. A Gen. 2016, 522, 90–100. [Google Scholar] [CrossRef]
  150. Zhang, D.; Hu, B.; Guan, D.; Luo, Z. Essential roles of defects in pure graphene/Cu2O photocatalyst. Catal. Commun. 2016, 76, 7–12. [Google Scholar] [CrossRef]
  151. Zhang, S.; Li, X.; Bai, X.; Zhao, W.; Zhao, S.; Chang, C.; Wang, S.; Duan, E. Facet-Engineered Cu2O Microparticles on Graphitic Carbon Nitride Nanosheets as Z-Scheme Heterojunction Photocatalysts for Degradation of Tetracycline. ACS Appl. Nano Mater. 2024, 7, 16127–16140. [Google Scholar] [CrossRef]
  152. Ding, S.; Bai, X.; Cui, L.; Shen, Q.; Zhang, X.; Jia, H.; Xue, J. Ag as an Electron Mediator in Porous Cu2O Nanostructures for Photocatalytic CO2 Reduction to CO. ACS Appl. Nano Mater. 2023, 6, 10539–10550. [Google Scholar] [CrossRef]
  153. Cai, J.; Liu, W.; Li, Z. One-pot self-assembly of Cu2O/RGO composite aerogel for aqueous photocatalysis. Appl. Surf. Sci. 2015, 358, 146–151. [Google Scholar] [CrossRef]
  154. Zhang, S.; Ou, X.; Xiang, Q.; Carabineiro, S.A.C.; Fan, J.; Lv, K. Research progress in metal sulfides for photocatalysis: From activity to stability. Chemosphere 2022, 303 Pt 2, 135085. [Google Scholar] [CrossRef]
  155. Di, J.; Jiang, W. Recent progress of low-dimensional metal sulfides photocatalysts for energy and environmental applications. Mater. Today Catal. 2023, 1, 100001. [Google Scholar] [CrossRef]
  156. Parihar, A.; Sharma, P.; Choudhary, N.K.; Khan, R.; Gupta, A.; Sen, R.K.; Prasad, H.C.; Ashiq, M. Green Synthesis of CdS and CdS/rGO Nanocomposites: Evaluation of Electrochemical, Antimicrobial, and Photocatalytic Properties. ACS Appl. Bio Mater. 2023, 6, 3706–3716. [Google Scholar] [CrossRef] [PubMed]
  157. Das, C.; Shafi, T.; Pan, S.; Naushad, M.; Dubey, B.K.; Chowdhury, S. MoS2 Nanoflowers Decorated on Graphene Aerogels for Visible-Light-Driven Photocatalytic Degradation of Tetracycline. ACS Appl. Nano Mater. 2023, 6, 12991–13000. [Google Scholar] [CrossRef]
  158. Wei, X.N.; Ou, C.L.; Fang, S.S.; Zheng, X.C.; Zheng, G.P.; Guan, X.X. One-pot self-assembly of 3D CdS-graphene aerogels with superior adsorption capacity and photocatalytic activity for water purification. Powder Technol. 2019, 345, 213–222. [Google Scholar] [CrossRef]
  159. Bano, Z.; Saeed, R.M.Y.; Zhu, S.; Xia, M.; Mao, S.; Lei, W.; Wang, F. Mesoporous CuS nanospheres decorated rGO aerogel for high photocatalytic activity towards Cr(VI) and organic pollutants. Chemosphere 2020, 246, 125846. [Google Scholar] [CrossRef]
  160. Jiang, Y.; Chowdhury, S.; Balasubramanian, R. Nitrogen-doped graphene hydrogels as potential adsorbents and photocatalysts for environmental remediation. Chem. Eng. J. 2017, 327, 751–763. [Google Scholar] [CrossRef]
  161. Shafi, T.; Das, C.; Dubey, B.K.; Chowdhury, S. Aerogels Made of Few-Layer WS2 Nanosheets and Nitrogen-Doped Graphene for Photocatalytic Degradation of Caffeine. ACS Appl. Nano Mater. 2024, 7, 1723–1737. [Google Scholar] [CrossRef]
  162. Zhang, L.; Wu, L.; Feng, Z.; Meng, Q.; Li, Y.; Duan, T. Adopting sulfur-atom sharing strategy to construct CoS2/MoS2 heterostructure on three-dimensional nitrogen-doped graphene aerogels: A novel photocatalyst for wastewater treatment. J. Environ. Chem. Eng. 2021, 9, 104771. [Google Scholar] [CrossRef]
  163. Wang, L.; Wang, K.; He, T.; Zhao, Y.; Song, H.; Wang, H. Graphitic Carbon Nitride-Based Photocatalytic Materials: Preparation Strategy and Application. ACS Sustain. Chem. Eng. 2020, 8, 16048–16085. [Google Scholar] [CrossRef]
  164. El Messaoudi, N.; Ciğeroğlu, Z.; Şenol, Z.M.; Elhajam, M.; Noureen, L. A comparative review of the adsorption and photocatalytic degradation of tetracycline in aquatic environment by g-C3N4-based materials. J. Water Process Eng. 2023, 55, 104150. [Google Scholar] [CrossRef]
  165. Zhang, J.Y.; Zhang, S.H.; Li, J.; Zheng, X.C.; Guan, X.X. Constructing of 3D graphene aerogel-g-C3N4 metal-free heterojunctions with superior purification efficiency for organic dyes. J. Mol. Liq. 2020, 310, 113242. [Google Scholar] [CrossRef]
  166. Zhang, R.; Huang, Z.; Li, C.; Zuo, Y.; Zhou, Y. Monolithic g-C3N4/reduced graphene oxide aerogel with in situ embedding of Pd nanoparticles for hydrogenation of CO2 to CH4. Appl. Surf. Sci. 2019, 475, 953–960. [Google Scholar] [CrossRef]
  167. Chen, Z.; Ma, S.; Hu, Y.; Lv, F.; Zhang, Y. Preparation of Bi-based porous and magnetic electrospun fibers and their photocatalytic properties in weak polar medium. Colloids Surf. A 2021, 610, 125718. [Google Scholar] [CrossRef]
  168. Zuarez Chamba, M.; Rajendran, S.; Herrera Robledo, M.; Priya, A.K.; Navas Cardenas, C. Bi-based photocatalysts for bacterial inactivation in water: Inactivation mechanisms, challenges, and strategies to improve the photocatalytic activity. Environ. Res. 2022, 209, 112834. [Google Scholar] [CrossRef]
  169. Yu, X.; Shi, J.; Feng, L.; Li, C.; Wang, L. A three-dimensional BiOBr/RGO heterostructural aerogel with enhanced and selective photocatalytic properties under visible light. Appl. Surf. Sci. 2017, 396, 1775–1782. [Google Scholar] [CrossRef]
  170. Yang, J.; Shi, Q.; Zhang, R.; Xie, M.; Jiang, X.; Wang, F.; Cheng, X.; Han, W. BiVO4 quantum tubes loaded on reduced graphene oxide aerogel as efficient photocatalyst for gaseous formaldehyde degradation. Carbon 2018, 138, 118–124. [Google Scholar] [CrossRef]
  171. Jafarzadeh, M. Recent Progress in the Development of MOF-Based Photocatalysts for the Photoreduction of Cr((VI)). ACS Appl. Mater. Interfaces 2022, 14, 24993–25024. [Google Scholar] [CrossRef] [PubMed]
  172. Zhang, Y.; Ma, D.; Li, J.; Zhi, C.; Zhang, Y.; Liang, L.; Mao, S.; Shi, J.W. Recent research advances of metal organic frameworks (MOFs) based composites for photocatalytic H2 evolution. Coord. Chem. Rev. 2024, 517, 215955. [Google Scholar] [CrossRef]
  173. Sandhu, Z.A.; Farwa, U.; Danish, M.; Raza, M.A.; Talib, A.; Amjad, H.; Riaz, R.; Al-Sehemi, A.G. Sustainability and photocatalytic performance of MOFs: Synthesis strategies and structural insights. J. Clean. Prod. 2024, 470, 143263. [Google Scholar] [CrossRef]
  174. Liu, Z.; Sun, D.; Wang, C.; You, B.; Li, B.; Han, J.; Jiang, S.; Zhang, C.; He, S. Zeolitic imidazolate framework-67 and its derivatives for photocatalytic applications. Coord. Chem. Rev. 2024, 502, 215612. [Google Scholar] [CrossRef]
  175. Nazir, M.A.; Javed, M.S.; Islam, M.; Assiri, M.A.; Hassan, A.M.; Jamshaid, M.; Najam, T.; Shah, S.S.A.; Rehman, A.U. MOF@graphene nanocomposites for energy and environment applications. Compos. Commun. 2024, 45, 101783. [Google Scholar] [CrossRef]
  176. Zhao, X.; Xu, M.; Song, X.; Zhou, W.; Liu, X.; Wang, H.; Huo, P. Integration of 3D macroscopic reduced graphene oxide aerogel with DUT-67 for selective CO2 photoreduction to CO in Gas-Solid reaction. Chem. Eng. J. 2022, 446, 137034. [Google Scholar] [CrossRef]
  177. Shah, S.J.; Luan, X.; Yu, X.; Su, W.; Wang, Y.; Zhao, Z.; Zhao, Z. Construction of 3D-graphene/NH(2)-MIL-125 nanohybrids via amino-ionic liquid dual-mode bonding for advanced acetaldehyde photodegradation under high humidity. J. Colloid Interface Sci. 2024, 663, 491–507. [Google Scholar] [CrossRef] [PubMed]
  178. Yang, J.; Chen, D.; Zhu, Y.; Zhang, Y.; Zhu, Y. 3D-3D porous Bi2WO6/graphene hydrogel composite with excellent synergistic effect of adsorption-enrichment and photocatalytic degradation. Appl. Catal. B Environ. 2017, 205, 228–237. [Google Scholar] [CrossRef]
  179. Jin, S.; Yang, Y.; Zhang, J.; Zheng, H. Preparation of a novel TiO2-graphene 3D framework material for efficient adsorption-photocatalytic removal of micro-organic contaminants from water. Mater. Chem. Phys. 2021, 263, 124339. [Google Scholar] [CrossRef]
  180. Dong, S.; Xia, L.; Chen, X.; Cui, L.; Zhu, W.; Lu, Z.; Sun, J.; Fan, M. Interfacial and electronic band structure optimization for the adsorption and visible-light photocatalytic activity of macroscopic ZnSnO3/graphene aerogel. Compos. Part B Eng. 2021, 215, 108765. [Google Scholar] [CrossRef]
  181. Liu, Q.; Shen, J.; Yang, X.; Zhang, T.; Tang, H. 3D reduced graphene oxide aerogel-mediated Z-scheme photocatalytic system for highly efficient solar-driven water oxidation and removal of antibiotics. Appl. Catal. B Environ. 2018, 232, 562–573. [Google Scholar] [CrossRef]
  182. Shen, Z.K.; Yuan, Y.J.; Pei, L.; Yu, Z.T.; Zou, Z. Black phosphorus photocatalysts for photocatalytic H2 generation: A review. Chem. Eng. J. 2020, 386, 123997. [Google Scholar] [CrossRef]
  183. Ganguly, P.; Harb, M.; Cao, Z.; Cavallo, L.; Breen, A.; Dervin, S.; Dionysiou, D.D.; Pillai, S.C. 2D Nanomaterials for Photocatalytic Hydrogen Production. ACS Energy Lett. 2019, 4, 1687–1709. [Google Scholar] [CrossRef]
  184. Samajdar, S.; Ghosh, S.; Thandavarayan, M.; Medda, S.K.; Manna, S.; Mohapatra, M. Construction of a 3D/2D Z-Scheme Heterojunction for Promoting Charge Separation and Augmented Photocatalytic Hydrogen Evolution. Energy Fuels 2023, 37, 14290–14302. [Google Scholar] [CrossRef]
  185. Zhang, X.; An, W.; Li, Y.; Hu, J.; Gao, H.; Cui, W. Efficient photo-catalytic hydrogen production performance and stability of a three-dimensional porous CdS NPs-graphene hydrogel. Int. J. Hydrogen Energy 2018, 43, 9902–9913. [Google Scholar] [CrossRef]
  186. Liu, J.; Wei, X.; Sun, W.; Guan, X.; Zheng, X.; Li, J. Fabrication of S-scheme CdS-g-C3N4-graphene aerogel heterojunction for enhanced visible light driven photocatalysis. Environ. Res. 2021, 197, 111136. [Google Scholar] [CrossRef]
  187. Seo, D.B.; Kwon, Y.M.; Kim, J.; Kang, S.; Yim, S.; Lee, S.S.; Kim, E.T.; Song, W.; An, K.S. Edge-Rich 3D Structuring of Metal Chalcogenide/Graphene with Vertical Nanosheets for Efficient Photocatalytic Hydrogen Production. ACS Appl. Mater. Interfaces 2024, 16, 28613–28624. [Google Scholar] [CrossRef]
  188. Guo, K.; Hussain, I.; Jie, G.A.; Fu, Y.; Zhang, F.; Zhu, W. Strategies for improving the photocatalytic performance of metal-organic frameworks for CO(2) reduction: A review. J. Environ. Sci. 2023, 125, 290–308. [Google Scholar] [CrossRef]
  189. Alkhatib, I.I.; Garlisi, C.; Pagliaro, M.; Al-Ali, K.; Palmisano, G. Metal-organic frameworks for photocatalytic CO2 reduction under visible radiation: A review of strategies and applications. Catal. Today 2020, 340, 209–224. [Google Scholar] [CrossRef]
  190. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photoelectrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  191. Hussain, S.; Wang, Y.; Guo, L.; He, T. Theoretical insights into the mechanism of photocatalytic reduction of CO2 over semiconductor catalysts. J. Photoch. Photobio. C 2022, 52, 100538. [Google Scholar] [CrossRef]
  192. Hussain, S.; Yang, X.; Yang, J.; Li, Q. Theoretical insights into the mechanism of photocatalytic reduction of CO2 and water splitting over II-VI zinc chalcogenide semiconductor. Mater. Today Sustain. 2024, 25, 100686. [Google Scholar] [CrossRef]
  193. Yuan, Z.; Zhu, X.; Gao, X.; An, C.; Wang, Z.; Zuo, C.; Dionysiou, D.D.; He, H.; Jiang, Z. Enhancing photocatalytic CO2 reduction with TiO(2)-based materials: Strategies, mechanisms, challenges, and perspectives. Environ. Sci. Ecotechnol. 2024, 20, 100368. [Google Scholar] [CrossRef] [PubMed]
  194. Park, J.; Jin, T.; Liu, C.; Li, G.; Yan, M. Three-Dimensional Graphene-TiO2 Nanocomposite Photocatalyst Synthesized by Covalent Attachment. ACS Omega 2016, 1, 351–356. [Google Scholar] [CrossRef] [PubMed]
  195. Zhang, J.; Shao, S.; Zhou, D.; Xu, Q.; Wang, T. ZnO nanowire arrays decorated 3D N-doped reduced graphene oxide nanotube framework for enhanced photocatalytic CO2 reduction performance. J. CO2 Util. 2021, 50, 101584. [Google Scholar] [CrossRef]
  196. Song, M.; Li, J.; Xu, M.; Xu, Z.; Song, X.; Liu, X.; Zhang, J.; Yang, Y.; Xie, X.; Zhou, W.; et al. Facile synthesis of MOF-808/RGO-based 3D macroscopic aerogel for enhanced photoreduction CO2. J. Colloid Interface Sci. 2024, 668, 471–483. [Google Scholar] [CrossRef] [PubMed]
  197. Xia, Y.; Cheng, B.; Fan, J.; Yu, J.; Liu, G. Near-infrared absorbing 2D/3D ZnIn2S4/N-doped graphene photocatalyst for highly efficient CO2 capture and photocatalytic reduction. Sci. China Mater. 2020, 63, 552–565. [Google Scholar] [CrossRef]
Figure 1. Scheme diagram of the photocatalytic process.
Figure 1. Scheme diagram of the photocatalytic process.
Gels 10 00626 g001
Figure 2. Systematic diagram of graphene and graphene-based derivatives. Reproduced with permission from Ref. [61], Copyright 2022, Elsevier.
Figure 2. Systematic diagram of graphene and graphene-based derivatives. Reproduced with permission from Ref. [61], Copyright 2022, Elsevier.
Gels 10 00626 g002
Figure 3. Various roles of 3D graphene in the photocatalytic system.
Figure 3. Various roles of 3D graphene in the photocatalytic system.
Gels 10 00626 g003
Figure 4. 3D graphene with different structures. (A) Fabrication process (i) and morphology (iiv) of graphene aerogel spheres. Reproduced with permission from Ref. [84], Copyright 2020, Elsevier; (B) Fabrication (i) and structural characteristics (iiv) of graphene foam. Reproduced with permission from Ref. [85], Copyright 2023, Elsevier; (C) Schematic illustration for the preparation of graphene aerogel. Reproduced with permission from Ref. [86], Copyright 2024, Elsevier; (D) Digital camera images (i,ii) and cross-sectional view SEM images (iii,iv) of graphene aerogel film. Reproduced with permission from Ref. [87], Copyright 2018, Elsevier; (E) Cross polarized-light optical images (i), SEM images (ii,iii), and photographs (iv,v) of graphene aerogel hollow fiber. Reproduced with permission from Ref. [88], Copyright 2022, Elsevier.
Figure 4. 3D graphene with different structures. (A) Fabrication process (i) and morphology (iiv) of graphene aerogel spheres. Reproduced with permission from Ref. [84], Copyright 2020, Elsevier; (B) Fabrication (i) and structural characteristics (iiv) of graphene foam. Reproduced with permission from Ref. [85], Copyright 2023, Elsevier; (C) Schematic illustration for the preparation of graphene aerogel. Reproduced with permission from Ref. [86], Copyright 2024, Elsevier; (D) Digital camera images (i,ii) and cross-sectional view SEM images (iii,iv) of graphene aerogel film. Reproduced with permission from Ref. [87], Copyright 2018, Elsevier; (E) Cross polarized-light optical images (i), SEM images (ii,iii), and photographs (iv,v) of graphene aerogel hollow fiber. Reproduced with permission from Ref. [88], Copyright 2022, Elsevier.
Gels 10 00626 g004
Figure 5. (A) Schematic diagram of the preparation process of 3D graphene by CVD (iiv); (B) SEM images of copper template (i), graphene grown on copper template (ii), and graphene network after evaporating copper template (iii,iv). Reproduced with permission form Ref. [91], Copyright 2017, American Chemical Society.
Figure 5. (A) Schematic diagram of the preparation process of 3D graphene by CVD (iiv); (B) SEM images of copper template (i), graphene grown on copper template (ii), and graphene network after evaporating copper template (iii,iv). Reproduced with permission form Ref. [91], Copyright 2017, American Chemical Society.
Gels 10 00626 g005
Figure 6. (A) Schematic of a ceramic tube (i) and SEM image of polystyrene microspheres wrapped with graphene oxide (ii); (B) Fabrication process of graphene MOP film; (C) Morphologies of graphene-oxide composite films: SEM images of graphene-SnO2 film (i,ii), TEM images of graphene-SnO2 film (iii,iv), and SEM images of graphene-Fe2O3 (v) and graphene-NiO (vi). Reproduced with permission from Ref. [94], Copyright 2016, American Chemical Society.
Figure 6. (A) Schematic of a ceramic tube (i) and SEM image of polystyrene microspheres wrapped with graphene oxide (ii); (B) Fabrication process of graphene MOP film; (C) Morphologies of graphene-oxide composite films: SEM images of graphene-SnO2 film (i,ii), TEM images of graphene-SnO2 film (iii,iv), and SEM images of graphene-Fe2O3 (v) and graphene-NiO (vi). Reproduced with permission from Ref. [94], Copyright 2016, American Chemical Society.
Gels 10 00626 g006
Figure 7. (A) The procedure of preparing graphene aerogels by a combination of hydrothermal treatment (different time), lyophilization and hydrazine reduction: 20 min (i), 30 min (ii), and 40 min (iii); (B) The cross-sectional SEM images of graphene aerogels with different hydrothermal times and their corresponding magnified views: 20 min (iiii), 30 min (ivvi), 40 min (viiix). Reproduced with permission from Ref. [97], Copyright 2024, Elsevier.
Figure 7. (A) The procedure of preparing graphene aerogels by a combination of hydrothermal treatment (different time), lyophilization and hydrazine reduction: 20 min (i), 30 min (ii), and 40 min (iii); (B) The cross-sectional SEM images of graphene aerogels with different hydrothermal times and their corresponding magnified views: 20 min (iiii), 30 min (ivvi), 40 min (viiix). Reproduced with permission from Ref. [97], Copyright 2024, Elsevier.
Gels 10 00626 g007
Figure 8. (A) Fabrication procedure of the phase change materials with radial scaffold; (B) Morphology of the aerogel and phase change material: schematic diagram for cross-sectional SEM observation (i) and cross-section SEM (ii,iii) of graphene/chitosan aerogel, schematic diagram for longitudinal section SEM observation (iv) and longitudinal section SEM (v,vi) of graphene/chitosan aerogel, digital image (vii) and SEM images of graphene/chitosan-PCM (viii,ix). Reproduced with permission from Ref. [98], Copyright 2024, Elsevier.
Figure 8. (A) Fabrication procedure of the phase change materials with radial scaffold; (B) Morphology of the aerogel and phase change material: schematic diagram for cross-sectional SEM observation (i) and cross-section SEM (ii,iii) of graphene/chitosan aerogel, schematic diagram for longitudinal section SEM observation (iv) and longitudinal section SEM (v,vi) of graphene/chitosan aerogel, digital image (vii) and SEM images of graphene/chitosan-PCM (viii,ix). Reproduced with permission from Ref. [98], Copyright 2024, Elsevier.
Gels 10 00626 g008
Figure 9. (A) Schematic illustration of the emulsion soft-template synthesis procedures for preparing porous graphene foams; (B) SEM images (i,ii) and TEM images (iiivi) of graphene using TMB as emulsion templates; (C) SEM images (i,ii) and TEM images (iiivi) of graphene using n-hexadecane as emulsion templates. Reproduced with permission from Ref. [101], Copyright 2014, Royal Society of Chemistry.
Figure 9. (A) Schematic illustration of the emulsion soft-template synthesis procedures for preparing porous graphene foams; (B) SEM images (i,ii) and TEM images (iiivi) of graphene using TMB as emulsion templates; (C) SEM images (i,ii) and TEM images (iiivi) of graphene using n-hexadecane as emulsion templates. Reproduced with permission from Ref. [101], Copyright 2014, Royal Society of Chemistry.
Gels 10 00626 g009
Figure 10. (A) The formation mechanism for graphene hydrogel; (B) Photographs of graphene oxide solution before and after hydrothermal reduction; (C) SEM image of the interior microstructures of graphene. Reproduced with permission from Ref. [104], Copyright 2010, American Chemical Society.
Figure 10. (A) The formation mechanism for graphene hydrogel; (B) Photographs of graphene oxide solution before and after hydrothermal reduction; (C) SEM image of the interior microstructures of graphene. Reproduced with permission from Ref. [104], Copyright 2010, American Chemical Society.
Gels 10 00626 g010
Figure 11. Schematic illustration of 3D printing (AC); SEM images (DF), ultralight structure (G), and mechanical properties (H,I) of graphene. Reproduced with permission from Ref. [132], Copyright 2019, Wiley.
Figure 11. Schematic illustration of 3D printing (AC); SEM images (DF), ultralight structure (G), and mechanical properties (H,I) of graphene. Reproduced with permission from Ref. [132], Copyright 2019, Wiley.
Gels 10 00626 g011
Figure 12. (A) Preparation schematic diagram of ZnO/graphene foam, inset is the corresponding sample photographs of each step; (B) SEM images of graphene/Ni (i), graphene (ii), and ZnO/graphene (iiivi). Reproduced with permission from Ref. [142], Copyright 2016, Elsevier.
Figure 12. (A) Preparation schematic diagram of ZnO/graphene foam, inset is the corresponding sample photographs of each step; (B) SEM images of graphene/Ni (i), graphene (ii), and ZnO/graphene (iiivi). Reproduced with permission from Ref. [142], Copyright 2016, Elsevier.
Gels 10 00626 g012
Figure 13. (A) Schematic illustration of the synthetic process of Bi2WO6/graphene; (B) SEM images of Bi2WO6 (i) and Bi2WO6/graphene (ii); (C) Schematic diagram of pollutants adsorption and photocatalytic degradation by Bi2WO6/graphene composite. Reproduced with permission from Ref. [178], Copyright 2017, Elsevier.
Figure 13. (A) Schematic illustration of the synthetic process of Bi2WO6/graphene; (B) SEM images of Bi2WO6 (i) and Bi2WO6/graphene (ii); (C) Schematic diagram of pollutants adsorption and photocatalytic degradation by Bi2WO6/graphene composite. Reproduced with permission from Ref. [178], Copyright 2017, Elsevier.
Gels 10 00626 g013
Figure 14. (A) SEM images of MoS2/graphene with different synthetic process: hydrothermal method (i) and chemical activation route (ii); (B) Schematic of the photocatalysis mechanism under visible−light irradiation; (C) Probable pathways for photocatalytic degradation of tetracycline. Reproduced with permission from Ref. [110], Copyright 2024, American Chemical Society.
Figure 14. (A) SEM images of MoS2/graphene with different synthetic process: hydrothermal method (i) and chemical activation route (ii); (B) Schematic of the photocatalysis mechanism under visible−light irradiation; (C) Probable pathways for photocatalytic degradation of tetracycline. Reproduced with permission from Ref. [110], Copyright 2024, American Chemical Society.
Gels 10 00626 g014
Figure 15. (A) A SEM image of 3D graphene (i), TEM images of TiO2/graphene (ii,iv), and the lattice diffraction pattern of the particle in the white circle in (iii); (B) Electron transfer pathways in the photocatalytic hydrogen production by TiO2/graphene. Reproduced with permission from Ref. [66], Copyright 2017, Springer.
Figure 15. (A) A SEM image of 3D graphene (i), TEM images of TiO2/graphene (ii,iv), and the lattice diffraction pattern of the particle in the white circle in (iii); (B) Electron transfer pathways in the photocatalytic hydrogen production by TiO2/graphene. Reproduced with permission from Ref. [66], Copyright 2017, Springer.
Gels 10 00626 g015
Figure 16. (A) Fabrication process of (MoS2 and WS2) nanosheets/3D graphene; (B) SEM images of 3D graphene (i), MoS2/graphene (ii), and WS2/graphene (iii); (C) Schematic illustration of the transfer process in the photocatalyst system (i) and the photocatalytic performance comparison of this work with other photocatalysts in the literature (ii). Reproduced with permission from Ref. [187], Copyright 2024, American Chemical Society.
Figure 16. (A) Fabrication process of (MoS2 and WS2) nanosheets/3D graphene; (B) SEM images of 3D graphene (i), MoS2/graphene (ii), and WS2/graphene (iii); (C) Schematic illustration of the transfer process in the photocatalyst system (i) and the photocatalytic performance comparison of this work with other photocatalysts in the literature (ii). Reproduced with permission from Ref. [187], Copyright 2024, American Chemical Society.
Gels 10 00626 g016
Figure 17. (A) Schematic illustration for the formation process of ZnIn2S4/N-graphene; (B) SEM images of N-graphene (i) and ZnIn2S4/N-graphene (ii); (C) Photogenerated charge transfer mechanism: valence band spectra (i), flat-band potentials (ii), and energy level diagram (iii). Reproduced with permission from Ref. [197], Copyright 2020, Springer.
Figure 17. (A) Schematic illustration for the formation process of ZnIn2S4/N-graphene; (B) SEM images of N-graphene (i) and ZnIn2S4/N-graphene (ii); (C) Photogenerated charge transfer mechanism: valence band spectra (i), flat-band potentials (ii), and energy level diagram (iii). Reproduced with permission from Ref. [197], Copyright 2020, Springer.
Gels 10 00626 g017
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, F.; Liu, J.; Hu, L.; Guo, C. Recent Progress of Three-Dimensional Graphene-Based Composites for Photocatalysis. Gels 2024, 10, 626. https://doi.org/10.3390/gels10100626

AMA Style

Zhang F, Liu J, Hu L, Guo C. Recent Progress of Three-Dimensional Graphene-Based Composites for Photocatalysis. Gels. 2024; 10(10):626. https://doi.org/10.3390/gels10100626

Chicago/Turabian Style

Zhang, Fengling, Jianxing Liu, Liang Hu, and Cean Guo. 2024. "Recent Progress of Three-Dimensional Graphene-Based Composites for Photocatalysis" Gels 10, no. 10: 626. https://doi.org/10.3390/gels10100626

APA Style

Zhang, F., Liu, J., Hu, L., & Guo, C. (2024). Recent Progress of Three-Dimensional Graphene-Based Composites for Photocatalysis. Gels, 10(10), 626. https://doi.org/10.3390/gels10100626

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