Next Article in Journal
Active Actuating of a Simply Supported Beam with the Flexoelectric Effect
Previous Article in Journal
Design of CNS-Li2SiO3 Permeable Protective Coatings and Effects on Mortar Matrix
Previous Article in Special Issue
Silver-Nanoparticles Embedded Pyridine-Cholesterol Xerogels as Highly Efficient Catalysts for 4-nitrophenol Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbon Gels-Modified TiO2: Promising Materials for Photocatalysis Applications

1
School of Science, Beijing Technology and Business University, Beijing 100048, China
2
Basic Experimental Center for Natural Science, University of Science and Technology Beijing, Beijing 100083, China
3
Key Laboratory of Photochemistry, Beijing National Laboratory for Molecular Sciences, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China
*
Author to whom correspondence should be addressed.
Submission received: 25 March 2020 / Revised: 2 April 2020 / Accepted: 2 April 2020 / Published: 8 April 2020
(This article belongs to the Special Issue Organic and Carbon Gels: Properties and Application)

Abstract

:
Carbon gels are a kind of porous organic polymer, which play pivotal roles in electrode, supercapacitor, hydrogen storage, and catalysis. Carbon gels are commonly prepared by the condensation of resorcinol and formaldehyde. The as-prepared polymers are further aged and sintered at a high temperature in an inert atmosphere to form cross-linked and intertwined porous structures. Owing to its large specific area and narrow pore size distribution, this kind of material is very appropriate for mass transfer, substrate absorption, and product desorption from the pores. In recent years, carbon gels have been discovered to function as effective hybrid materials with TiO2 for photocatalytic applications. They could act as efficient deep-traps for photo-induced holes, which decreases the recombination probability of photo-induced carriers and lengthens their lifetime. In this mini-review, we will discuss the state-of-the-art paragon examples of carbon gels/TiO2 composite materials applied in photo(electro)catalysis. The major challenges and gaps of its application in this field will also be emphasized.

1. Introduction

Nowadays, environmental pollution and energy shortage are two main crises facing the whole world. The excessive utilization of fossil fuels produces various pollutants, which was discharged into water and air environments. There is an urgent need for the use of green and renewable energy resources for human society. Sunlight, considered as an inexhaustible and environmentally friendly energy source, has aroused the attention from both academics and industries. As Fujishima and Honda first reported TiO2 could be used to catalyze the water-splitting process evolving hydrogen and oxygen under sunlight illumination in 1972 [1], and Carey et al. initially used TiO2 photocatalysis as an advanced oxidation technique for the removal of pollutants in aqueous solution [2,3], photocatalysis has been developed rapidly and becomes the focus of various disciplines including catalysis [4,5,6,7], materials chemistry [8,9,10], environmental chemistry [11], energy chemistry [12,13], surface chemistry [14,15,16], and processing chemistry and chemical engineering [17,18]. Among various photocatalysts, TiO2 being extremely stable under light irradiation, highly acidic, and basic conditions; non-toxic, earth-abundant, and easily-recyclable; and reusable without much loss of activity, has been profoundly investigated from the 1970s up till now [19,20,21,22,23,24,25,26,27]. Moreover, with the high photo-induced hole oxidation potential (Evb+ = 2.7 V vs. NHE (Normal Hydrogen Electrode) at pH = 7); the ability to produce similarly highly oxidative OH• radicals (E = 2.8 V vs. NHE), H2O2, and HOO•; and the appropriate photo-induced conduction band electron reduction potential (E = −0.5 V vs. NHE at pH = 7), which could be easily trapped by dioxygen generating superoxide radical anion, all of these reactive oxygen species, along with the photo-induced holes on TiO2 surface, could almost decompose all the organic pollutants by unselective thorough mineralization with sequential hydrogen abstraction; halogen abstraction; and addition to R•, C=C, and C≡C bonds until all the organic pollutants finally transform to CO2, H2O, and inorganic ion species [7,11,28]. Although TiO2 photocatalysis has garnered a plethora of successful examples in water environment remediation applications [29,30,31,32,33], there are still gaps and shortcomings that need to be overcome for TiO2 photocatalysis to be applied in real industrial processes. The following main issues need to be addressed. Firstly, owing to the wide band gap (Eg = 3.2 eV for anatase and 3.0 eV for rutile), TiO2-based nanomaterial could only be excited under UV (ultra-violet) irradiation (λ < 387.5 nm). This means that only less than 5% of the solar spectrum energy could be utilized by TiO2 photocatalysis. Approximately 95% of the solar energy is wasted as heat. Furthermore, current TiO2 photocatalysis displays a low quantum yield, even in UV spectrum. Poor photon energy to chemical energy efficiency is obtained for the existing TiO2 system. The quantum yield is in the range below 20%. The poor photocatalysis performance of TiO2 is attributed to the following factors: (1) poor visible-light absorption, (2) facile hole-electron recombination, and (3) poor adsorption and catalytic activity towards non-polar and hydrophobic compounds. To overcome these obstacles, various strategies have been developed. Metal and non-metal atoms [34] doping such as Fe [35], Cu [36], Ni [37] and N [38,39,40], C [41], and S [42], which could introduce extra traps below the bottom of the conduction band and the top of the valence band, were incorporated in lowering the band gap to facilitate visible-light absorption ability [43]. However, the addition of these dopant atoms could in another way decrease its photocatalytic efficiency, as the trapping sites would also be the recombination sites for electrons and holes. Coupling other photocatalysts with TiO2 to construct heterojunction structure is another approach to improve TiO2 photocatalysis performance. A number of inorganic and organic semiconductor photocatalysts, including CdS [44], BiVO4 [45], AgBr [46], RGO [47], and g-C3N4 [48], have been combined with TiO2 to generate a hybrid photocatalyst system, which displays excellent performance for various photocatalytic applications including water-splitting, CO2 reduction, water and air decontamination, and organic synthesis [49,50,51,52].
Carbon gels, as a kind of aerogels, were discovered as early as 1931 by Kistler [53]. The author successfully prepared the gels using the supercritical liquid evaporation method, which guaranteed that the jelly internal structure did not experience either change or shrinkage. This is a milestone discovery for colloid chemistry. However, the synthesis and preparation of carbon gels was not rediscovered and researched until more than half a century later. Pekala and co-workers innovatively developed the four-step procedure to synthesize resorcinol–formaldehyde (RF) gels as shown in Figure 1 [54]. Firstly, resorcinol and formaldehyde were polycondensed in an alkaline solution. Then, the as-formed RF mixtures were subjected to aging conditions at 85 °C for several days. Later, the acid-treatment and solvent-exchange processes were required before the key critical-point drying procedure. Finally, CO2 was introduced to flow in and replace the residue solvent molecules at 45 °C. The dried RF gels were characterized by various techniques to prove the existence of the organic aerogel structure. This example resembles the renaissance of organic carbon gels [55,56,57,58,59]. Nowadays, carbon gels have been extensively investigated and applied in hydrogen and methane storage [60,61,62,63,64,65,66,67,68,69], electric energy storage [70,71,72,73], thermal transport [74,75], and catalysis [76,77,78]. Carbon gels have been demonstrated as ideal materials for these applications owing to their large specific surface area, high electric conductivity, and porous structure [79,80]. These features would also be critical merits for catalyst and co-catalyst materials for photocatalysis applications [81,82]. Carbon gels have been successfully incorporated into TiO2 photocatalysis to improve the overall performance for photocatalytic water-splitting and environment remediation [83]. This review will discuss the state-of-the-art paragon examples of carbon gels-modified TiO2-nanomaterials including the preparation, characterization, and activity of these composite photocatalysts. Furthermore, the gaps and challenges of this area will also be outlined.

2. Carbon Gels-Modified TiO2 Photocatalysis

Owing to their large specific surface area, ordered porous structure, and high electric conductivity, carbon gels have been applied in various fields including hydrogen storage, electrode materials, and heterogeneous catalyst support. Although carbon gels are often used as conductive materials, they also can be used to construct a composite semiconductor photocatalyst with TiO2.
In 2010, Zhao and co-workers firstly used carbon gels to modify the TiO2 photocatalyst to improve its adsorption and conductivity performance [84]. They prepared a TiO2/carbon aerogel (CA) photoelectrode material. The CA was synthesized by the base-catalyzed polycondensation between resorcinol and formaldehyde. The as-synthesized wet gels were initially solvent-exchanged by acetone to replace water. The organic gel was transformed to CA by heating in an argon atmosphere at 950 °C. The CA material was immersed into the sol–gel process of TiO2 formation. The as-formed mixed TiO2/CA was sintered at different high temperatures to yield the final photoelectrode material. Various characterization techniques such as scanning electronic microscopy (SEM), X-ray diffraction (XRD), Raman spectrometry, and N2 isothermal adsorption and desorption experiments were applied to determine the morphology (as shown in Figure 2), crystal structure, and porosity property. The TiO2/CA photoelectrode possesses both excellent photocatalytic properties owing to the anatase TiO2 semiconductor photoresponsive component and outstanding electrochemical properties of CA material, such as its extremely high conductivity, low electrochemical impedance, and large specific surface area. Thus, the TiO2/CA photoelectrode demonstrated much higher photocurrent density and degradation efficiency for highly opaque methylene blue wastewater under a −0.6 V potential bias and 365 nm light irradiation compared with the TiO2/ITO photoelectrode. The authors attributed the high efficiency to the following factors. Firstly, applying −0.6 V bias would form an electrical double layer between the electrode and bulk solution, causing the pollutants methylene blue, which possesses positive charges, to adsorb on TiO2/CA electrode surface. The excellent conductivity of CA materials would enhance the electrosorption process and the large specific surface area would promote the dispersion of TiO2 nano-crystallite on the electrode film. The electrosorption effect would increase the transparency of the opaque wastewater, which would greatly increase the photocatalytic degradation efficiency of TiO2.
In 2011, Wu and co-workers reported that mesoporous and macroporous carbon aerogels could couple with TiO2 to construct a hybrid CA/TiO2 composite photocatalyst [85]. The composite photocatalyst was prepared by a sol–gel process mixing Ti(iPrO)4 with CA in an alcoholic solution. The as-prepared CA was sintered at 400 °C under N2 atmosphere. Two CA materials with different pore diameters were hybridized with TiO2. The one with mesoporous structure displayed much higher activity on methyl orange degradation compared with pristine TiO2 and TiO2/CA with macroporous structure as shown in Figure 3. By analyzing XRD (X-ray diffraction), SEM (scanning electron microscopy), and BET (Brunner-Emmet-Teller measurements) experimental results, the authors determined that the mesoporous structure of TiO2/CA125 photocatalyst promoted the substrates adsorption and products desorption because of the suitable pore size. The micropores could not accommodate the considerably large methyl orange dye molecule entering the pore, while the macropore is too large to confine the dye molecule inside the pore for effective adsorption and surface photocatalytic reaction.
In 2011, Zhao and co-workers reported that CA/TiO2 hybrid material could act as a very promising candidate for a photocatalysis enhanced electroadsorption (PES)-based dye pollutant elimination process [86]. They prepared the CA/TiO2 electrode by dispersing the as-prepared CA material into titania tetrabutylate sol–gel. The CA material was synthesized from the poly-condensation of resorcinol and formaldehyde, the displacement of water by immersing in acetone, and final sintering under argon at a high temperature forming the final porous aerosol microstructure. The morphology, surface composition, and properties were characterized by SEM, EDS (Energy Dispersive Spectroscopy), TG/DTA (Thermogravimetric Analysis/Differential Thermal Analysis), BET, EIS (Electrochemical Impedance Spectroscopy), and Raman spectrometry. Furthermore, the material demonstrated excellent alizarin red (AR) dye pollutant elimination ability by combining adsorption, electroadsorption, and photocatalysis in a three-in-all strategy. In a 400 mg/L high concentration AR pollutant solution, the hybrid material exhibited 97.3% TOC (Total Organic Carbon) removal in 240 min for the PES process in comparison with 59% in the photocatalysis process without applied electric field and 22% in the electrosorption process without illumination. Not only did it possess high AR removal efficiency, but this material also exhibited very good stability, providing 92.3% AR removal after five cycles. Moreover, the authors proposed a mechanism for the synergistic effect of photocatalysis and electrosorption as shown in Figure 4. The adsorbed dye molecules labile bonds were initially cleaved via single-electron-transfer by TiO2 photo-induced hole species or hydrogen abstraction by hydroxyl radicals and other reactive oxygen species (ROS), and the dye molecule radical cations were hydroxylated and mineralized step by step via radical addition, atom abstraction, and bond cleavage mechanism. The photo-induced electrons were extracted to the other electrode by circuit to inhibit hole-electron recombination, and thus improve the mineralization efficiency. Owing to the excellent porosity and conductivity, CA could act as very effective adsorbent and electrode materials. Under applied positive potential, negative-charged dye-molecules would be attracted to the positive electrode, enhancing adsorption. The TiO2 photocatalyst mineralized the dye, which avoided saturated adsorption of CA material, while CA adsorbed and enriched dye pollutant to keep high pollutant concentration on the TiO2 surface to make the photocatalytic reaction proceed rapidly. Photocatalysis and electroadsorption synergistically ensured highly efficient high-concentration dye solution decontamination, which is important not only in academics, but also in industrial engineering.
Furthermore, Zhao and co-workers discovered that TiO2/CA material could have the capacity to act as an effective photoelectrode to extend the TiO2 absorption spectrum to the visible-light region and mineralize Rhodamine 6G dye pollutants under visible-light irradiation in an aerobic condition under −0.9 V bias [87]. They figured out that H2O2 was generated on the electrode surface by the reduction of dioxygen. Further, the in-situ generated H2O2 coordinated with TiO2, forming a visible-light responsive surface Ti–peroxide complex. Upon light irradiation, this surface complex was excited and injected electrons to the TiO2 conduction band. The conduction band electrons’ decomposed surface adsorbed H2O2 to hydroxyl radicals. The hydroxyl radicals that possess strong oxidative ability (E1/2 = 2.4 V vs. NHE) can be directly applied to mineralize pollutants. This mechanism was shown in Figure 5. To demonstrate the efficiency of this photoelectrocatalysis system, azo dye Rhodamine 6G was applied as a model pollutant in the activity test. The results showed that, upon –0.9 V bias and visible-light irradiation (λ > 420 nm), the TiO2/CA photocathode could reach removal efficiency for Rhodamine 6G at 90.3% and TOC at 83.3% in 300 min. GC-MS (Gas Chromatography- Mass Spectroscopy) and HPLC (High-performance Liquid Chromatography) tracing of intermediates indicated that less intermediates were generated under photoelectrocatalytic conditions than under traditional photocatalytic conditions. This catalyst system demonstrated excellent pollutant removal efficiency and low energy consumption. More importantly, this example provided us a new approach to explore the visible-light responsive unconventional TiO2-based photocatalytic materials.
Zhao and co-workers reported that TiO2 nanorod cluster (NRC) material could be loaded on carbon aerogel by a hydrothermal/calcination method [88]. The as-prepared TiO2 NRC/CA composite material could be used as photocathode for the decomposition and mineralization of a notorious pollutant p-nitroaniline. They compared the pollutant removal efficiency and activity of TiO2 NRC/CA with TiO2 NRC/graphite material. The results indicated that TiO2 NRC/CA material could provide superior 98.2% PNA (p-Nitroaniline) removal efficiency in 180 min under visible-light irradiation with −0.6 V bias in comparison with 79.8% for TiO2 NRC/graphite. From various characterization experiments, the authors ascribed the higher efficiency of TiO2 NRC/CA to the CA’s three-dimensional structure, which provided greater surface area and larger conductivity. More TiO2 photocatalyst could be loaded on CA rather than on graphite. Further, more H2O2 and hydroxyl radicals could be formed on electrode surface. Owing to the greater level of TiO2 loading, more surface Ti–peroxide complex and photocurrent could be yielded under visible-light irradiation. All these factors made TiO2 NRC/CA excellent photocathode materials for p-nitroaniline pollutants mineralization. Moreover, the authors proposed the decomposition pathways for this photoelectrocatalytic system as shown in Figure 6. They discovered the new important intermediates 1,2,4-trihydroxybenzene during the decomposition processes of p-nitroaniline and reasonably explained its mineralization mechanism.
Shi and co-workers discovered that TiO2 photocatalytic activity of DMP (dimethyl phthalate) degradation could be enhanced by hybridizing a CA material and constructing secondary hydrophilic pores [89]. They synthesized TiO2/CA composite photocatalysts with different CA pore diameters. They discovered that materials comprising CA with 9.3 nm mesopores exhibited the highest photocatalytic degradation efficiency, while CA with smaller pore diameters of 3.4 and 4.3 nm generated inferior results even compared with the pristine TiO2 photocatalyst. Contact angle analysis and DMP adsorption tests indicated that CA-9.3 possessed the least ability to adsorb DMP owing to its larger wettability and hydrophilicity. However, as TiO2 nanoparticles dispersed in both inner and outer space of CA-9.3, photo-induced electrons on the outer surface migrated to the inner surface through CA conductive porous structure, inhibiting electron-hole recombination and improving the charge separation process. Thus, the generation of hydroxyl radicals by photo-induced holes on the outer surface was enhanced by this effect, which increased the DMP photodegradation rate. Moreover, the generation rate of hydroxyl radical by other routes was also increased. In the inner space of hydrophilic secondary pore, dioxygen, photo-electrons, and water molecules were enriched by this micro-reactor. This encapsulation effect enhanced the hydroxyl radical generation by the dioxygen reduction route in comparison with the lack of porous structure of pristine TiO2. This mechanism was illustrated in Figure 7.
Wang and co-workers reported that Ce-doped TiO2/CA could be applied as potent visible-light responsive electrosorptive photocatalytic materials for 4-chlorophenol degradation [90]. They discovered that cerium doping could introduce defect sites above the top of the valence band red-shifting TiO2 absorption spectrum from 387 nm to 532 nm. Furthermore, carbon aerogel was hybridized with Ce-TiO2 photocatalyst by an impregnation method. The as-synthesized Ce-TiO2/CA material was used as photoelectrode to degrade 4-chlorophenol under 0.6 V bias and 500 W Xenon lamp irradiation. This material exhibited 75 times larger photocurrent than Ce-TiO2/FTO (fluorine-doped tin oxide), which was ascribed to the lesser electron-hole recombination rate by the CA porous structure, which facilitated the charge separation processes, as mentioned above. Furthermore, for the 4-chlorophenol (4-CP) degradation, Ce-TiO2/CA generated 97.3% 4-CP removal in 4.5 h, in sharp contrast with 65.4% for Ce-TiO2/FTO under other identical photoelectrochemical conditions. This enhancing effect was mainly because of the highly-developed porous structure and excellent conductivity, which were pivotal for 4-CP adsorption and the charge separation and migration processes. This mechanism was illustrated in Figure 8.
Szilágyi and co-workers reported an atomic-layer-deposition (ALD) method to prepare resorcinol–formaldehyde aerogel (RFA) and resorcinol–formaldehyde carbon aerogel (RFCA)/TiO2 composite photocatalysts [91]. They initially prepared RFA and RFCA by the modified Pekala’s three-step method; the first step to synthesize RF hydrogel, the second step dry RF hydrogel in supercritical condition forming RFA, and the last step to yield RFCA by sintering RFA under a N2 atmosphere at a high temperature. The as-synthesized RFA and RFCA were utilized as substrates for ALD of TiO2 at 80 °C and 250 °C. TiO2 prepared at 80 °C was amorphous, while that at 250 °C was crystalline. Furthermore, the authors conducted the photo-degradation experiments of methyl orange dye. Surprisingly, they discovered that RFCA/amorphous TiO2 exhibited better photocatalytic activity than RFCA/crystalline TiO2. Further, the RFCA material without TiO2 deposition exhibited the best photocatalytic activity towards methyl orange degradation. These results were shown in Figure 9. The authors ascribed these unconventional results to the fact that the oxide deposition reduced the surface area and the functional group content of RFCA, thus blocking and decreasing the activity sites. The ALD treatment at a higher temperature more significantly deteriorated the RFCA porous structure and decreased its functional group contents. This report was interesting, demonstrating that amorphous carbonaceous material such as resorcinol–formaldehyde carbon aerogel could not only exhibit semiconductive photocatalytic activity, but also generate higher ability towards organic pollutant degradation under illumination than traditional metal-oxide materials.
The performance of the state-of-the-art examples of TiO2/CA photo(electro)catalysis is summarised in Table 1.
Besides the activity improvements of the TiO2/CA photo(electro)catalyst, its life cycle assessment (LCA) of the environmental impacts should also be considered. Although we lack the data of the environmental effects of the CA synthesis and the hybrid process of TiO2 and CA nano-materials, according to the previously reported LCA of environmental impacts for seven different fabrication routes of TiO2 nanomaterials by mass unit, surface-area unit, and photocatalytic activity unit [92], or the reported LCA cradle to gate environmental impacts of five different non-doped and metal-doped TiO2 nanomaterials prepared by the sol–gel method by the function unit of photocatalytic carbamazepine and methyl orange degradation [93], we can infer that the TiO2/CA photo(electro)catalyst will generate more environmental impacts because the fabrication of CA materials requires an intensive energy-input process and the use of a considerable amount of organic precursor resorcinol and formaldehyde, while the use of formaldehyde will contribute to environmental concern. Further, during the photo(electro)catalytic process of pollutant removal, the added bias would generate much greater consumption of electricity energy. However, the greatly enhanced pollutants removal efficiency would come as a trade-off to the added extra environment impacts for TiO2/CA materials compared with non-doped TiO2 photocatalysts. Further, a thorough investigation using quantitative software to assess the environmental impacts of the TiO2/CA photo(electro)catalyst “from cradle to grave” based on the function unit of the photo(electro)catalytic removal of certain organic pollutants by LCA should be conducted in the future.

3. Conclusions

We have outlined the state-of-the-art examples of carbon aerogels materials hybridized with TiO2 applied as both catalysts for photoelectrochemical cells and photocatalysts for environmental pollutants degradation. Although still in its blossoming stage in comparison with other traditional inorganic metal oxides, metal chalcogenides, and surface plasmonic metallic photocatalysts and photo-electrocatalysts, CA-modified TiO2 material still exhibited miscellaneous advantages. Firstly, carbon aerogels modification could greatly increase the surface area of TiO2 owing to the great porosity and highly-developed hierarchical porous structures. The increase of surface area would enhance the adsorption of pollutants on TiO2. Furthermore, the introduction of porous structure would enhance electron-hole pair separation by the facilitated TiO2 photoelectron transfer between inner surface and outer surface in the interconnected pore structure. Secondly, owing to CA’s excellent conductivity, upon hybridization with CA, TiO2 semiconductive material became much more conductive, and the electron-transfer impedance was greatly reduced. This effect was crucial for the increase of photocurrent during the photoelectrochemical degradation of organic pollutants. Last, but not least, the doping of carbon aerogel would introduce defect sites in TiO2 bulk solid, thus generating a dopant energy level in the top of the valence band and in the bottom of the conduction band, thus providing the narrower band gap for more efficient photon absorption. This would generate an apparent red-shift phenomenon for TiO2, extending its absorption limit from 387 nm to approximately 532 nm. Compared with other crystalline carbonaceous materials, metal oxides, metal chalcogenides, and metal surface plasmonic photocatalysts, CA/TiO2 photocatalysts still have a long way to proceed for more mature applications in environment remediation. The main limitations and challenges for CA/TiO2 photocatalysts include the following aspects. Initially, the degradation efficiency should be further improved to fulfill the requirements of practical use in industry wastewater treatment. To meet this end, an elaborate, intricate, and novel synthesizing method should be developed. Much higher performance of hydroxyl radical generation and pollutants adsorption should be achieved. Secondly, in order to improve the activity of the composite photocatalyst, the mechanism of how the photoelectrons, photo-induced holes, dioxygen, water, organic pollutants, and the degradation intermediates interacted in the composite photocatalyst, and in particular, how the porous structure influenced the whole photocatalytic process should be more clearly demonstrated. We do believe that if these issues can be resolved successfully, the future of CA/TiO2 photocatalysis would be promising.

Author Contributions

Conceptualization, D.M.; writing—original draft preparation, D.M.; writing—review and editing, D.M., A.L. and C.C.; Literature search: D.M. and J.L. supervision, D.M.; funding acquisition, D.M. and A.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (21703005) and the fundamental research funds for the central universities (FRF-TP-18-104A1).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  2. Carey, J.H.; Lawrence, J.; Tosine, H.M. Photodechlorination of PCB’s in the presence of titanium dioxide in aqueous suspensions. Bull. Environ. Contam. Toxicol. 1976, 16, 697–701. [Google Scholar] [CrossRef] [PubMed]
  3. Frank, S.N.; Bard, A.J. Heterogeneous photocatalytic oxidation of cyanide ion in aqueous solutions at titanium dioxide powder. J. Am. Chem. Soc. 1977, 99, 303–304. [Google Scholar] [CrossRef]
  4. 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]
  5. Ishitani, O.; Inoue, C.; Suzuki, Y. Photocatalytic reduction of carbon dioxide to methane and acetic acid by an aqueous suspension of metal-deposited TiO2. J. Photochem. Photobiol. A Chem. 1993, 72, 269–271. [Google Scholar] [CrossRef]
  6. Yu, J.; Low, J.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced Photocatalytic CO2-Reduction Activity of Anatase TiO2 by Coexposed {001} and {101} Facets. J. Am. Chem. Soc. 2014, 136, 8839–8842. [Google Scholar] [CrossRef]
  7. Ma, D.; Liu, A.; Li, S.; Lu, C.; Chen, C. TiO2 photocatalysis for C–C bond formation. Catal. Sci. Technol. 2018, 8, 2030–2045. [Google Scholar] [CrossRef]
  8. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications, and Applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef]
  9. Zhang, H.; Liu, G.; Shi, L.; Liu, H.; Wang, T.; Ye, J. Engineering coordination polymers for photocatalysis. Nano Energy 2016, 22, 149–168. [Google Scholar] [CrossRef]
  10. Xiong, F.; Yu, Y.-Y.; Wu, Z.; Sun, G.; Ding, L.; Jin, Y.; Gong, X.-Q.; Huang, W. Methanol Conversion into Dimethyl Ether on the Anatase TiO2(001) Surface. Angew. Chem. Int. Ed. 2015, 55, 623–628. [Google Scholar] [CrossRef]
  11. Chen, C.; Ma, W.; Zhao, J. Semiconductor-mediated photodegradation of pollutants under visible-light irradiation. Chem. Soc. Rev. 2010, 39, 4206. [Google Scholar] [CrossRef] [PubMed]
  12. Li, H.; Shang, J.; Ai, Z.; Zhang, L. Efficient Visible Light Nitrogen Fixation with BiOBr Nanosheets of Oxygen Vacancies on the Exposed {001} Facets. J. Am. Chem. Soc. 2015, 137, 6393–6399. [Google Scholar] [CrossRef] [PubMed]
  13. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for Selective Photoreduction of CO2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef] [PubMed]
  14. Heller, A. Chemistry and Applications of Photocatalytic Oxidation of Thin Organic Films. Accounts Chem. Res. 1995, 28, 503–508. [Google Scholar] [CrossRef]
  15. Kikuchi, Y.; Sunada, K.; Iyoda, T.; Hashimoto, K.; Fujishima, A. Photocatalytic bactericidal effect of TiO2 thin films: dynamic view of the active oxygen species responsible for the effect. J. Photochem. Photobiol. A Chem. 1997, 106, 51–56. [Google Scholar] [CrossRef]
  16. Saito, T.; Iwase, T.; Horie, J.; Morioka, T. Mode of photocatalytic bactericidal action of powdered semiconductor TiO2 on mutans streptococci. J. Photochem. Photobiol. B Boil. 1992, 14, 369–379. [Google Scholar] [CrossRef]
  17. Bet-Moushoul, E.; Mansourpanah, Y.; Farhadi, K.; Tabatabaei, M. TiO2 nanocomposite based polymeric membranes: A review on performance improvement for various applications in chemical engineering processes. Chem. Eng. J. 2016, 283, 29–46. [Google Scholar] [CrossRef]
  18. Sheikh, M.; Pazirofteh, M.; Dehghani, M.; Asghari, M.; Rezakazemi, M.; Valderrama, C.; Cortina, J. Application of ZnO nanostructures in ceramic and polymeric membranes for water and wastewater technologies: A review. Chem. Eng. J. 2019, 123475, 123475. [Google Scholar] [CrossRef]
  19. Fujishima, A.; Zhang, X.; Tryk, D. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  20. Ma, D.; Zhai, S.; Wang, Y.; Liu, A.; Chen, C. TiO2 Photocatalysis for Transfer Hydrogenation. Molecules 2019, 24, 330. [Google Scholar] [CrossRef] [Green Version]
  21. Ma, D.; Zhai, S.; Wang, Y.; Liu, A.; Chen, C. Synthetic Approaches for C-N Bonds by TiO2 Photocatalysis. Front. Chem. 2019, 7, 635. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Y.; Liu, A.; Ma, D.; Li, S.; Lu, C.; Li, T.; Chen, C. TiO2 Photocatalyzed C–H Bond Transformation for C–C Coupling Reactions. Catalysts 2018, 8, 355. [Google Scholar] [CrossRef] [Green Version]
  23. Lang, X.; Chen, X.; Zhao, J. Heterogeneous visible light photocatalysis for selective organic transformations. Chem. Soc. Rev. 2014, 43, 473–486. [Google Scholar] [CrossRef] [PubMed]
  24. Lang, X.; Ma, W.; Chen, C.; Ji, H.; Zhao, J. Selective Aerobic Oxidation Mediated by TiO2 Photocatalysis. Accounts Chem. Res. 2013, 47, 355–363. [Google Scholar] [CrossRef]
  25. Lang, X.; Zhao, J.; Chen, X. Cooperative photoredox catalysis. Chem. Soc. Rev. 2016, 45, 3026–3038. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, Y.; Wei, Y.; Song, W.; Chen, C.; Zhao, J. Photocatalytic Hydrodehalogenation for the Removal of Halogenated Aromatic Contaminants. ChemCatChem 2018, 11, 258–268. [Google Scholar] [CrossRef]
  27. Wu, T.; Liu, G.; Zhao, J.; Hidaka, H.; Serpone, N. Mechanistic study of the TiO2-assisted photodegradation of squarylium cyanine dye in methanolic suspensions exposed to visible light. New J. Chem. 2000, 24, 93–98. [Google Scholar] [CrossRef]
  28. Teoh, W.Y.; Scott, J.A.; Amal, R. Progress in Heterogeneous Photocatalysis: From Classical Radical Chemistry to Engineering Nanomaterials and Solar Reactors. J. Phys. Chem. Lett. 2012, 3, 629–639. [Google Scholar] [CrossRef]
  29. Tan, T.T.Y.; Beydoun, D.; Amal, R. Photocatalytic Reduction of Se(VI) in Aqueous Solutions in UV/TiO2 System: Kinetic Modeling and Reaction Mechanism. J. Phys. Chem. B 2003, 107, 4296–4303. [Google Scholar] [CrossRef]
  30. Augugliaro, V.; Camera-Roda, G.; Loddo, V.; Palmisano, G.; Palmisano, L.; Soria, J.; Yurdakal, S. Heterogeneous Photocatalysis and Photoelectrocatalysis: From Unselective Abatement of Noxious Species to Selective Production of High-Value Chemicals. J. Phys. Chem. Lett. 2015, 6, 1968–1981. [Google Scholar] [CrossRef]
  31. Liu, G.; Wu, T.; Zhao, J.; Hidaka, H.; Serpone, N. Photoassisted Degradation of Dye Pollutants. 8. Irreversible Degradation of Alizarin Red under Visible Light Radiation in Air-Equilibrated Aqueous TiO2Dispersions. Environ. Sci. Technol. 1999, 33, 2081–2087. [Google Scholar] [CrossRef]
  32. Serpone, N.; Emeline, A. Semiconductor Photocatalysis — Past, Present, and Future Outlook. J. Phys. Chem. Lett. 2012, 3, 673–677. [Google Scholar] [CrossRef] [PubMed]
  33. Wu, T.; Lin, T.; Zhao, J.; Hidaka, H.; Serpone, N. TiO2-Assisted Photodegradation of Dyes. 9. Photooxidation of a Squarylium Cyanine Dye in Aqueous Dispersions under Visible Light Irradiation. Environ. Sci. Technol. 1999, 33, 1379–1387. [Google Scholar] [CrossRef]
  34. Choi, W.; Termin, A.; Hoffmann, M.R. The Role of Metal Ion Dopants in Quantum-Sized TiO2: Correlation between Photoreactivity and Charge Carrier Recombination Dynamics. J. Phys. Chem. 1994, 98, 13669–13679. [Google Scholar] [CrossRef]
  35. Zhu, J.; Chen, F.; Zhang, J.; Chen, H.; Anpo, M. Fe3+-TiO2 photocatalysts prepared by combining sol–gel method with hydrothermal treatment and their characterization. J. Photochem. Photobiol. A Chem. 2006, 180, 196–204. [Google Scholar] [CrossRef]
  36. Valero, J.M.; Obregón, S.; Colón, G. Active Site Considerations on the Photocatalytic H2 Evolution Performance of Cu-Doped TiO2 Obtained by Different Doping Methods. ACS Catal. 2014, 4, 3320–3329. [Google Scholar] [CrossRef]
  37. Tang, J.; Grampp, G.; Liu, Y.; Wang, B.-X.; Tao, F.-F.; Wang, L.-J.; Liang, X.-Z.; Xiao, H.-Q.; Shen, Y.-M. Visible Light Mediated Cyclization of Tertiary Anilines with Maleimides Using Nickel(II) Oxide Surface-Modified Titanium Dioxide Catalyst. J. Org. Chem. 2015, 80, 2724–2732. [Google Scholar] [CrossRef]
  38. Xing, M.; Zhang, J.; Chen, F. New approaches to prepare nitrogen-doped TiO2 photocatalysts and study on their photocatalytic activities in visible light. Appl. Catal. B Environ. 2009, 89, 563–569. [Google Scholar] [CrossRef]
  39. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  40. Khan, S.U.; Al-Shahry, M.; Ingler, W.B. Efficient Photochemical Water Splitting by a Chemically Modified n-TiO2. Science 2002, 297, 2243–2245. [Google Scholar] [CrossRef]
  41. Shim, J.; Seo, Y.-S.; Oh, B.-T.; Cho, M. Microbial inactivation kinetics and mechanisms of carbon-doped TiO2 (C-TiO2) under visible light. J. Hazard. Mater. 2016, 306, 133–139. [Google Scholar] [CrossRef] [PubMed]
  42. Umebayashi, T.; Yamaki, T.; Itoh, H.; Asai, K. Band gap narrowing of titanium dioxide by sulfur doping. Appl. Phys. Lett. 2002, 81, 454–456. [Google Scholar] [CrossRef]
  43. Daghrir, R.; Drogui, P.; Robert, D. Modified TiO2 For Environmental Photocatalytic Applications: A Review. Ind. Eng. Chem. Res. 2013, 52, 3581–3599. [Google Scholar] [CrossRef]
  44. Liu, S.; Zhang, N.; Tang, Z.-R.; Xu, Y.-J. Synthesis of One-Dimensional CdS@TiO2 Core–Shell Nanocomposites Photocatalyst for Selective Redox: The Dual Role of TiO2 Shell. ACS Appl. Mater. Interfaces 2012, 4, 6378–6385. [Google Scholar] [CrossRef] [PubMed]
  45. Singh, A.P.; Kodan, N.; Mehta, B.R.; Held, A.; Mayrhofer, L.; Moseler, M. Band Edge Engineering in BiVO4/TiO2 Heterostructure: Enhanced Photoelectrochemical Performance through Improved Charge Transfer. ACS Catal. 2016, 6, 5311–5318. [Google Scholar] [CrossRef]
  46. Hirakawa, T.; Kamat, P.V. Photoinduced Electron Storage and Surface Plasmon Modulation in Ag@TiO2 Clusters. Langmuir 2004, 20, 5645–5647. [Google Scholar] [CrossRef]
  47. Yu, H.; Xiao, P.; Tian, J.; Wang, F.; Yu, J. Phenylamine-Functionalized rGO/TiO2 Photocatalysts: Spatially Separated Adsorption Sites and Tunable Photocatalytic Selectivity. ACS Appl. Mater. Interfaces 2016, 8, 29470–29477. [Google Scholar] [CrossRef]
  48. Wang, X.-J.; Yang, W.-Y.; Li, F.-T.; Xue, Y.-B.; Liu, R.-H.; Hao, Y.-J. In Situ Microwave-Assisted Synthesis of Porous N-TiO2/g-C3N4 Heterojunctions with Enhanced Visible-Light Photocatalytic Properties. Ind. Eng. Chem. Res. 2013, 52, 17140–17150. [Google Scholar] [CrossRef]
  49. Ma, D.; Liu, A.; Lu, C.; Chen, C. Photocatalytic Dehydrogenation of Primary Alcohols: Selectivity Goes against Adsorptivity. ACS Omega 2017, 2, 4161–4172. [Google Scholar] [CrossRef] [Green Version]
  50. Ma, D.; Yan, Y.; Ji, H.; Chen, C.; Zhao, J. Photocatalytic activation of pyridine for addition reactions: an unconventional reaction feature between a photo-induced hole and electron on TiO2. Chem. Commun. 2015, 51, 17451–17454. [Google Scholar] [CrossRef]
  51. Liu, Y.; Zhang, M.; Tung, C.-H.; Wang, Y. TiO2 Photocatalytic Cyclization Reactions for the Syntheses of Aryltetralones. ACS Catal. 2016, 6, 8389–8394. [Google Scholar] [CrossRef]
  52. Cherevatskaya, M.; Neumann, M.; Füldner, S.; Harlander, C.; Kümmel, S.; Dankesreiter, S.; Pfitzner, A.; Zeitler, K.; König, B. Visible-Light-Promoted Stereoselective Alkylation by Combining Heterogeneous Photocatalysis with Organocatalysis. Angew. Chem. Int. Ed. 2012, 51, 4062–4066. [Google Scholar] [CrossRef] [PubMed]
  53. Kistler, S.S. Coherent Expanded Aerogels and Jellies. Nature 1931, 127, 741. [Google Scholar] [CrossRef]
  54. Pekala, R.W. Organic aerogels from the polycondensation of resorcinol with formaldehyde. J. Mater. Sci. 1989, 24, 3221–3227. [Google Scholar] [CrossRef]
  55. Hanzawa, Y.; Kaneko, K.; Pekala, R.W.; Dresselhaus, M.S. Activated Carbon Aerogels. Langmuir 1996, 12, 6167–6169. [Google Scholar] [CrossRef]
  56. Heinrich, T.; Klett, U.; Fricke, J. Aerogels?Nanoporous materials part I: Sol-gel process and drying of gels. J. Porous Mater. 1995, 1, 7–17. [Google Scholar] [CrossRef]
  57. Lee, K.T.; Oh, S.M. Novel synthesis of porous carbons with tunable pore size by surfactant-templated sol-gel process and carbonisation. Chem. Commun. 2002, 2722–2723. [Google Scholar] [CrossRef]
  58. Baumann, T.F.; Satcher, J.H. Homogeneous Incorporation of Metal Nanoparticles into Ordered Macroporous Carbons. Chem. Mater. 2003, 15, 3745–3747. [Google Scholar] [CrossRef]
  59. Baumann, T.F.; Satcher, J.H. Template-directed synthesis of periodic macroporous organic and carbon aerogels. J. Non-Cryst. Solids 2004, 350, 120–125. [Google Scholar] [CrossRef]
  60. Feaver, A.; Cao, G. Activated carbon cryogels for low pressure methane storage. Carbon 2006, 44, 590–593. [Google Scholar] [CrossRef]
  61. Zhang, S.; Gross, A.F.; Van Atta, S.L.; López, M.; Liu, P.; Ahn, C.C.; Vajo, J.J.; Jensen, C.M. The synthesis and hydrogen storage properties of a MgH2 incorporated carbon aerogel scaffold. Nanotechnology 2009, 20, 204027. [Google Scholar] [CrossRef] [PubMed]
  62. Lin, K.-S.; Mai, Y.-J.; Chiu, S.-W.; Yang, J.-H.; Chan, S.L.I. Synthesis and Characterization of Metal Hydride/Carbon Aerogel Composites for Hydrogen Storage. J. Nanomater. 2012, 2012, 1–9. [Google Scholar] [CrossRef]
  63. Utke, R.; Milanese, C.; Javadian, P.; Jepsen, J.; Laipple, D.; Karmi, F.; Puszkiel, J.; Jensen, T.R.; Marini, A.; Klassen, T.; et al. Nanoconfined 2LiBH4–MgH2–TiCl3 in carbon aerogel scaffold for reversible hydrogen storage. Int. J. Hydrog. Energy 2013, 38, 3275–3282. [Google Scholar] [CrossRef] [Green Version]
  64. Gosalawit−Utke, R.; Nielsen, T.K.; Pranzas, K.; Saldan, I.; Pistidda, C.; Karimi, F.; Laipple, D.; Skibsted, J.; Jensen, T.R.; Klassen, T.; et al. 2LiBH4–MgH2 in a Resorcinol–Furfural Carbon Aerogel Scaffold for Reversible Hydrogen Storage. J. Phys. Chem. C 2011, 116, 1526–1534. [Google Scholar] [CrossRef]
  65. Tian, H.; Buckley, C.E.; Wang, S.; Zhou, M. Enhanced hydrogen storage capacity in carbon aerogels treated with KOH. Carbon 2009, 47, 2128–2130. [Google Scholar] [CrossRef]
  66. Gross, A.F.; Vajo, J.J.; Van Atta, S.L.; Olson, G.L. Enhanced Hydrogen Storage Kinetics of LiBH4 in Nanoporous Carbon Scaffolds. J. Phys. Chem. C 2008, 112, 5651–5657. [Google Scholar] [CrossRef]
  67. Feaver, A.; Sepehri, S.; Shamberger, P.J.; Stowe, A.; Autrey, T.; Cao, G. Coherent Carbon Cryogel−Ammonia Borane Nanocomposites for H2 Storage. J. Phys. Chem. B 2007, 111, 7469–7472. [Google Scholar] [CrossRef]
  68. McNicholas, T.P.; Wang, A.; O’Neill, K.; Anderson, R.; Stadie, N.; Kleinhammes, A.; Parilla, P.; Simpson, L.; Ahn, C.C.; Wang, Y.; et al. H2 Storage in Microporous Carbons from PEEK Precursors. J. Phys. Chem. C 2010, 114, 13902–13908. [Google Scholar] [CrossRef] [Green Version]
  69. Kabbour, H.; Baumann, T.F.; Satcher, J.H.; Saulnier, A.; Ahn, C.C. Toward New Candidates for Hydrogen Storage: High-Surface-Area Carbon Aerogels. Chem. Mater. 2006, 18, 6085–6087. [Google Scholar] [CrossRef]
  70. Mayer, S.T.; Pekala, R.; Kaschmitter, J.L. The Aerocapacitor: An Electrochemical Double?Layer Energy?Storage Device. J. Electrochem. Soc. 1993, 140, 446. [Google Scholar] [CrossRef] [Green Version]
  71. Saliger, R.; Fischer, U.; Herta, C.; Fricke, J. High surface area carbon aerogels for supercapacitors. J. Non-Cryst. Solids 1998, 225, 81–85. [Google Scholar] [CrossRef]
  72. Long, J.W.; Dening, B.M.; McEvoy, T.M.; Rolison, D.R. Carbon aerogels with ultrathin, electroactive poly(o-methoxyaniline) coatings for high-performance electrochemical capacitors. J. Non-Cryst. Solids 2004, 350, 97–106. [Google Scholar] [CrossRef]
  73. Bordjiba, T.; Mohamedi, M.; Dao, L.H. Synthesis and electrochemical capacitance of binderless nanocomposite electrodes formed by dispersion of carbon nanotubes and carbon aerogels. J. Power Sources 2007, 172, 991–998. [Google Scholar] [CrossRef]
  74. Worsley, M.A.; Satcher, J.H.; Baumann, T.F. Synthesis and Characterization of Monolithic Carbon Aerogel Nanocomposites Containing Double-Walled Carbon Nanotubes. Langmuir 2008, 24, 9763–9766. [Google Scholar] [CrossRef]
  75. Worsley, M.A.; Satcher, J.H.; Baumann, T.F. Enhanced thermal transport in carbon aerogel nanocomposites containing double-walled carbon nanotubes. J. Appl. Phys. 2009, 105, 84316. [Google Scholar] [CrossRef]
  76. Baker, W.S.; Long, J.W.; Stroud, R.M.; Rolison, D.R. Sulfur-functionalized carbon aerogels: a new approach for loading high-surface-area electrode nanoarchitectures with precious metal catalysts. J. Non-Cryst. Solids 2004, 350, 80–87. [Google Scholar] [CrossRef]
  77. Marie, J.; Berthon-Fabry, S.; Achard, P.; Chatenet, M.; Pradourat, A.; Chaînet, E. Highly dispersed platinum on carbon aerogels as supported catalysts for PEM fuel cell-electrodes: comparison of two different synthesis paths. J. Non-Cryst. Solids 2004, 350, 88–96. [Google Scholar] [CrossRef]
  78. Du, H.; Li, B.; Kang, F.; Fu, R.; Zeng, Y. Carbon aerogel supported Pt–Ru catalysts for using as the anode of direct methanol fuel cells. Carbon 2007, 45, 429–435. [Google Scholar] [CrossRef]
  79. Lu, A.-H.; Spliethoff, B.; Schüth, F. Aqueous Synthesis of Ordered Mesoporous Carbon via Self-Assembly Catalyzed by Amino Acid. Chem. Mater. 2008, 20, 5314–5319. [Google Scholar] [CrossRef]
  80. Lin, C.; Ritter, J.A. Carbonization and activation of sol–gel derived carbon xerogels. Carbon 2000, 38, 849–861. [Google Scholar] [CrossRef]
  81. Nardecchia, S.; Carriazo, D.; Ferrer, M.L.; Gutiérrez, M.C.; Del Monte, F. Three dimensional macroporous architectures and aerogels built of carbon nanotubes and/or graphene: synthesis and applications. Chem. Soc. Rev. 2013, 42, 794–830. [Google Scholar] [CrossRef] [PubMed]
  82. Ghosh, S.; Maiyalagan, T.; Basu, R.N.; Thandavarayan, M. Nanostructured conducting polymers for energy applications: towards a sustainable platform. Nanoscale 2016, 8, 6921–6947. [Google Scholar] [CrossRef] [PubMed]
  83. Yin, X.; Liu, Q.; Chen, Y.; Xu, A.; Wang, Y.; Tu, Y.; Han, W. Preparation, characterization and environmental application of the composite electrode TiO2-NTs/SnO2-Sb with carbon aerogels. J. Chem. Technol. Biotechnol. 2019, 94, 3124–3133. [Google Scholar] [CrossRef]
  84. Wu, M.; Jin, Y.; Zhao, G.; Li, M.; Li, N. Electrosorption-promoted Photodegradation of Opaque Wastewater on A Novel TiO2/Carbon Aerogel Electrode. Environ. Sci. Technol. 2010, 44, 1780–1785. [Google Scholar] [CrossRef] [PubMed]
  85. Cui, H.; Zhao, Z.; Liang, Y.; Shi, J.; Wu, D.; Liu, H.; Fu, R. Influence of Carbon Aerogel (CA) Pore Structure on Photodegradation of Methyl Orange over TiO2/CA. Chin. J. Catal. (Chin. Version) 2011, 32, 321–324. [Google Scholar] [CrossRef]
  86. Jin, Y.; Wu, M.; Zhao, G.; Li, M. Photocatalysis-enhanced electrosorption process for degradation of high-concentration dye wastewater on TiO2/carbon aerogel. Chem. Eng. J. 2011, 168, 1248–1255. [Google Scholar] [CrossRef]
  87. Jin, Y.; Zhao, G.; Wu, M.; Lei, Y.; Li, M.; Jin, X. In Situ Induced Visible-Light Photoeletrocatalytic Activity from Molecular Oxygen on Carbon Aerogel-Supported TiO2. J. Phys. Chem. C 2011, 115, 9917–9925. [Google Scholar] [CrossRef]
  88. Wang, Y.; Zhang, Y.-N.; Zhao, G.; Wu, M.; Li, M.; Li, N.; Zhang, Y.; Zhang, Y. Electrosorptive photocatalytic degradation of highly concentrated p-nitroaniline with TiO2 nanorod-clusters/carbon aerogel electrode under visible light. Sep. Purif. Technol. 2013, 104, 229–237. [Google Scholar] [CrossRef]
  89. Cui, H.; Liang, Z.-X.; Zhang, J.; Liu, H.; Shi, J. Enhancement of the photocatalytic activity of a TiO2/carbon aerogel based on a hydrophilic secondary pore structure. RSC Adv. 2016, 6, 68416–68423. [Google Scholar] [CrossRef]
  90. Wang, Y.; Pan, Z.; Qin, D.; Bai, S.; Peng, Q. Preparation of Ce-TiO2/carbon aerogel electrode and its performance in degradation of 4-chlorophenol. J. Rare Earths 2018, 36, 374–378. [Google Scholar] [CrossRef]
  91. Justh, N.; Mikula, G.J.; Bakos, L.P.; Nagy, B.; László, K.; Parditka, B.; Erdélyi, Z.; Takáts, V.; Mizsei, J.; Szilágyi, I.M. Photocatalytic properties of TiO2@polymer and TiO2@carbon aerogel composites prepared by atomic layer deposition. Carbon 2019, 147, 476–482. [Google Scholar] [CrossRef] [Green Version]
  92. Wu, F.; Zhou, Z.; Hicks, A. Life Cycle Impact of Titanium Dioxide Nanoparticle Synthesis through Physical, Chemical, and Biological Routes. Environ. Sci. Technol. 2019, 53, 4078–4087. [Google Scholar] [CrossRef] [PubMed]
  93. Fernandes, S.; Esteves da Silva, J.C.G.; Pinto da Silva, L. Life Cycle Assessment of the Sustainability of Enhancing the Photodegradation Activity of TiO2 with Metal-Doping. Materials 2020, 13, 1487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. A schematic diagram of the reaction of resorcinol and formaldehyde to synthesize carbon aerogel. Copied with permission from [54], Copyrights 1989 Springer [54].
Figure 1. A schematic diagram of the reaction of resorcinol and formaldehyde to synthesize carbon aerogel. Copied with permission from [54], Copyrights 1989 Springer [54].
Materials 13 01734 g001
Figure 2. Scanning electronic microscopy (SEM) images of carbon aerogel (CA) (A × 500, B × 80 k) and TiO2/CA (C × 500, D × 80 k, E × 450 k). Copied with the permission from [84], Copyrights 2010 ACS [84].
Figure 2. Scanning electronic microscopy (SEM) images of carbon aerogel (CA) (A × 500, B × 80 k) and TiO2/CA (C × 500, D × 80 k, E × 450 k). Copied with the permission from [84], Copyrights 2010 ACS [84].
Materials 13 01734 g002
Figure 3. The photocatalytic degradation curves of methyl orange dye over TiO2 and TiO2/CA photocatalysts. Copied with the permission from [85], Copyrights 2011 Elsevier [85].
Figure 3. The photocatalytic degradation curves of methyl orange dye over TiO2 and TiO2/CA photocatalysts. Copied with the permission from [85], Copyrights 2011 Elsevier [85].
Materials 13 01734 g003
Figure 4. Mechanism of photocatalysis enhanced electroadsorption (PES) degradation of alizarin red (AR) dye on the TiO2/CA photoelectrode. Copied with the permission from [86], Copyrights 2011 Elsevier [86]. CB: Conduction Band. VB: Valence band.
Figure 4. Mechanism of photocatalysis enhanced electroadsorption (PES) degradation of alizarin red (AR) dye on the TiO2/CA photoelectrode. Copied with the permission from [86], Copyrights 2011 Elsevier [86]. CB: Conduction Band. VB: Valence band.
Materials 13 01734 g004
Figure 5. Mechanism of in-situ induced visible-light photoelectrocatalysis from molecular oxygen on carbon aerogels-supported TiO2. Copied with the permission from [87], Copyrights 2011 ACS [87].
Figure 5. Mechanism of in-situ induced visible-light photoelectrocatalysis from molecular oxygen on carbon aerogels-supported TiO2. Copied with the permission from [87], Copyrights 2011 ACS [87].
Materials 13 01734 g005
Figure 6. Mechanism of the decomposition of intermediates in TiO2 nanorod cluster (NRC)/CA photoelectrocatalysis over p-nitroaniline pollutants. Copied with the permission from [88], Copyrights 2013 Elsevier [88].
Figure 6. Mechanism of the decomposition of intermediates in TiO2 nanorod cluster (NRC)/CA photoelectrocatalysis over p-nitroaniline pollutants. Copied with the permission from [88], Copyrights 2013 Elsevier [88].
Materials 13 01734 g006
Figure 7. Schematic illustration of dimethyl phthalate (DMP) degradation in TiO2/CA-9.3 under light irradiation. The red ring describes the reactions occurring in the secondary pores. The processes include the following: (I) on the external surface, OH• formation with H2O reduction and oxidation; (II) the photoelectrons generated on the external surface migrate to the internal surface through the carbon layer; (III) on the internal surface, the photoelectrons are trapped by O2 to produce O2•−; (IV) HOc are generated with the participation of H2O and O2•− and (V) DMP is oxidized by active HO• on both the external and internal surface. Copied with the permission from [89], Copyrights 2016 RSC [89].
Figure 7. Schematic illustration of dimethyl phthalate (DMP) degradation in TiO2/CA-9.3 under light irradiation. The red ring describes the reactions occurring in the secondary pores. The processes include the following: (I) on the external surface, OH• formation with H2O reduction and oxidation; (II) the photoelectrons generated on the external surface migrate to the internal surface through the carbon layer; (III) on the internal surface, the photoelectrons are trapped by O2 to produce O2•−; (IV) HOc are generated with the participation of H2O and O2•− and (V) DMP is oxidized by active HO• on both the external and internal surface. Copied with the permission from [89], Copyrights 2016 RSC [89].
Materials 13 01734 g007
Figure 8. Schematics of the mechanism of Ce-TiO2/carbon aerogel electrode in the photoelectrocatalytic degradation of 4-chlorophenol pollutant. Copied with the permission from [90], Copyrights 2018 Elsevier [90].
Figure 8. Schematics of the mechanism of Ce-TiO2/carbon aerogel electrode in the photoelectrocatalytic degradation of 4-chlorophenol pollutant. Copied with the permission from [90], Copyrights 2018 Elsevier [90].
Materials 13 01734 g008
Figure 9. Photocatalytic results of (A) resorcinol–formaldehyde aerogel (RFA) and its composite, (B) resorcinol–formaldehyde carbon aerogel (RFCA) and its composites, and (C) specific surface area and heating corrected data of RFCA and its composites. Copied with the permission from [91], Copyrights 2019 Elsevier [91].
Figure 9. Photocatalytic results of (A) resorcinol–formaldehyde aerogel (RFA) and its composite, (B) resorcinol–formaldehyde carbon aerogel (RFCA) and its composites, and (C) specific surface area and heating corrected data of RFCA and its composites. Copied with the permission from [91], Copyrights 2019 Elsevier [91].
Materials 13 01734 g009
Table 1. Summary of photo(electro)catalytic performance of TiO2/carbon aerogel (CA) materials for pollutants removal. NRC, nanorod cluster; RFCA, resorcinol–formaldehyde carbon aerogel.
Table 1. Summary of photo(electro)catalytic performance of TiO2/carbon aerogel (CA) materials for pollutants removal. NRC, nanorod cluster; RFCA, resorcinol–formaldehyde carbon aerogel.
CatalystConditionPollutantRemoval Rate ConstantRef.
TiO2/CA365 nm UV-irradiation under −0.6 V biasMethylene Blue (150 mg/L)10.27 × 10−3 min−1[84]
TiO2/CA 125300 W Hg lampMethyl Orange (10 mg/L)46.2 × 10−3 min−1[85]
TiO2/CA80 W UV light (320–400 nm, peak at 365 nm) 0.6 V biasAlizarin Red (400 mg/L)9.24 × 10−3 min−1[86]
TiO2/CA300 W Xe lamp (420–800 nm) 100 mW/cm2 −0.9 V biasRhodamine 6G (50 mg/L)3.61 × 10−3 min−1[87]
TiO2NRC/CA500 W Xe lamp (peak at 420 nm) −0.6 V biasp-Nitroaniline (150 mg/L)23.1 × 10−3 min−1[88]
TiO2/CA-9.3300 W Xe lamp (320–700 nm peak at 420 nm) 85 mW/cm2Dimethyl phthalate (2 mg/L)12.6 × 10−3 min−1[89]
Ce-TiO2/CA500 W Xe lamp 0.6 V bias4-Chlorophenol (100 mg/L)9.24 × 10−3 min−1[90]
RFCA/TiO2/80 °C2 parallel UV lights (18 W UV-A blacklights)Methyl Orange (8 × 10−5 mol/L)3.3 × 10−3 min−1[91]

Share and Cite

MDPI and ACS Style

Ma, D.; Li, J.; Liu, A.; Chen, C. Carbon Gels-Modified TiO2: Promising Materials for Photocatalysis Applications. Materials 2020, 13, 1734. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13071734

AMA Style

Ma D, Li J, Liu A, Chen C. Carbon Gels-Modified TiO2: Promising Materials for Photocatalysis Applications. Materials. 2020; 13(7):1734. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13071734

Chicago/Turabian Style

Ma, Dongge, Jundan Li, Anan Liu, and Chuncheng Chen. 2020. "Carbon Gels-Modified TiO2: Promising Materials for Photocatalysis Applications" Materials 13, no. 7: 1734. https://0-doi-org.brum.beds.ac.uk/10.3390/ma13071734

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