Next Article in Journal
Catalysts for Syngas Production
Previous Article in Journal
Trends and Outlook of Computational Chemistry and Microkinetic Modeling for Catalytic Synthesis of Methanol and DME
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Porous Hybrid Materials Based on Mesotetrakis(Hydroxyphenyl) Porphyrins and TiO2 for Efficient Visible-Light-Driven Hydrogen Production

MOE Laboratory of Bioinorganic and Synthetic Chemistry, Lehn Institute of Functional Materials, School of Chemistry, and School of Materials Science and Engineering, Sun Yat-sen University, Guangzhou 510275, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(6), 656; https://doi.org/10.3390/catal10060656
Submission received: 7 May 2020 / Revised: 28 May 2020 / Accepted: 7 June 2020 / Published: 11 June 2020
(This article belongs to the Section Photocatalysis)

Abstract

:
A series of highly robust nano-micro hybrid materials based on meso-tetra(4-hydroxyphenyl) porphyrins (M = H, Pd, Zn) and titanium dioxide (denoted as THPP-TiO2, THPP-Pd-TiO2, and THPP-Zn-TiO2) have been prepared by a facile sol-gel method for the first time. When Pt nanoparticles are incorporated in these hybrids, Pt/THPP-Pd-TiO2 achieves good H2 production activity (2025.4 μmol g−1 h−1 and 12.03 μmol m−2 h−1), higher than that of Pt/THPP-Zn-TiO2 (1239.8 μmol g−1 h−1 and 7.46 μmol m−2 h−1) and Pt/THPP-TiO2 (576.8 μmol g−1 h−1 and 4.02 μmol m−2 h−1), owing to the different central metal ions in porphyrins. The best activity of Pt/THPP-Pd-TiO2 would be attributed to the two-center catalysis from coordination Pd metal ions and Pt nanoparticles, while the higher activity of Pt/THPP-Zn-TiO2 than Pt/THPP-TiO2 could be ascribed to the more effective light harvesting and electron transfer between THPP-Zn and TiO2. In addition, the hybridized Pt/THPP-Pd-TiO2 catalyst exhibits unattenuated hydrogen production stability even after recycling the experiment 10 times (cumulative turnover number of 5111 after 50 h), far superior to that of the surface-sensitized Pt/THPP-Pd/TiO2 catalyst with analogous components, due to the more stable Ti-O bonds between four phenols in porphyrins and TiO2 for the hybrid system. The present study provides a promising approach for constructing stable organic–inorganic hybrid systems with unique hierarchical structures for efficient light absorption and electron transfer.

Graphical Abstract

1. Introduction

Photocatalytic hydrogen generation over the semiconductor-based photocatalyst under solar light irradiation is a promising approach to solving current energy and environmental issues [1,2,3,4]. The active and stable photocatalysts are vital for highly efficient hydrogen production. As a good candidate photocatalyst, TiO2 has been widely used owing to its low toxicity, abundance and stability against the light irradiation [5,6,7,8,9,10]. The photoactivity of TiO2 is closely relevant to particle size, crystal form, surface morphology, and pore structure, and therefore the proper engineering of TiO2 in photocatalytic system can greatly improve its activity. TiO2 with a large surface area and micro/mesoporous structures can facilitate reactant/product transfer and enhance optical absorption efficiency, which shows a perfect performance in dye-sensitized TiO2 photocatalysis [11,12,13]. However, pure TiO2 has poor response to visible light due to its wide band gap and fast electron-hole recombination, which limit its photocatalytic activity. In general, noble metal (Au, Ag, and Pt) is usually loaded on TiO2, which can effectively inhibit fast charge carrier recombination [14].
To enhance the light response of the TiO2-based photocatalytic systems, the two most efficient strategies, bandgap engineering (cation/anion doping, or semiconductor heterojunction construction, etc.) [15,16,17,18] and dye sensitization [11,12,13], have been proposed. In this regard, sensitizing TiO2 with photosensitizer is a facile but effective way to realize the visible-light-driven photocatalysis of TiO2. Among all the reported dye sensitizers, metal complexes and metal-free organic dyes are widely applied in TiO2-based photocatalytic hydrogen evolution [11]. Porphyrins and their metallic derivatives with conjugate structures have attracted much attention due to their excellent light absorption, prospective photoexcitation, electron-transfer ability, low cost and excellent thermal and chemical stability [19,20]. Moreover, their photoelectronic properties can be easily tuned by introducing diverse functional groups on the peripheries or metal ions on the centers of their macrocycles. Thus, porphyrins are promising candidates for constructing functional nanomaterials. R. Banerjee et al. reported the synthesis of a porphyrin-based porous polymer and its application in sensitizing TiO2 nanoparticles by ball-milling method, and the obtained composite showed higher activity for H2 evolution than the parent precursor under visible light irradiation [21]. Liu et al. reported photocatalytic systems with Sn-porphyrins as the sensitizer by surface sensitization, and the H2 evolution activities of a series of Sn-porphyrins with different functional groups and Pt nanoparticles with different surface stabilizers were investigated [22]. Nevertheless, the common dye-sensitized TiO2 photocatalytic systems are still unsatisfactory because the dye would shed from the semiconductor and follow by photodegradation [23]. Moreover, charge injection from dye to TiO2 is an inefficient process, which limits the efficiency of photogenerated charges [24,25]. Due to insufficient adsorption stability and low charge mobility, it is necessary to anchor the dyes onto TiO2 by a covalent bond for exploring durable and efficient photosynthesis systems.
Recently, organic/inorganic hybrid materials have been proven as very promising in obtaining new materials with desirable properties because the combined advantageous characteristics of organic and inorganic parts can realize the optimization and complementarity of their properties [26,27,28,29,30,31]. To prepare TiO2-based nano hybrid materials, many approaches, such as dispersion polymerization [32], hydrothermal method [33], template-assisted growth [34], and sol-gel method [35,36,37] have been developed, among which the sol-gel is a facile method for preparing multi-component and highly dispersive materials.
Herein, with meso-tetra(4-hydroxyphenyl) porphyrins (M = H, Pd, Zn, denoted as THPP, THPP-Pd and THPP-Zn) (Scheme 1) as visible-light absorption antennas, respectively, a series of novel TiO2-based organic/inorganic hybrid nanocomposites (denoted as THPP-TiO2, THPP-Pd-TiO2, and THPP-Zn-TiO2) were fabricated using four firm Ti-O bonds as the interfacial linkers by a simple sol-gel method. When Pt nanoparticles (Pt NPs) were loaded onto the hybrids, these hybrid systems could display stable photocatalytic H2 evolution properties under visible light irradiation, showing an increasing trend in the sequence of Pt/THPP-Pd-TiO2 > Pt/THPP-Zn-TiO2 > Pt/THPP-TiO2 due to the effect of central metal ions in porphyrins. Compared with analogous surface-sensitized Pt/THPP-Pd/TiO2 catalyst, the photocatalytic stability of the hybridized Pt/THPP-Pd-TiO2 was significantly improved. In addition, the mechanism of electron transfer and photocatalytic H2 production in the hybrid materials was investigated.

2. Results and Discussion

The synthesis of the three porphyrins, THPP-TiO2, THPP-Pd-TiO2, THPP-Zn-TiO2 and their materials loaded with Pt NPs, and comparative Pt/THPP-Pd/TiO2 was included in the experimental section. The X-ray diffraction (XRD) patterns of the samples are shown in Figure 1. Five diffraction peaks at 25.53°, 38.12°, 48.22°, 54.18°, and 62.94° of Gel-TiO2, THPP-TiO2, THPP-Pd-TiO2 and THPP-Zn-TiO2 samples corresponded to (101), (103), (200), (105), and (204) planes of the crystalline anatase structure of titanium dioxide. The diffraction peaks of THPP-TiO2, THPP-Pd-TiO2, and THPP-Zn-TiO2 matched well to those of pure Gel-TiO2, indicating that the presence of porphyrins in the hybrid did not change the crystallinity of TiO2. The relatively regular diffraction peaks without the peaks of other impurities showed the high purity of these samples.
The surface areas of the photocatalyst play an important role in the hydrogen production in the photocatalytic system. Nitrogen adsorption–desorption isotherms were measured at 77 K to obtain the Brunauer–Emmett–Teller (BET) surface areas and pore size distributions of the samples (Figures S1 and S2 and Table S1). The BET surface areas of THPP-TiO2, THPP-Pd-TiO2, THPP-Zn-TiO2, Gel-TiO2, and, for comparison, P25-TiO2 were evaluated to be 143.4, 166.1, 168.3, 162.9, and 55.4 m2 g−1, respectively. The results suggested that the sol-gel method could increase the BET surface areas of TiO2. In addition, the micro/mesoporous structure, comprising both micropores (1~2 nm) and mesopores (3~5 nm), could be observed in porphyrin-TiO2 and Gel-TiO2 materials, in contrast to that (29 nm) of mesoporous P25-TiO2, which shortened the transport length and contributed to the faster diffusion of the electrons from the bulk to the surface, thus inhibiting the charge recombination and increasing the catalytic efficiency.
The morphologies of the as-prepared samples have been observed by the scanning electron microscopy (SEM) and transmission electron microscopy (TEM). The SEM image in Figure 2A reveals that the porphyrin-TiO2 presented a random aggregation of sheets which were closely associated with each other. The SEM elemental mapping of THPP-TiO2, THPP-Pd-TiO2, THPP-Zn-TiO2, indicated that the elements C and N for THPP-TiO2, C, N, and Pd for THPP-Pd-TiO2, and C, N, and Zn for THPP-Zn-TiO2dispersed homogenously throughout the whole hybrid material, respectively (Figure 2B–D).
The TEM image of the porphyrin-TiO2 samples (Figure 3A–C) showed that the hybrid materials were composed of nanospheres with an average diameter of 10 nm. The lattice fringe spacing was estimated to be 0.357 nm, in accordance with the (101) plane of anatase TiO2 (Figure 3D–F). The selected area electron diffraction (SAED) results of the porphyrin-TiO2 samples confirmed the formation of the anatase phase in TiO2 (Figure 3G–I).
UV-vis absorption spectra of three porphyrins in DMF solution and in solid-state, Gel-TiO2, and the hybrid materials based on porphyrins and TiO2 are shown in Figure 4. All the porphyrins exhibited strong absorption in the visible region, ranging from 400 to 700 nm. The peak centered at about 420 nm was ascribed to the Soret absorption band originating from the a1u(π)–eg(π) transition (S0–S2), while the absorption bands ranging from 500 to 700 nm were attributed to the weaker Q bands of a2u(π)–eg(π) transitions (S0–S1). The absorption intensities of the three porphyrins were measured to be close to the values reported in the literature [38]. Compared with THPP, the absorption peak numbers of THPP-Pd and THPP-Zn in the region above 500 nm were reduced from 4 to 2, owing to the increased porphyrin ring symmetry after coordination with metal, which reduced the degree of molecular orbital splitting and improved the degeneracy. Although the maxmium absorption wavelengths (λmax) of the Soret absorption bands were similar, the Q-band absorption wavelengths increased in the order of THPP-Pd < THPP < THPP-Zn. Particularly, THPP-Zn had the highest extinction coefficient and most obvious red-shift among those porphyrins, because it was easier for ZnPyd to occur the metal-to-ligand charge transfer (MLCT) owing to the weaker interaction between Zn2+ (3d10) and Py macrocycle ion without any single electron [39]. This better absorption and charge transfer properties in relation to Zn2+ ions were favorable for the H2 production reaction.
When the porphyrins was doped into TiO2 by the sol-gel method, the hybrids exhibited a red-shift and significantly extended the absorption profile, which most probably originated from the strong interaction between the –OH moieties of porphyrins and –Ti atoms of the TiO2, in agreement with our previous work [40]. For the Gel-TiO2, a strong absorption peak at 325 nm with an absorption edge of 450 nm corresponded to a 3.25 eV band gap. In contrast to Gel-TiO2 with a single optical band gap in Tauc plots (Figure S3), the hybrid materials based on porphyrins and TiO2 showed multiple band gaps. These narrow band gaps (2.6–2.8 eV) could be caused by the doping of the porphyrins, while the wide band gap of 3.0–3.3 eV for the three porphyrin-based hybrids should come from the inorganic component TiO2, respectively.
To prove the feasibility of conducting photocatalytic H2 production, the first oxidation potentials of THPP, THPP-Pd, and THPP-Zn were measured by cyclic voltammograms (CVs) to obtain HOMO value (1.12, 1.08, and 1.10 V vs. NHE) (Figure S4 and Table S2). The zero–zero excitation energies were estimated from the intersection of the absorption and emission spectra as 1.90, 2.19, and 2.05 eV for THPP, THPP-Pd, and THPP-Zn, respectively (Figure S5 and Table S2). The LUMO levels of THPP, THPP-Pd, and THPP-Zn (−0.78, −1.11, and −0.95 V vs. NHE) were calculated from the HOMO values of the porphyrins and the zero-zero excitation energies. The valence-band potential of Gel-TiO2 was reported to be −0.5 V vs. NHE in our previous literature [41], which was more positive than the LUMO values of the three porphyrins and more negative than the redox potential of H+/H2, and thus facilitated electron injection from the porphyrins to TiO2 and favoured the proton reduction to realize photocatalytic H2 generation. Among the three porphyrins, the higher LUMO level of THPP-Pd than those of THPP and THPP-Zn implied that there is a stronger driving force to promote electron transfer from THPP-Pd to TiO2, which could accelerate the rate of hydrogen production. On the other hand, the first oxidation potentials of the porphyrins were more positive than that of the redox couple (0.84 ± 0.12 V vs. NHE) of triethanolamine as a sacrificial agent [42], beneficial to promoting dye regeneration and suppressing the recombination between the injected electrons and the dye cation radical (Figure 5).
For evaluating the influence of coordination metal ions on the photoactivity of the as-prepared materials, the photocatalytic experiments were performed in the visible light region with the Pt/THPP-TiO2, Pt/THPP-Pd-TiO2 and Pt/THPP-Zn-TiO2 catalysts using 0.5 wt% Pt particles as co-catalyst and TEOA as sacrificial electron donor. Figure 6 illustrated the H2 evolution plots of these hybrid systems and the corresponding data were summarized in Table S3. Pt/THPP-Pd-TiO2 catalyst exhibited higher hydrogen production activity (2025.4 μmol g−1 h−1 based on the catalyst mass and 12.03 μmol m−2 h−1 based on the surface area) than those of Pt/THPP-Zn-TiO2 (1239.8 μmol g−1 h−1 and 7.46 μmol m−2 h−1) and Pt/THPP-TiO2 (576.8 μmol g−1 h−1 and 4.02 μmol m−2 h−1). The result indicated that coordinated Pd metal ions play an essential role in the photoactivity of Pt/THPP-Pd-TiO2. For the known photochemical molecular devices (PMDs), the N^N-chelated NnMX2 (M = Pd2+, Pt2+ or Rh2+, X = Cl, Br or I, n = 2 or 4) coordination polyhedron was commonly employed as a catalyst centre, in which the dissociation of the terminal X anions during the photocatalytic process was found [43,44]. It was also reported that Pd(Py)4 motifs possessing a PdN4 environment with four monodentate pyridyl ligands could be used as the catalytic centres for photocatalytic H2 evolution, where the breaking of the Pd-N bond and the formation of hydride intermediate may happen [45,46]. Thus, we deduced that the higher amount of H2 exhibited by the Pt/THPP-Pd-TiO2 hybrid may be attributed to the two-center catalysis from coordination Pd metal ions of porphyrin dye and Pt nanoparticles. In addition, the better catalytic activity of Pt/THPP-Zn-TiO2 than Pt/THPP-TiO2 could be attributed to the more effective light harvesting and charge transfer between THPP-Zn and TiO2 mentioned above.The apparent quantum yields (AQYs) of corresponding samples were measured and followed the order of Pt/THPP-Pd-TiO2 > Pt/THPP-Zn-TiO2 > Pt/THPP-TiO2 (Table S4), in consistance with the results of their photocatalytic H2 production.
In order to confirm the efficiencies of the photocatalysts and their stability under UV light irradiation, control experiments of the three catalysts under complete UV-vis irradiation have been performed (Figure S6) and all catalysts exhibited higher H2 production activities than those under visible light, owing to the effective UV-responses of the porphyrins and TiO2 substrate. It was noteworthy that Pt/THPP-Zn-TiO2 showed the highest activity under complete UV-vis irradiation. This result could be ascribed to the better absorption of THPP-Zn compared to THPP-Pd and THPP in the UV band, which could be observed in the UV-vis absorption spectra of THPP, THPP-Pd, and THPP-Zn (Figure 4).
To check the contribution of the Pd coordination metal ions to H2 production, the control experiments with THPP-TiO2, THPP-Pd-TiO2 and THPP-Zn-TiO2 without Pt NPs for photocatalysis have been performed (Figure 7 and Table S3). It was found that the H2 generation rate of THPP-Pd-TiO2 without Pt NPs was 230.1 μmol g−1 h−1 (1.37 μmol m−2 h−1), while no obvious amount of H2 was detected for THPP-TiO2 and THPP-Zn-TiO2. These results supported that Pd coordination metal ions were conducive to the photocatalytic H2 evolution activity of THPP-Pd-TiO2.
To study the difference of catalytic activity of the hybrid system Pt/THPP-Pd-TiO2 and the surface sensitization system Pt/THPP-Pd/TiO2, the control experiments of two catalysts were carried out (Figure 8 and Table S5). In the initial 5 h light irradiation, two catalysts with 0.5 wt% Pt loading and approximate porphyrin amounts exhibited the same H2 evolution activity. However, in second and third run, H2-production activity of hybrid Pt/THPP-Pd-TiO2 were rising from initial 2025.4 to 2755.9 μmol g−1 h−1 (12.03 to 16.38 μmol m−2 h−1) in contrast with the continuous decline in the activity of surface sensitization Pt/THPP-Pd/TiO2 to 1110.3 μmol g−1 h−1 (6.81 μmol m−2 h−1) after the third run. During the photocatalytic reaction process, the THPP-Pd dye seriously detached from the Pt/THPP-Pd/TiO2 material under visible light irradiation, but the detaching phenomena were not contributed to the Pt/THPP-Pd-TiO2 catalysts even after they were used for 10 times and recycled. As shown in Figure 9 and Table S5, the Pt/THPP-Pd-TiO2 catalyst showed an outstanding stability without a decrease in photocatalysis. The TONH2 calculated as the number of moles of H2 divided by the number of moles of Pt was 5111 after 50 h, indicating the higher photocatalytic activity and stability of the hybrid materials than those of the common surface dye-sensitized system. The inorganic component TiO2 in the hybrid materials was expected to be able to protect the THPP-Pd, and the stable linkages between THPP-Pd and TiO2 were beneficial for efficient electron transfer from THPP-Pd* to TiO2. Thus, organic–inorganic components in the hybrid system, in other words, have a synergistic effect on the photocatalytic reaction.

3. Conclusions

In summary, we have, for the first time, developed a series of unique hierarchical hybrid materials (THPP-TiO2, THPP-Pd-TiO2, and THPP-Zn-TiO2) by combining TiO2 and porphyrins through a sol-gel method, which possessed a highly accessible surface area as well as delivered excellent charge separation and transport efficiency due to the stable Ti–O linkage, resulting in efficient and persistent photocatalytic water splitting activities when loaded with Pt nanoparticles. The absorption spectra, electrochemical properties and charge transfer processes of those porphyrins were investigated to vary along with the central metal ions. The effects of central metal ions led to the catalytic properties of the hybrids for H2 production indicating a rising trend in the sequence of Pt/THPP-Pd-TiO2 > Pt/THPP-Zn-TiO2 > Pt/THPP-TiO2. Especially, Pt/THPP-Pd-TiO2 showed the best H2 production activity (2025.4 μmol g−1 h−1 and 12.03 μmol m−2 h−1) among the three hybrids under visible light irradiation owing to two catalytic centers from coordination Pd metal ions of porphyrin dye and Pt nanoparticles. The superior activity of Pt/THPP-Zn-TiO2 to Pt/THPP-TiO2 could be attributed to the more effective optical absorption of THPP-Zn and electron transfer between THPP-Zn and TiO2. Comparatively, the hybrid Pt/THPP-Pd-TiO2 exhibited superior stability to the analogous surface dye-sensitized Pt/THPP-Pd/TiO2 system under similar photocatalytic conditions and an impressive TON of 5111 has been achieved through a ten-cycle photoreaction without a decrease for 50 h. The present results reveal that varying the central metal ions in porphyrins or conducting interface and microstructure modulation is a facile strategy to optimize the photogenerated charge transfer pathway and utilize the visible light of sunlight to significantly enhance the efficiency of the solar-to hydrogen conversion efficiency.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/6/656/s1, Figure S1: N2 adsorption-desorption isotherms of Gel-TiO2, THPP-TiO2, THPP-Pd-TiO2 and THPP-Zn-TiO2 at 77 K, Figure S2: The pore size distributions of A) THPP-TiO2, B) THPP-Pd-TiO2 and C) THPP-Zn-TiO2, Table S1: BET surface areas, pore volumes and pore widths of Gel-TiO2, THPP-TiO2, THPP-Pd-TiO2 and THPP-Zn-TiO2 derived from 77 K N2 sorption isotherms, Figure S3: Tauc plots of THPP-TiO2, THPP-Pd-TiO2, THPP-Zn-TiO2 and Gel-TiO2, Figure S4: Cyclic voltammogram of THPP, THPP-Pd and THPP-Zn in 0.1 M TBAPF6 THF solution with a scan rate of 40 mV s-1, Figure S5. Absorption and emission spectra of A) THPP, B) THPP-Pd, and C) THPP-Zn measured in THF solution, Table S2: The energy levels and electrochemical data of THPP, THPP-Pd and THPP-Zn, Table S3: H2 production data and calculated TON values of hybrid catalysts of THPP, THPP-Pd and THPP-Zn with or without Pt NPs under visible light in 5 h, Table S4: AQY results for H2 production of Pt/THPP-TiO2, Pt/THPP-Pd-TiO2 and Pt/THPP-Zn-TiO2 catalysts (6 mg for each sample) after 0.5 h irradiation under 425 nm LED light source, Figure S6: H2 production curves of Pt/THPP-TiO2, Pt/THPP-Pd-TiO2, and Pt/THPP-Zn-TiO2 in an aqueous solution of 10% TEOA under complete UV-vis irradiation, Table S5: H2 production data and calculated TON values of Pt/THPP-Pd-TiO2 and Pt/THPP-Pd/TiO2 in cyclic photocatalytic experiment.

Author Contributions

J.-M.L. designed the research and wrote the paper. J.-F.H. directed the experiments and helped in paper preparation. L.-Y.H. carried out most of the syntheses and measurements. Y.L. and S.Q. helped in experiments and data analyses. All authors discussed the results and commented on the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation Project of China (grant no. 21572280, 21975291, and 61772053), the NSF of Guangdong Province (grant no. 2019A1515011640), Science and Technology Plan Project of Guangzhou (grant no. 201707010114), and the FRF for the Central Universities (grant no. 19lgpy12).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, S.; Takata, T.; Domen, K. Particulate photocatalysts for overall water splitting. Nat. Rev. Mater. 2017, 2, 17050. [Google Scholar] [CrossRef]
  2. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef] [PubMed]
  3. Tilley, S.D. Recent advances and emerging trends in photo-electrochemical solar energy conversion. Adv. Energy Mater. 2019, 9, 1802877. [Google Scholar] [CrossRef] [Green Version]
  4. Qi, J.; Zhang, W.; Cao, R. Solar-to-hydrogen energy conversion based on water splitting. Adv. Energy Mater. 2018, 8, 1701620. [Google Scholar] [CrossRef]
  5. Chen, X.; Li, C.; Grätzel, M.; Kostecki, R.; Mao, S.S. Nanomaterials for renewable energy production and storage. Chem. Soc. Rev. 2012, 41, 7909–7937. [Google Scholar] [CrossRef]
  6. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C 2012, 13, 169–189. [Google Scholar] [CrossRef]
  7. Kubacka, A.; Ferna’ndez-Garcı’a, M.; Colo’n, G. Advanced nanoarchitectures for solar photocatalytic applications. Chem. Rev. 2012, 112, 1555–1614. [Google Scholar] [CrossRef] [PubMed]
  8. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  9. Miyoshi, A.; Nishioka, S.; Maeda, K. Water splitting on rutile TiO2-based photocatalysts. Chem. Eur. J. 2018, 24, 18204–18219. [Google Scholar] [CrossRef]
  10. Lee, C.-Y.; Zou, J.; Bullock, J.; Wallace, G.G. Emerging approach in semiconductor photocatalysis: Towards 3D architectures for efficient solar fuels generation in semi-artificial photosynthetic systems. J. Photochem. Photobiol. C 2019, 39, 142–160. [Google Scholar] [CrossRef]
  11. Cecconi, B.; Manfredi, N.; Montini, T.; Fornasiero, P.; Abbotto, A. Dye-sensitized solar hydrogen production: The emerging role of metal-free organic sensitizers. Eur. J. Org. Chem. 2016, 31, 5194–5215. [Google Scholar] [CrossRef]
  12. Ho, P.-Y.; Wang, Y.; Yiu, S.-C.; Yu, W.-H.; Ho, C.-L.; Huang, S. Starburst triarylamine donor-based metal-free photosensitizers for photocatalytic hydrogen production from water. Org. Lett. 2017, 5, 1048–1051. [Google Scholar] [CrossRef] [PubMed]
  13. Tiwari, A.; Duvva, N.; Rao, V.N.; Venkatakrishnan, S.M.; Giribabu, L.; Pal, U. Tetrathiafulvalene scaffold-based sensitizer on hierarchical porous TiO2: Efficient light-harvesting material for hydrogen production. J. Phys. Chem. C 2019, 123, 70–81. [Google Scholar] [CrossRef]
  14. Su, R.; Tiruvalam, R.; Logsdail, A.J.; He, Q.; Downing, C.A.; Jensen, M.T.; Dimitratos, N.; Kesavan, L.; Wells, P.P.; Bechstein, R.; et al. Designer titania-supported Au-Pd nanoparticles for efficient photocatalytic hydrogen production. ACS Nano 2014, 8, 3490–3497. [Google Scholar] [CrossRef] [PubMed]
  15. Ma, Y.; Wang, X.; Jia, Y.; Chen, X.; Han, H.; Li, C. Titanium dioxide-based nanomaterials for photocatalytic fuel generations. Chem. Rev. 2014, 114, 9987–10043. [Google Scholar] [CrossRef]
  16. Fang, W.; Xing, M.; Zhang, J. Modifications on reduced titanium dioxide photocatalysts: A review. J. Photochem. Photobiol. C 2017, 32, 21–39. [Google Scholar] [CrossRef]
  17. Navarro, R.M.; Alvarez-Galvan, M.C.; Villoria de la Mano, J.A.; Al-Zahrani, S.M.; Fierro, J.L.G. A framework for visible-light water splitting. Energy Environ. Sci. 2010, 3, 1865–1882. [Google Scholar] [CrossRef]
  18. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  19. Urbani, M.; Grätzel, M.; Nazeeruddin, M.K.; Torres, T. Meso-substituted porphyrins for dye-sensitized solar cells. Chem. Rev. 2014, 114, 12330–12396. [Google Scholar] [CrossRef]
  20. Song, H.; Liu, Q.; Xie, Y. Porphyrin-sensitized solar cells: Systematic molecular optimization, coadsorption and cosensitization. Chem. Commun. 2018, 54, 1811–1824. [Google Scholar] [CrossRef]
  21. Mukherjee, G.; Thote, J.; Aiyappa, H.B.; Kandambeth, S.; Banerjee, S.; Vanka, K.; Banerjee, R. A porous porphyrin organic polymer (PPOP) for visible light triggered hydrogen production. Chem. Commun. 2017, 53, 4461–4464. [Google Scholar] [CrossRef] [PubMed]
  22. Koposova, E.; Liu, X.; Pendin, A.A.; Thiele, B.; Shumilova, G.; Ermolenko, Y.; Offenhäusser, A.; Mourzina, Y. Influence of Meso-substitution of the Porphyrin ring on enhanced hydrogen evolution in a photochemical system. J. Phys. Chem. C 2016, 120, 13873–13890. [Google Scholar] [CrossRef]
  23. Huang, J.-F.; Lei, Y.; Xiao, L.-M.; Chen, X.-L.; Zhong, Y.-H.; Qin, S.; Liu, J.-M. Photocatalysts for H2 generation from starburst triphenylamine/carbazole donor-based metal-free dyes and porous anatase TiO2 cube. ChemSusChem 2020, 13, 1037–1043. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.; Hong, J.; Zhang, W.; Xu, R. Carbon nitride nanosheets for photocatalytic hydrogen evolution: Remarkably enhanced activity by dye sensitization. Catal. Sci. Technol. 2013, 3, 1703–1711. [Google Scholar] [CrossRef]
  25. Min, S.X.; Lu, G.X. Dye-sensitized reduced graphene oxide photocatalysts for highly efficient visible-light-driven water reduction. J. Phys. Chem. C 2011, 115, 13938–13945. [Google Scholar] [CrossRef]
  26. Chen, Q.; Marco, N.D.; Yang, Y.; Song, T.-B.; Chen, C.-C.; Zhao, H.; Hong, Z.; Zhou, H.; Yang, Y. Under the spotlight: The organic-inorganic hybrid halide perovskite for optoelectronic applications. Nano Today 2015, 10, 355–396. [Google Scholar] [CrossRef] [Green Version]
  27. Pan, Z.; Yao, L.; Zhai, J.; Yao, X.; Chen, H. Interfacial coupling effect in organic/inorganic nanocomposites with high energy density. Adv. Mater. 2018, 30, 1705662. [Google Scholar]
  28. Chen, Y.; Shi, J. Chemistry of mesoporous organosilica in nanotechnology: Molecularly organic-inorganic hybridization into frameworks. Adv. Mater. 2016, 28, 3235–3272. [Google Scholar] [CrossRef]
  29. Gobbi, M.; Orgiu, E.; Samorì, P. When 2D materials meet molecules: Opportunities and challenges of hybrid organic/inorganic van der waals heterostructures. Adv. Mater. 2018, 30, 1706103. [Google Scholar] [CrossRef] [Green Version]
  30. Lazzara, G.; Cavallaro, G.; Panchal, A.; Fakhrullin, R.; Stavitskaya, A.; Vinokurov, V.; Lvov, Y. An assembly of organic-inorganic composites using halloysite clay nanotubes. Curr. Opin. Colloid Interface Sci. 2018, 35, 42–50. [Google Scholar] [CrossRef]
  31. Kaushik, A.; Kumar, R.; Arya, S.K.; Nair, M.; Malhotra, B.D.; Bhansali, S. Organic-inorganic hybrid nanocomposite-based gas sensors for environmental monitoring. Chem. Rev. 2015, 115, 4571–4606. [Google Scholar] [CrossRef] [PubMed]
  32. Matsuyama, K.; Mishima, K. Preparation of poly(methyl methacrylate)-TiO2 nanoparticle composites by pseudo-dispersion polymerization of methyl methacrylate in supercritical CO2. J. Supercrit. Fluids 2009, 49, 256–264. [Google Scholar] [CrossRef]
  33. Liu, N.; Chen, X.; Zhang, J.; Schwank, J.W. A review on TiO2-based nanotubes synthesized via hydrothermal method: Formation mechanism, structure modification, and photocatalytic applications. Catal. Today 2014, 225, 34–51. [Google Scholar] [CrossRef]
  34. Bhat, T.S.; Mali, S.S.; Korade, S.D.; Shaikh, J.S.; Karanjkar, M.M.; Hong, C.K.; Kim, J.H.; Patil, P.S. Mesoporous architecture of TiO2 microspheres via controlled template assisted route and their photoelectrochemical properties. J. Mater. Sci. Mater. El. 2017, 28, 304–316. [Google Scholar] [CrossRef]
  35. Macwan, D.P.; Dave, P.N.; Chaturvedi, S. A review on nano-TiO2 sol-gel type syntheses and its applications. J. Mater. Sci. 2011, 46, 3669–3686. [Google Scholar] [CrossRef]
  36. Akpan, U.G.; Hameed, B.H. The advancements in sol-gel method of doped-TiO2 photocatalysts. Appl. Catal. A 2010, 375, 1–11. [Google Scholar] [CrossRef]
  37. Zhang, K.; Wang, X.; Guo, X.; Dai, J.; Xiang, J. Fast sol-gel method to prepare mesoporous TiO2 with high photocatalytic activity. Nano 2015, 10, 1550001. [Google Scholar] [CrossRef]
  38. Wang, J.; Liu, D.; Liu, Q.; Peng, T.; Li, R.; Zhou, S. Effects of the central metal ions on the photosensitization of metalloporphyrins over carbon nitride for visible-light-responsive H2 production. Appl. Surf. Sci. 2019, 464, 255–261. [Google Scholar] [CrossRef]
  39. Liao, M.-S.; Scheiner, S. Electronic structure and bonding in metal porphyrins, metal = Fe, Co, Ni, Cu, Zn. J. Chem. Phys. 2002, 117, 205–219. [Google Scholar] [CrossRef] [Green Version]
  40. Chen, Y.-F.; Huang, J.-F.; Shen, M.-H.; Liu, J.-M.; Huang, L.-B.; Zhong, Y.-H.; Qin, S.; Guo, J.; Su, C.-Y. A porous hybrid material based on calixarene dye and TiO2 demonstrating high and stable photocatalytic performance. J. Mater. Chem. A 2019, 7, 19852–19861. [Google Scholar] [CrossRef]
  41. Zhong, Y.-H.; Lei, Y.; Huang, J.-F.; Xiao, L.-M.; Chen, X.-L.; Luo, T.; Qin, S.; Guo, J.; Liu, J.-M. Design of an alkalinous pyridyl acceptor-based calix[4]arene dye and synthesis of stable calixarene-TiO2 porous hybrid materials for efficient photocatalysis. J. Mater. Chem. A 2020, 8, 8883–8891. [Google Scholar] [CrossRef]
  42. Sun, H.; Hoffman, M.Z. Reductive quenching of the excited states of ruthenium(II) complexes containing 2,2’-bipyridine, 2,2’-bipyrazine and 2,2’-bipyrimidine ligands. J. Phys. Chem. A 1994, 98, 11719–11726. [Google Scholar] [CrossRef]
  43. Pfeffer, M.G.; Schäfer, B.; Smolentsev, G.; Uhlig, J.; Nazarenko, E.; Guthmuller, J.; Kuhnt, C.; Wächtler, M.; Dietzek, B.; Sundström, V.; et al. Palladium versus platinum: The metal in the catalytic center of a molecular photocatalyst determines the mechanism of the hydrogen production with visible light. Angew. Chem. Int. Ed. 2015, 54, 5044–5048. [Google Scholar] [CrossRef]
  44. Tschierlei, S.; Presselt, M.; Kuhnt, C.; Yartsev, A.; Pascher, T.; Sundström, V.; Karnahl, M.; Schwalbe, M.; Schäfer, B.; Rau, S.; et al. Photophysics of an intramolecular hydrogen-evolving Ru-Pd photocatalyst. Chem. Eur. J. 2009, 15, 7678–7688. [Google Scholar] [CrossRef]
  45. Dempsey, J.L.; Brunschwig, B.S.; Winkler, J.R.; Gray, H.B. Hydrogen evolution catalyzed by cobaloximes. Acc. Chem. Res. 2009, 42, 1995–2004. [Google Scholar] [CrossRef] [Green Version]
  46. Manbeck, G.F.; Brewer, K.J. Photoinitiated electron collection in polyazine chromophores coupled to water reduction catalysts for solar H2 production. Coord. Chem. Rev. 2013, 257, 1660–1675. [Google Scholar] [CrossRef]
Scheme 1. Molecular structures of THPP, THPP-Pd and THPP-Zn.
Scheme 1. Molecular structures of THPP, THPP-Pd and THPP-Zn.
Catalysts 10 00656 sch001
Figure 1. The XRD patterns of Gel-TiO2, THPP-TiO2, THPP-Pd-TiO2 and THPP-Zn-TiO2.
Figure 1. The XRD patterns of Gel-TiO2, THPP-TiO2, THPP-Pd-TiO2 and THPP-Zn-TiO2.
Catalysts 10 00656 g001
Figure 2. (A) TEM of THPP-TiO2, THPP-Pd-TiO2, and THPP-Zn-TiO2, and the elemental mappings image of (B) THPP-TiO2, (C) THPP-Pd-TiO2 and (D) THPP-Zn-TiO2.
Figure 2. (A) TEM of THPP-TiO2, THPP-Pd-TiO2, and THPP-Zn-TiO2, and the elemental mappings image of (B) THPP-TiO2, (C) THPP-Pd-TiO2 and (D) THPP-Zn-TiO2.
Catalysts 10 00656 g002
Figure 3. TEM of (A) THPP-TiO2, (B) THPP-Pd-TiO2 and (C) THPP-Zn-TiO2; HRTEM of (D) THPP-TiO2, (E) THPP-Pd-TiO2 and (F) THPP-Zn-TiO2, and SAED of (G) THPP-TiO2, (H) THPP-Pd-TiO2 and (I) THPP-Zn-TiO2.
Figure 3. TEM of (A) THPP-TiO2, (B) THPP-Pd-TiO2 and (C) THPP-Zn-TiO2; HRTEM of (D) THPP-TiO2, (E) THPP-Pd-TiO2 and (F) THPP-Zn-TiO2, and SAED of (G) THPP-TiO2, (H) THPP-Pd-TiO2 and (I) THPP-Zn-TiO2.
Catalysts 10 00656 g003
Figure 4. UV-vis absorption spectra of THPP, THPP-Pd, and THPP-Zn (A) in DMF solution and (B) in solid-state, and (C) UV-vis solid-state diffraction spectra of THPP-TiO2, THPP-Pd-TiO2, THPP-Zn-TiO2 and Gel-TiO2.
Figure 4. UV-vis absorption spectra of THPP, THPP-Pd, and THPP-Zn (A) in DMF solution and (B) in solid-state, and (C) UV-vis solid-state diffraction spectra of THPP-TiO2, THPP-Pd-TiO2, THPP-Zn-TiO2 and Gel-TiO2.
Catalysts 10 00656 g004aCatalysts 10 00656 g004b
Figure 5. Proposed mechanism of photocatalytic H2 production over Pt/porphyrin-TiO2.
Figure 5. Proposed mechanism of photocatalytic H2 production over Pt/porphyrin-TiO2.
Catalysts 10 00656 g005
Figure 6. H2 production curves of Pt/THPP-TiO2, Pt/THPP-Pd-TiO2, and Pt/THPP-Zn-TiO2 in an aqueous solution of 10% TEOA under visible-light irradiation (λ > 420 nm).
Figure 6. H2 production curves of Pt/THPP-TiO2, Pt/THPP-Pd-TiO2, and Pt/THPP-Zn-TiO2 in an aqueous solution of 10% TEOA under visible-light irradiation (λ > 420 nm).
Catalysts 10 00656 g006
Figure 7. H2 production curves of THPP-TiO2, THPP-Pd-TiO2, and THPP-Zn-TiO2 in an aqueous solution of 10% TEOA under visible-light irradiation (λ > 420 nm).
Figure 7. H2 production curves of THPP-TiO2, THPP-Pd-TiO2, and THPP-Zn-TiO2 in an aqueous solution of 10% TEOA under visible-light irradiation (λ > 420 nm).
Catalysts 10 00656 g007
Figure 8. Time courses of H2 production over Pt/THPP-Pd-TiO2 and Pt/THPP-Pd/TiO2 under λ > 420 nm light irradiation.
Figure 8. Time courses of H2 production over Pt/THPP-Pd-TiO2 and Pt/THPP-Pd/TiO2 under λ > 420 nm light irradiation.
Catalysts 10 00656 g008
Figure 9. Photocatalytic H2 production against the recyclability of Pt/THPP-Pd-TiO2 under visible light irradiation (λ > 420 nm) in 100 mL H2O/TEOA (9: 1 v/v).
Figure 9. Photocatalytic H2 production against the recyclability of Pt/THPP-Pd-TiO2 under visible light irradiation (λ > 420 nm) in 100 mL H2O/TEOA (9: 1 v/v).
Catalysts 10 00656 g009

Share and Cite

MDPI and ACS Style

Huang, L.-Y.; Huang, J.-F.; Lei, Y.; Qin, S.; Liu, J.-M. Porous Hybrid Materials Based on Mesotetrakis(Hydroxyphenyl) Porphyrins and TiO2 for Efficient Visible-Light-Driven Hydrogen Production. Catalysts 2020, 10, 656. https://doi.org/10.3390/catal10060656

AMA Style

Huang L-Y, Huang J-F, Lei Y, Qin S, Liu J-M. Porous Hybrid Materials Based on Mesotetrakis(Hydroxyphenyl) Porphyrins and TiO2 for Efficient Visible-Light-Driven Hydrogen Production. Catalysts. 2020; 10(6):656. https://doi.org/10.3390/catal10060656

Chicago/Turabian Style

Huang, Li-Yuan, Jian-Feng Huang, Yang Lei, Su Qin, and Jun-Min Liu. 2020. "Porous Hybrid Materials Based on Mesotetrakis(Hydroxyphenyl) Porphyrins and TiO2 for Efficient Visible-Light-Driven Hydrogen Production" Catalysts 10, no. 6: 656. https://doi.org/10.3390/catal10060656

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