Next Article in Journal
Photocatalytic Degradation of Ciprofloxacin by UV Light Using N-Doped TiO2 in Suspension and Coated Forms
Next Article in Special Issue
CuO/ZnO/CQDs@PAN Nanocomposites with Ternary Heterostructures for Enhancing Photocatalytic Performance
Previous Article in Journal
Limestone Calcination Kinetics in Microfluidized Bed Thermogravimetric Analysis (MFB-TGA) for Calcium Looping
Previous Article in Special Issue
Enhancement and Mechanism of Rhodamine B Decomposition in Cavitation-Assisted Plasma Treatment Combined with Fenton Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Butanol Isomerization on Photothermal Hydrogen Production over Ti@TiO2 Core-Shell Nanoparticles

Institute for the Separation Chemistry in Marcoule (ICSM), University Montpellier, UMR 5257, CEA, CNRS, ENSCM, Marcoule, F-30207 Bagnols sur Cèze, France
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1662; https://doi.org/10.3390/catal12121662
Submission received: 10 November 2022 / Revised: 13 December 2022 / Accepted: 15 December 2022 / Published: 17 December 2022

Abstract

:
In this work, we reported for the first time the effect of butanol isomerization on the photothermal production of hydrogen in the presence of a noble, metal-free Ti@TiO2 core-shell photocatalyst. The experiments were performed in aqueous solutions of 1-BuOH, 2-BuOH, and t-BuOH under Xe lamp irradiation (vis/NIR: 8.4 W, UV: 0.6 W) at 35–69 °C. The increase in temperature significantly enhanced H2 formation, indicating a strong photothermal effect in the studied systems. However, in dark conditions, H2 emission was not observed even at elevated temperatures, which clearly points out the photonic origin of H2 photothermal formation. The rate of H2 production followed the order of 1-BuOH >> 2-BuOH > t-BuOH in the entire range of studied temperatures. In the systems with 1-BuOH and 2-BuOH, hydrogen was the only gaseous product measured online in the outlet carrier argon using mass spectrometry. By contrast, a mixture of H2, CH4, and C2H6 was detected for t-BuOH, indicating a C–C bond scission with this isomer during photocatalytic degradation. The apparent activation energies, Ea, with 1-BuOH/2-BuOH isomers (20–21 kJ·mol−1) was found to be larger than for t-BuOH (13 kJ·mol−1). The significant difference in thermal response for 1-BuOH/2-BuOH and t-BuOH isomers was ascribed to the difference in the photocatalytic mechanisms of these species. The photothermal effect with 1-BuOH/2-BuOH isomers can be explained by the thermally induced transfer of photogenerated, shallowly trapped electron holes to highly reactive free holes at the surface of TiO2 and the further hole-mediated cleavage of the O-H bond. In the system with t-BuOH, another mechanism could also contribute to the overall process through hydrogen abstraction from the C–H bond by an intermediate OH radical, leading to CH3 group ejection. Formation of OH radicals during light irradiation of Ti@TiO2 nanoparticle suspension in water has been confirmed using terephthalate dosimetry. This analysis also revealed a positive temperature response of OH radical formation.

Graphical Abstract

1. Introduction

Environmentally friendly industry is likely to be a dominant end-use for hydrogen as the world strives to achieve net zero by 2050 [1,2]. Catalytic processes play a key role in the production of low-carbon hydrogen [3,4]. More specifically, photocatalytic hydrogen evolution from bio-based sources, such as alcohols, has a great potential for green hydrogen production and can also be used effectively for the production of value-added aldehydes, carboxylic acids, and esters [5]. Note that, although CO2 is also released in these processes, it is considered to be a part of a clean cycle when the sacrificial reagent has a biological origin [6]. Introducing heat into photocatalytic processes has attracted much attention in the last decade because it may significantly improve photoconversion efficiency [7,8]. However, the data about dominating photothermal mechanisms during H2 production are still scarce in the literature. Recent studies of the H/D kinetic isotope effect revealed that the photothermal effect in the process of reforming aqueous glycerol over a Ti@TiO2 core-shell photocatalyst is mainly attributed to the electron-hole-mediated splitting of the O-H bond from glycerol [9]. On the other hand, the dynamics of intermediates at the catalyst surface could also contribute to the overall reaction kinetics through the entropy of activation. It was concluded that this effect reduces the influence of temperature on photocatalytic hydrogen production. Herein, we focus on photothermal H2 production from aqueous solutions of butanol isomers in the presence of Ti@TiO2 nanoparticles (NPs). Numerous studies of photocatalytic alcohol reforming over bare TiO2 and noble metal-loaded TiO2 NPs in aqueous solutions, and in a gas phase, revealed the involvement of α-H atom abstraction in the limiting stage for primary and secondary alcohol isomers, followed by the formation of aldehydes or ketones, respectively [10,11,12,13]. On the other hand, the photoreforming of t-BuOH in the gas phase over TiO2 yields butyraldehyde as a primary product [14,15]. In aqueous solutions, photocatalytic degradation of t-BuOH over platinized TiO2 is accompanied by C–C bond scission and formation of an H2, CH4, C2H6 gas mixture [16]. To the best of our knowledge, the influence of butanol isomerization on photothermal H2 production in aqueous solutions has not been reported in the literature. Meanwhile, a comparative study of butanol isomers would provide new insights onto the photothermal reaction mechanism.

2. Results and Discussion

The experiments were performed using Ti@TiO2 core-shell NPs obtained by sonohydrothermal treatment (SHT) of metallic titanium Ti0 NPs in pure water. The experimental details are described in Materials and Methods. HR TEM images (Figure 1) show the quasi-spherical air-passivated Ti0 NPs (50–150 nm) before SHT treatment. These particles tend to exhibit a nanocrystalline shell composed of 10–20 nm anatase TiO2 crystals after SHT treatment. Detailed analysis of Ti@TiO2 NPs has been reported in our earlier studies [17,18,19]. It was shown that Ti@TiO2 NPs provide an extended photo response in the UV/vis/NIR spectral range due to the interband/intraband transitions of the metallic core and the bandgap of the semiconducting titanium oxide shell (3.52 ± 0.01 eV [18]), thus allowing efficient solar light harvesting. Photocatalytic activity of this material was ascribed to the effective electron-hole separation between the Ti0 core and the nanocrystalline TiO2 shell. This conclusion is supported by the fact that the non-coated Ti0 NPs do not exhibit photocatalytic activity despite strong light absorption [17,18].
Figure 2 depicts a typical profile of H2 emission obtained using a 0.5 M 1-BuOH aqueous solution irradiated using an Xe lamp in the presence of suspended Ti@TiO2 NPs. The 2-BuOH and t-BuOH solutions exhibit similar behavior. A stepwise increase in the bulk temperature causes a significant stepwise increase in H2 concentration in the outlet gas, indicating a strong photothermal effect in the studied system. However, when the light is cut off, H2 formation is not observed even at elevated temperatures, which clearly indicates the photonic origin of H2 formation. It should be noted that the strong evaporation of BuOH isomers does not allow us to study H2 photothermal production at T > 69 °C.
In the entire range of the studied temperatures, the H2 formation rate fell in the sequence of 1-BuOH >> 2-BuOH > t-BuOH as shown in Figure 3. This trend can be explained by the increased adsorption strength of primary alcohols on the metal oxide surface as compared to secondary and tertiary isomers [20]. It is worth noting that the isomers of alcohols exhibit such tendencies with many other TiO2-based photocatalysts loaded with noble metals, M/TiO2, where M = Pt, Pd, Au, Rh [21,22,23]. Table 1 summarizes the values of the apparent activation energy, Ea, calculated from Arrhenius plots obtained in the temperature range of 35–69 °C for all studied systems. Thus, it can be noticed that, despite the significant difference in reaction rate, both systems with 1-BuOH and 2-BuOH are characterized by very similar values of Ea. In addition, these values are comparable with the Ea measured previously for reforming glycerol with the same catalyst [19], which implies a similarity in the reaction mechanism for these sacrificial reagents. Earlier studies have noticed that the photothermal effect during the photolysis of glycerol with Ti@TiO2 NPs is attributed to the thermally induced transfer of photogenerated, shallowly trapped electron holes to highly reactive free holes, h+, at the surface of TiO2 [9]. Further electron-hole-mediated cleavage of the O-H bond from glycerol could lead to H2 formation. Similarly, the photothermal H2 production on Ti@TiO2 in the presence of 1-BuOH/2-BuOH isomers can be expressed by the Equations (1)–(4):
Scavenger adsorption:
Ti@TiO2 + RCH2OH/R1R2CHOH ⇄ Ti@TiO2·RCH2OH/R1R2CHOH
Photoexcitation:
Ti@TiO2·RCH2OH/R1R2CHOH + → [e−Ti@TiO2h+]·RCH2OH/R1R2CHOH
Hole-mediated oxidation:
RCH2OH/R1R2CHOH + 2h+ → RCHO/R1R2CO + 2H+
Molecular hydrogen formation:
2H+ + 2e → H2
where RCH2OH and R1R2CHOH correspond to 1-BuOH and 2-BuOH, respectively. After light absorption by the catalyst, the adsorbed RCH2OH and/or R1R2CHOH species are oxidized by photogenerated holes to form aldehydes and/or ketones and two protons. Molecular hydrogen is then obtained upon the reduction of produced H+ cations by the photogenerated electrons. It is interesting to note that the photolysis of a 1-BuOH aqueous solution in the presence of TiO2 loaded with noble metals leads to the formation of H2, CO2, C3H8 while that of 2-BuOH leads to H2, CO2, C2H6, CH4 in the gas phase, respectively [10,24]. In our case, H2 was the only gaseous photocatalytic product observed for both isomers. The absence of CO2 emissions could be related to the low adsorption of the formed aldehydes/ketones on the surface of our catalyst, thus inhibiting their further oxidation into CO2. This conclusion is supported by the absence of solution acidification during photolysis, indicating that organic acids are not formed during the process. However, for the system with t-BuOH, the formation of CH4 and C2H6 was detected at 35–52 °C, along with H2 formation, which agrees with the observation of CH3 radicals during photolysis of t-BuOH in the gas phase over TiO2 [14]. At higher temperatures, strong evaporation of t-BuOH did not allow for the proper measurements of hydrocarbons using mass spectrometry.
In order to highlight the positive effect of temperature on h+ mobility, we studied the photothermal production of OH radicals in Milli-Q water in the presence of Ti@TiO2 NPs and without the addition of any alcohols. Recently, it was reported that OH radicals can be produced by oxidation of bridged O-H groups at the surface of TiO2-based catalysts with the photogenerated h+ [25]. Herein, the kinetics of OH radical formation was studied using terephthalate dosimetry [26]. This analytical technique is based on OH-induced hydroxylation of the terephthalate (TPH) ion and measurement of the formed 2-hydroxyterephtalate (2-HTPH) using fluorescence spectroscopy. The experimental details are described in Materials and Methods. Figure 4 clearly points out the acceleration of OH radical formation expressed as 2-HTPH concentration with the increase in bulk temperature in accordance with the above-discussed photothermal mechanism of H2 production. The deceleration of 2-HTPH accumulation with irradiation time is most likely related to the photocatalytic oxidation of 2-HTPH.
On the basis of the above results, it is evident that a rather low Ea value for t-BuOH in comparison with 1-BuOH/2-BuOH isomers (Table 1) reflects the significant influence of C-C bond scission on the photothermal effect with t-BuOH. As aforementioned, the photothermal reforming of primary and secondary alcohols in the presence of TiO2-based photocatalysts deals with the mobility of photogenerated h+ [9,27,28]. In the system with t-BuOH, there are two possible reaction pathways for t-BuOH photoconversion: h+-mediated O-H bond cleavage similar to that with 1-BuOH/2-BuOH isomers (Reactions 5,6), and hydrogen abstraction from the C-H bond by the intermediate OH radical, leading to CH3 group ejection (Reactions 8,9) [16]:
[e−Ti@TiO2h+]·(CH3)3COH → [e−Ti@TiO2]·(CH3)3CO + H+
[e−Ti@TiO2] + H+ → H + Ti@TiO2
H + H → H2
TiO2-OH + h+ → TiO2 + OH
(CH3)3COH + OH → CH2=C(CH3)OH + CH3 + H2O
CH3 + H+ + e → CH4
CH3 + CH3 → C2H6
The secondary reactions (10,11) leading to the emission of methane and ethane could influence the production of H2. Actually, the Ea measured experimentally involves not only the enthalpy of activation, ΔH, but also the entropy of activation, ΔS, which is sensitive to the reorganization of the solvent network during product formation and the migration of intermediates as well. In terms of the Eyring transition state theory, the free energy of activation, ΔG, can be approximated as [29,30]:
E a Δ G = Δ H T Δ S
Hydrogen bonding in aqueous solutions induces the ordering of alcohol molecules at the surface of the catalyst. The ejection of the methyl group from t-BuOH would provide much stronger modification of the H–bonding network and, consequently, larger activation entropy than in the case of O–H bond cleavage for 1- and 2-BuOH isomers, which finally would lead to the drop of photothermal effect for t-BuOH.

3. Materials and Methods

3.1. Chemical Reagents

1-BuOH, 2-BuOH, t-BuOH (all > 99%, Sigma-Aldrich, St. Louis, MI, USA), terephthalic acid, TPH (99%, Acros Organics, Waltham, MA, USA), 2-hydroxyterephthalic acid, 2-HTPH (97%, Sigma-Aldrich), NaOH (98%, Alfa Aesar, Haverill, MA, USA), KH2PO4 (99%, Sigma-Aldrich), and Na2HPO4 (99%, Sigma-Aldrich) were used as received without further purification. All solutions were prepared using Milli-Q grade ultrapure water (18.2 MΩ·cm at 25 °C). Pure Ar (99.999%) was supplied by Air Liquid.

3.2. Catalyst Preparation

Ti@TiO2 NPs were obtained by sonohydrothermal treatment (20 kHz, 200 °C) of commercially available Ti0 NPs (American Elements) as described previously [19]. Typically, 2 g of air-passivated Ti0 NPs was dispersed in 50 mL of Milli-Q water using an ultrasonic bath. The suspension was then transferred into the sonohydrothermal reactor and heated at 200 °C (autogenic pressure P = 19 bar) under simultaneous ultrasonic treatment (f = 20 kHz, Pac = 17 W) for 3 h. After cooling, the obtained particles were recovered using centrifugation, washed with deionized water, and dried at room temperature under reduced pressure.

3.3. Photocatalytic Experiments

The photocatalytic study was performed in a thermostated glass-made gas-flow cell as shown in Figure 5. The temperature inside the reactor was controlled by an external thermostat. In a typical run, 7.8 mg of photocatalyst was dispersed ultrasonically in 65 mL of an aqueous solution of alcohol before transferring it into the photoreactor. The cell used for the photocatalytic experiments was equipped with two inlets, one to purge the gas and another to measure the temperature during photothermal treatment. The argon gas flow through the reactor was controlled by a volumetric flowmeter and kept constant at 45 mL·min−1 during the experiment. Photocatalytic experiments were carried out using the white light of an ASB-XE-175W xenon lamp. During the experiments, the lamp was placed at a distance of 8 cm away from the reactor. The light intensity delivered onto the reactor at such distance was measured with an X1-1 Optometer (Gigahertz-Optik) using UV-3710-4 (300–420 nm) and RW-3705-4 (400–1100 nm) calibrated detectors. The obtained values were normalized to the irradiated surface area, and the calculated light power was equal to 8.4 W and 0.6 W for vis/NIR and UV spectral ranges, respectively. During the photocatalytic treatment, the suspensions inside the reactor were stirred continuously, and the temperature was increased stepwise up to 69 °C. The cell used for performing the photocatalytic experiments was also equipped with an outlet connected to a bench-top magnetic sector mass spectrometer (Themo Scientific PRIMA BT, Waltham, MA, USA), allowing for the continuous online analysis of the produced gases. The H2 formation was quantified using external calibration curves prepared with standard gas mixtures in argon (Air Liquide).

3.4. Terephthalate Dosimetry

The solution for terephthalate dosimetry was prepared as follows: 332 mg of terephthalic acid was dissolved under mechanical stirring at an ambient temperature in a 500 mL buffer solution prepared with 200 mg of NaOH, 590 mg of KH2PO4, and 980 mg of Na2HPO4 dissolved in Milli-Q water. The obtained solution was then diluted to 1 L with water. Then, 7.8 mg of photocatalyst was dispersed into 65 mL of the prepared solution using an ultrasonic bath and further introduced into the photocatalytic cell. During photolysis, sample aliquots were withdrawn from the reactor every 5 min and filtered through 0.2 μm PTFE filters. The concentration of 2-hydroxyterephthalic acid formed upon the reaction of terephthalic acid with OH radicals was measured using fluorescence at 425 nm using a Fluoromax-4 device equipped with a Horiba NanoLED laser providing an excitation wavelength of 315 nm. A calibration curve has been obtained using standard solutions of 2-hydroxyterephthalic acid.

4. Conclusions

The photothermal hydrogen production from the aqueous solutions of 1-BuOH, 2-BuOH, and t-BuOH isomers studied in this work provides new insights into reaction mechanisms. The experiments were performed using innovative, noble, metal-free core-shell Ti@TiO2 nanoparticles obtained through hydrothermal oxidation of metallic titanium nanoparticles assisted by power ultrasound. We found that the isomerization of butanol strongly influences the kinetics of photocatalytic hydrogen evolution and its thermal response. The 1-BuOH isomer exhibits the highest H2 formation rate, which corroborates with the strongest adsorption of 1-BuOH at the surface of TiO2 compared to other butanol isomers. Analysis of the gaseous products revealed a significant difference between 1-BuOH/2-BuOH and t-BuOH. Photolysis of 1-BuOH/2-BuOH solutions yields solely H2 as a gaseous photocatalytic product without emission of CO2 or hydrocarbons. However, the photocatalytic degradation of t-BuOH leads to the formation of H2, CH4, and C2H6 indicating the scission of the C–C bond for tertiary isomers. A similar trend is observed in the thermal response of the photocatalytic H2 formation. The apparent activation energies, Ea, are very similar for 1-BuOH and 2-BuOH isomers (20–21 kJ·mol−1), and are also fairly close to that of glycerol measured in our earlier studies with the same catalyst. On the other hand, the Ea value for t-BuOH is much lower (13 kJ·mol−1), indicating a weak photothermal effect for this isomer. This difference has been attributed to the more complex mechanism in the case of tertiary isomers involving a large number of intermediates and the contribution of photogenerated OH radicals to the degradation of t-BuOH. Formation of OH radicals during photoexcitation of Ti@TiO2 nanoparticles was confirmed using terephtalate dosimetry.

Author Contributions

Conceptualization, S.I.N.; methodology, S.E.H. and T.C.; formal analysis, S.I.N., S.E.H. and T.C.; investigation, S.E.H. and M.B.; data curation, S.E.H. and T.C.; writing—original draft preparation, S.E.H. and M.B.; writing—review and editing, S.I.N. and S.E.H.; visualization, S.E.H. and S.I.N.; supervision, S.I.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abeab, J.O.; Popoola, A.P.I.; Ajenifuja, E.; Popoola, O.M. Hydrogen energy, economy and storage: Review and recommendation. Int. J. Hydrogen Energy 2019, 44, 15072–15086. [Google Scholar] [CrossRef]
  2. Oureshi, F.; Yusuf, M.; Kamyab, H.; Zaidi, S.; Khalil, M.J.; Khan, M.A.; Alam, M.A.; Masood, F.; Bazli, L.; Chelliapan, S.; et al. Current trends in hydrogen production, storage and applications in India: A review. Sustain. Energy Technol. Assess. 2022, 53, 102677. [Google Scholar]
  3. Voiry, D.; Shin, H.S.; Loh, K.P.; Chhowalla, M. Low-dimensional catalysts for hydrogen evolution and CO2 reduction. Nat. Rev. Chem. 2018, 2, 0105. [Google Scholar] [CrossRef]
  4. Qureshi, F.; Yusuf, M.; Kamyab, H.; Vo, D.-V.N.; Chelliapan, S.; Joo, S.-W.; Vasseghian, Y. Latest eco-friendly avenues on hydrogen production towards a circular bioeconomy: Current challenges, innovative insights, and future perspectives. Renew. Sustain. Energy Rev. 2022, 168, 112916. [Google Scholar] [CrossRef]
  5. Navarro, R.M.; Sanchez-Sanchez, M.C.; Alvarez-Galvan, M.C.; Valle, F.D.; Fierro, J.L.G. Hydrogen production from renewable sources: Biomass and photocatalytic opportunities. Energy Environ. Sci. 2009, 2, 35–54. [Google Scholar] [CrossRef]
  6. Qureshi, F.; Yusuf, M.; Pasha, A.A.; Khan, H.W.; Imteyaz, B.; Irshad, K. Sustainable and energy efficient hydrogen production via glycerol reforming technique: A review. Int. J. Hydrogen Energy 2022, 47, 41397–41420. [Google Scholar] [CrossRef]
  7. Kollmannsberger, S.L.; Walenta, C.A.; Courtois, C.; Tschurl, M.; Heiz, U. Thermal control of selectivity in photocatalytic, water-free alcohol photoreforming. ACS Catal. 2018, 8, 11076–11084. [Google Scholar] [CrossRef]
  8. Fang, S.; Hu, Y.H. Thermo-photo catalysis: A whole greater than the sum of its parts. Chem. Soc. Rev. 2022, 51, 3609–3647. [Google Scholar] [CrossRef]
  9. El Hakim, S.; Chave, T.; Nikitenko, S.I. Deciphering the reaction mechanisms of photothermal hydrogen production using H/D kinetic isotope effect. Cat. Sci. Technol. 2022, 12, 5252–5256. [Google Scholar] [CrossRef]
  10. Linsebigler, A.L.; Lu, G.Q.; Yates, J.T. Photocatalysis on TiO2 surfaces—Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  11. Panayotov, D.A.; Morris, J.R. Surface chemistry of Au/TiO2: Thermally and photolytically activated reactions. Surf. Sci. Rep. 2016, 71, 77–271. [Google Scholar] [CrossRef]
  12. Osterloh, F.E. Photocatalysis versus photosynthesis: A sensitivity analysis of devices for solar energy conversion and chemical transformations. ACS Energy Lett. 2017, 2, 445–453. [Google Scholar] [CrossRef]
  13. Fabian, D.M.; Hu, S.; Singh, N.; Houle, F.A.; Hisatomi, T.; Domen, K.; Osterloh, F.E.; Ardo, S. Particle suspension reactors and materials for solar-driven water splitting. Energy Environ. Sci. 2015, 8, 2825–2850. [Google Scholar] [CrossRef] [Green Version]
  14. Walenta, C.A.; Kollmannsberger, S.L.; Courtois, C.; Tschurl, M.; Heiz, U. Photocatalytic selectivity switch to C-C scission: α-methyl ejection of tert-butanol on TiO2(110). Phys. Chem. Chem. Phys. 2018, 20, 7105–7111. [Google Scholar] [CrossRef]
  15. Courtois, C.; Walenta, C.A.; Tschurl, M.; Heiz, U.; Friend, C.M. Regulating photochemical selectivity with temperature: Isobutanol on TiO2(110). J. Am. Chem. Soc. 2020, 142, 13072–13080. [Google Scholar] [CrossRef]
  16. Nishimoto, S.; Ohtani, B.; Shirai, H.; Kagiya, T. Photocatalytic degradation and dimerization of t-butyl alcohol by aqueous suspension of platinized titanium dioxide. J. Chem. Soc. Perkin Trans. 1986, 2, 661–665. [Google Scholar] [CrossRef]
  17. Nikitenko, S.I.; Chave, T.; Cau, C.; Brau, H.-P.; Flaud, V. Photothermal hydrogen production using noble-metal-free Ti@TiO2 core-shell nanoparticles under visible-NIR light irradiation. ACS Catal. 2015, 5, 4790–4795. [Google Scholar] [CrossRef]
  18. Nikitenko, S.I.; Chave, T.; Le Goff, X. Insights into the photothermal hydrogen production from glycerol aqueous solutions over noble metal-free Ti@TiO2 core-shell nanoparticles. Part. Part. Syst. Charact. 2018, 35, 1800265. [Google Scholar] [CrossRef]
  19. El Hakim, S.; Chave, T.; Nada, A.A.; Roualdes, S.; Nikitenko, S.I. Tailoring noble metal-free Ti@TiO2 photocatalyst for boosting photothermal hydrogen production. Front. Catal. 2021, 1, 669260. [Google Scholar] [CrossRef]
  20. López-Tendallo, F.J.; Hidalgo-Carrillo, J.; Montes-Jiménez, V.; Sánchez-López, E.; Urbano, F.J.; Marinas, A. Photocatalytic production of hydrogen from binary mixtures of C-3 alcohols on Pt/TiO2: Influence of alcohol structure. Catal. Today 2019, 328, 2–7. [Google Scholar] [CrossRef]
  21. López, C.R.; Pulido Melián, E.; Ortega Méndez, J.A.; Santiago, D.E.; Doña Rodríguez, J.M.; González Díaz, O. Comparative study of alcohols as sacrificial agents in H2 production by heterogenous photocatalysis using Pt/TiO2 catalysts. J. Photochem. Photobiol. A Chem. 2015, 312, 45–54. [Google Scholar] [CrossRef]
  22. Yasuda, M.; Matsumoto, T.; Yamashita, T. Sacrificial hydrogen production over TiO2-based photocatalysis: Polyols, carboxylic acids, and saccharides. Renew. Sustain. Energy Rev. 2018, 81, 1627–1635. [Google Scholar] [CrossRef]
  23. Kennedy, J.; Bahruji, H.; Bowker, M.; Davies, P.R.; Bouleghlimat, E.; Issarapanacheewin, S. Hydrogen generation by photocatalytic reforming of potential biofuels: Polyols, cyclic alcohols, and saccharides. J. Photochem. Photobiol. A Chem. 2018, 356, 451–456. [Google Scholar] [CrossRef]
  24. Bahruji, H.; Bowker, M.; Davies, P.R.; Pedrono, F. New insights into the mechanism of photocatalytic reforming on Pd/TiO2. Appl. Catal. B Environ. 2011, 107, 205–209. [Google Scholar] [CrossRef]
  25. Nosaka, Y.; Nosaka, A. Understanding hydroxyl radical (OH) generation process in photocatalysis. ACS Energy Lett. 2016, 1, 356–359. [Google Scholar] [CrossRef] [Green Version]
  26. Mark, G.; Tauber, A.; Laupert, R.; Schuchmann, H.-P.; Schulz, D.; Mues, A.; von Sonntag, C. OH-radical formation by ultrasound in aqueous solution. Part II: Terephtalate and Fricke dosimetry and the influence of various conditions on the sonolytic yield. Ultrason. Sonochem. 1998, 5, 41–52. [Google Scholar] [CrossRef]
  27. Bahnemann, D.W.; Hilgendorff, M.; Memming, R. Charge carrier dynamics at TiO2 particles: Reactivity of free and trapped holes. J. Phys. Chem. B 1997, 101, 4265–4275. [Google Scholar] [CrossRef]
  28. Kim, G.; Choi, H.J.; Kim, H.; Kim, J.; Monllor-Satoca, D.; Kim, M.; Park, H. Temperature-boosted photocatalytic H2 production and charge transfer kinetics on TiO2 under UV and visible light. Photochem. Photobiol. Sci. 2016, 15, 1247–1253. [Google Scholar] [CrossRef]
  29. Eyring, H. The activated complex in chemical reactions. J. Chem. Phys. 1935, 3, 107–115. [Google Scholar] [CrossRef]
  30. Jain, P.; Yu, S. Isotope effects in plasmonic photosynthesis. Angew. Chem. Int. Ed. 2020, 59, 22480–22483. [Google Scholar]
Figure 1. Typical HR TEM images of Ti0 (A) and Ti@TiO2 (B) NPs. In the inset, the distance of 0.35 nm corresponds to (101) plane of TiO2 anatase. The analysis was performed with a Jeol 2200FS (200 kV) device.
Figure 1. Typical HR TEM images of Ti0 (A) and Ti@TiO2 (B) NPs. In the inset, the distance of 0.35 nm corresponds to (101) plane of TiO2 anatase. The analysis was performed with a Jeol 2200FS (200 kV) device.
Catalysts 12 01662 g001
Figure 2. Typical temperature-dependent profile of H2 emission during photolysis of 0.5 M 1-BuOH with white light of an Xe lamp under Ar flow (45 mL·min−1) in the presence of Ti@TiO2 NPs. H2 was measured online in the outlet gas using mass spectrometry.
Figure 2. Typical temperature-dependent profile of H2 emission during photolysis of 0.5 M 1-BuOH with white light of an Xe lamp under Ar flow (45 mL·min−1) in the presence of Ti@TiO2 NPs. H2 was measured online in the outlet gas using mass spectrometry.
Catalysts 12 01662 g002
Figure 3. Rate of H2 production as a function of bulk temperature from 0.5 M aqueous solutions of BuOH isomers irradiated with an Xe lamp (vis/NIR: 8.4 W, UV: 0.6 W) under Ar flow in the presence of Ti@TiO2 NPs.
Figure 3. Rate of H2 production as a function of bulk temperature from 0.5 M aqueous solutions of BuOH isomers irradiated with an Xe lamp (vis/NIR: 8.4 W, UV: 0.6 W) under Ar flow in the presence of Ti@TiO2 NPs.
Catalysts 12 01662 g003
Figure 4. Kinetics of photocatalytic OH radical formation as a function of bulk temperature measured using terephthalate dosimetry in the presence of Ti@TiO2 NPs and in the absence of butanol isomers (Xe lamp, Ar flow).
Figure 4. Kinetics of photocatalytic OH radical formation as a function of bulk temperature measured using terephthalate dosimetry in the presence of Ti@TiO2 NPs and in the absence of butanol isomers (Xe lamp, Ar flow).
Catalysts 12 01662 g004
Figure 5. Image of the photocatalytic cell before (A) and during (B) illumination with xenon lamp.
Figure 5. Image of the photocatalytic cell before (A) and during (B) illumination with xenon lamp.
Catalysts 12 01662 g005
Table 1. The apparent activation energy of H2 photocatalytic formation for BuOH isomers (0.5 M) irradiated with an Xe lamp in the presence of Ti@TiO2 NPs under Ar flow.
Table 1. The apparent activation energy of H2 photocatalytic formation for BuOH isomers (0.5 M) irradiated with an Xe lamp in the presence of Ti@TiO2 NPs under Ar flow.
AlcoholEa, kJ·mol−1
1-BuOH21 ± 2
2-BuOH20 ± 2
t-BuOH13 ± 2
Glycerol25 ± 5 *
* Data [16].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

El Hakim, S.; Bathias, M.; Chave, T.; Nikitenko, S.I. Influence of Butanol Isomerization on Photothermal Hydrogen Production over Ti@TiO2 Core-Shell Nanoparticles. Catalysts 2022, 12, 1662. https://doi.org/10.3390/catal12121662

AMA Style

El Hakim S, Bathias M, Chave T, Nikitenko SI. Influence of Butanol Isomerization on Photothermal Hydrogen Production over Ti@TiO2 Core-Shell Nanoparticles. Catalysts. 2022; 12(12):1662. https://doi.org/10.3390/catal12121662

Chicago/Turabian Style

El Hakim, Sara, Mathéo Bathias, Tony Chave, and Sergey I. Nikitenko. 2022. "Influence of Butanol Isomerization on Photothermal Hydrogen Production over Ti@TiO2 Core-Shell Nanoparticles" Catalysts 12, no. 12: 1662. https://doi.org/10.3390/catal12121662

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