Next Article in Journal
5-Hydroxymethylfurfural and Furfural Base-Free Oxidation over AuPd Embedded Bimetallic Nanoparticles
Next Article in Special Issue
SiO2@TiO2 Composite Synthesis and Its Hydrophobic Applications: A Review
Previous Article in Journal
Liquid Phase Furfural Oxidation under Uncontrolled pH in Batch and Flow Conditions: The Role of In Situ Formed Base
Previous Article in Special Issue
Energy-Saving UHMW Polymeric Flow Aids: Catalyst and Polymerization Process Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photoreduction of a Pd-Doped Mesoporous TiO2 Photocatalyst for Hydrogen Production under Visible Light

Chemical Reactor Engineering Centre (CREC), Faculty of Engineering, Western University, London, ON N6A 5B9, Canada
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(1), 74; https://doi.org/10.3390/catal10010074
Submission received: 16 November 2019 / Revised: 26 December 2019 / Accepted: 27 December 2019 / Published: 3 January 2020
(This article belongs to the Special Issue Commemorative Issue in Honor of Professor Hugo de Lasa)

Abstract

:
Photoreduction with visible light can enhance the photocatalytic activity of TiO2 for the production of hydrogen. In this article, we present a strategy to photoreduce a palladium-doped TiO2 photocatalyst by using near-UV light prior to its utilization. A sol-gel methodology was employed to prepare the photocatalysts with different metal loadings (0.25–5.00 wt% Pd). The structural and morphological characteristics of the synthesized Pd-TiO2 were analyzed by using X-ray Diffraction (XRD), BET Surface Area (SBET), TemperatureProgrammed Reduction (TPR), Chemisorption and X-ray Photoelectron Spectroscopy (XPS). Hydrogen was produced by water splitting under visible light irradiation using ethanol as an organic scavenger. Experiments were developed in the Photo-CREC Water-II (PCW-II) Reactor designed at the CREC-UWO (Chemical Reactor Engineering Centre). It was shown that the mesoporous 0.25 wt% Pd-TiO2 with 2.5 1eV band gap exhibits, under visible light, the best hydrogen production performance, with a 1.58% Quantum Yield being achieved.

Graphical Abstract

1. Introduction

In the last few years, researchers have been investigating clean and emission-free renewable energy resources [1]. In this respect, hydrogen is an energy vector that has attracted the attention of many scientists in both academia and industry [2,3,4,5,6,7]. Hydrogen is one of the most abundant elements on earth that can be obtained from a variety of organic sources [8]. Besides its abundance, hydrogen has a high calorific value (143 MJ/kg) and clean emissions, with the combustion of hydrogen producing water as the only by-product [9,10].
Photocatalysis provides a promising method for hydrogen production [11]. This method involves photons and semiconductors that generate electron-hole pairs, facilitating water splitting [12]. Photoexcited electron-hole pairs can be separated efficiently using sacrificial agents, which allow the formation of hydrogen with reduced electron-hole pair recombination [13,14,15,16,17]. Nowadays, however, this process faces challenges in being implemented using visible light, given its low photon conversion efficiency [18].
In order to achieve photocatalytic water splitting, feasible photocatalysts must meet the following criteria: (a) They must display suitable band gaps to absorb visible light, (b) They must be chemically stable under redox conditions, (c) They must have a low cost, (d) They must be recyclable, (e) They must be chemically resistant and (f) They must be adaptable for large-scale hydrogen production [19,20].
Until today, synthesized titanium dioxide (TiO2) has been the most frequently used semiconductor due to its oxidation and chemical resistance properties, accessibility, and affordability [21]. TiO2 contains three allotropic phases: anatase, rutile, and brookite [22]. However, anatase appears to be the most desirable phase for photocatalytic hydrogen production [23]. Synthesized mesoporous semiconductors may also add significant beneficial features to photocatalysis given that they may contribute with: (a) favorable electronic structures, (b) adequate properties for light absorption, (c) sufficient electron transport properties, (d) good photo-semiconductor excitation and (e) suitable chemical species transport properties [24].
Photocatalytic reactions can be realized via sunlight irradiation [25]. TiO2 is however, limited by its 3.2 eV band gap, being able the use about 4% of the solar spectrum only. Thus, modified TiO2-based photocatalysts with smaller band gaps are required for more efficient sunlight utilization [26]. One possible approach to modify photocatalysts is by loading a noble metal as a co-catalyst. For instance, it has been proven that loading a noble metal on TiO2 is an effective method to improve its photoactivity [27]. In this respect, in prior studies, our group has investigated the effect of platinum on TiO2 under near-UV light [28,29].
Nevertheless, one can consider the use of less expensive noble metal co-catalysts with lower Fermi levels than TiO2, such as palladium. In this respect, palladium is anticipated to decrease the electron-hole recombination, leading to an energy band gap reduction [30]. Abdelaal et al. [31], for instance, prepared a Pd-TiO2 photocatalyst for methylene blue degradation under a 300 W Xenon lamp irradiation. Once the Pd metal was added, authors observed a diminished band gap of 3.04 eV. Likewise, Espino et al. [32], doped TiO2 with palladium for sodium diclofenac, isoproturon, and phenol degradation in water using a 400 W mercury lamp. They noticed a decreased band gap of 2.78 eV.
In the present study, a mesoporous Pd-TiO2 photocatalyst active under visible light is considered. A key preparation step, photoreducing the Pd-TiO2 using near-UV light is implemented to enhance its photocatalyst activity [33]. It is shown, that the near-UV light photoreduced mesoporous Pd-TiO2 displays a 2.51 eV energy band gap. The near-UV light photoreduced Pd-TiO2 photocatalyst is evaluated in a Photo-CREC Water-II Reactor (PCW-II Reactor) under visible light [34]. Ethanol is employed as an organic scavenger to help with electron-hole separation [33]. The promoted mesoporous Pd-TiO2 energy band gap reduction enhances light absorption in the visible region [35]. This leads as is reported in the present study, to enhanced hydrogen production and increased Quantum Yields (QY).

2. Results and Discussion

2.1. Photocatalyst Characterization

2.1.1. Adsorption-Desorption Isotherms

The BET-surface area (SBET), the pore diameter (Dp), and the pore volume (Vp) of the prepared photocatalysts are reported in Table 1. These mesoporous photocatalysts display an increase in the specific surface area versus the one obtained for Degussa P-25. For the TiO2 doped with palladium in amounts greater than 0.25 wt%, a decrease in the specific surface area and an increase in the average pore diameter were noticed, with this being attributed to a moderate blocking of the TiO2 pores with Pd [36].
On the other hand, and when utilizing the Barrett–Joyner–Halenda (BJH) methodology, the mesoporous pore size distribution was found to be unimodal for the 0.25 and 0.50 wt% Pd–500 °C thermally treated TiO2 photocatalysts, with pore sizes in the 18–22 nm range. However, and for the photocatalysts with Pd loadings equal or larger than 1.0 wt%, the unimodal pore size distribution evolved towards a bimodal pore size distribution, with a second peak at 16–35 nm.

2.1.2. Hydrogen Chemisorption

By employing hydrogen pulse chemisorption, it was possible to determine the fraction of dispersed Pd on the photocatalyst [37]. Specifically, by assuming that only one hydrogen molecule chemisorbs on a single Pd site, the metal percent dispersion was calculated as reported in Table 2. This shows that increasing the metal loading decreases the metal dispersion, with the metal loading remaining at a high 75% for the 0.25 wt% Pd–TiO2.

2.1.3. X-ray Diffraction (XRD)

The mass fractions of anatase and rutile were determined from the relative XRD diffraction band intensities, using an anatase sample with a 100% TiO2 crystalline phase as a reference [25]. XRD patterns obtained for the different photocatalysts, are shown in Figure 1. XRD peaks for DP25, anatase and rutile are also given in Figure 1 in order to compare them with those of the Pd-TiO2 photocatalysts. XRD peaks at 25°, 38°, 48°, 54°, 63°, 69°, 70.5° and 75° 2θ diffraction angles were assigned to anatase (101), (004), (200), (105), (204), (116), (220) and (215) crystal planes [JCPDS No. 73-1764], whereas XRD peaks at 40.12° and 46.66° were assigned to Pd (111) and (200) crystalline planes, respectively [JCPDS No. 87-0638].
Figure 2 reports a comparative analysis of the XRD diffractograms for the photocatalysts before and after reduction. One can see a peak at 34° of the 2θ angle scale, which corresponds to (002) reflections of a tetragonal palladium oxide phase [JCPDS 41-1107]. Furthermore, the peaks at 40° and 46° of the 2θ angle scale relate to the Pd° [JCPDS No. 87-0638]. Thus, there is a structural difference in the semiconductor material after the reduction process, with the absence of XRD detectable palladium oxides and the formation of metallic palladium h k l (1 1 1) and (2 0 0).
Furthermore, the average size of the crystallites was calculated based on XRD peak broadening using the Scherrer Equation. The calculated crystallite sizes were between 9 and 14 nm and they are reported in Table 3.
Additionally, the lattice constants of the tetragonal anatase unit cells were calculated for the anatase phase (h k l) = (1 0 1). The resulting a = b = 3.7779 and c = 3.4888 parameters at 25° in the 2θ angle scale, showed that pure anatase was present in the photocatalysts [38].

2.1.4. Temperature Programmed Reduction (TPR)

Temperature Programmed Reduction (TPR) is one key parameter that could influence photocatalyst performance. TPR showed that once the photocatalyst was under hydrogen flows, the palladium oxide species were reduced [39], with four TPR peaks being observed. The first negative peak at 68 °C was attributed to the decomposition of palladium β-hydride, which occurred as soon as the hydrogen and palladium came into contact, at the beginning of the analysis. This negative peak at 68 °C may be larger at higher Pd loadings in excess of 0.25 wt% [40], with this being attributed to the higher Pd dispersions. The second broad peak was assigned to palladium oxide reduction. This peak started at 200 °C and was completed at 300 °C. This broad peak was attributed to the broad palladium particle size distribution (18–22 nm) [41], with larger particle sizes moderately increasing the palladium oxide reduction temperature.
Furthermore, as reported in Figure 3, the 0.25 wt%Pd-TiO2 semiconductor showed double peaks in the 400–600 °C range, which were attributed to the Ti+4 ions surface reduction [42]. Similar trends to the ones reported in Figure 3 were found for all doped photocatalysts, with these consisting of 0.25 wt% to 5.00 wt% Pd-TiO2. It should be noted, however, that only the second peak in the 200–300 °C range was considered in all the calculations, in order to establish the amount of reducible palladium.
In summary, one can see that for the Pd-TiO2 of the present study, the Pd reduction temperature was above 350 °C, which suggests strong metal-support interactions, potentially leading to high photocatalytic hydrogen production activity [43].

2.1.5. Band Gap

Figure 4 illustrates the UV-visible absorption spectra of the Pd-TiO2 at different metal loadings while applying the Kubelka–Munk (K–M) model, following the Tauc plot methodology. When using the Kubelka–Munk (K–M) method, the bad gap is determined by drawing a straight line in the (α hv)0.5 versus (hv) plot with α representing the absorption coefficient, h being the Planck constant (6.34 × 1034 J s/photon) and v denoting the radiation frequency [44]. Following this, the location of the band gap is determined at the straight line intersection with the x-axis. Figure 4 reports the Tauc plots of the doped TiO2 with 0.25 wt%Pd, 0.5 wt%Pd,1 wt%Pd, 2.5 wt%Pd, and 5 wt%Pd loadings.
For instance, one can observe in Figure 4, that for the 0.25 wt% Pd-TiO2, a linear extrapolation yields a 2.51 eV band gap. This 2.51 eV band gap corresponds to a 494 nm photon wavelength (λ) given λ = hc/Ebg, with h being the Planck constant (6.34 × 1034 J s/photon), c being the speed of light under vacuum (3.00 × 108 m/s2), and Ebg being the electron band gap. One should note, that 2.51 eV represents a significantly reduced Ebg. with respect to the 2.99 eV band gap for mesoporous TiO2 without Pd, as reported in Table 4. It is also observed in Table 4 that using Pd loadings above 0.25 wt%, yields to a reversed trend in the Ebg, with band gaps increasing steadily instead.
In this respect, one could assign the band gap decrease of the 0.25 wt% Pd-TiO2 to the Fermi level changes. These changes can be assigned to the sp-d orbital exchange interactions between the band electrons and the localized d electrons of the Pd 3d ions substituting the Ti4+ cations. The s-d and p-d exchange interactions give rise to a downward shift of the conduction band edge and an upward shift of the valence band edge, leading to a band gap narrowing [45,46,47,48]. However, at higher than 0.25%wt Pd loadings, it is speculated that the band gap increase is due to the dominant d-d transitions over the sp-d transitions.
Thus, one can consider, that low noble metal loadings (e.g., 0.25%wt Pd-TiO2) facilitate both charge collection and light absorption [49]. Low Pd loadings give rise to localized energy levels in the band gap of the TiO2. In this case, the valence band electrons of the TiO2 are excited at wavelengths longer than 400 nm [50]. Alternatively, excessive noble metal loading may lead to smaller photocatalyst specific surface areas, with larger metal crystallites formed with PdO inclusions [51]. In this respect, XPS confirmed that the PdO presence shields incident photons, blocking light absorption and preventing the generation of semiconductor electron-hole pairs [52].

2.1.6. X-ray Photoelectron Spectroscopy (XPS)

For chemical state identification, quantitative XPS analyses were performed on: a) 0.25 wt%–5 wt% Pd-TiO2 before photoreduction and b) 0.25%Pd-TiO2 after photoreduction. The Pd 3d3/2 and Pd 3d5/2 spin orbital splitting photoelectrons were observed in both the photoreduced and non-photoreduced 0.25 wt% Pd-TiO2 photocatalysts.
Figure 5 reports the photoreduced and non-photoreduced 0.25 wt% Pd-TiO2 photocatalyst XPS peaks. XPS peaks were analyzed via band deconvolution, at the 334.54 eV, 336.38 eV, 339.69 eV, and 341.54 eV characteristic binding energies.
Table 5 reports the observed binding energies for the 0.25 wt% Pd–TiO2 with the full-widths-at-half-maximum (FWHM) and percentual areas.
Thus, one can see that according to Table 5 and Figure 5, the 0.25 wt% Pd–TiO2 photocatalyst displays in the XPS, the two most intense peaks, at 334.43 and 339.69 eV. These peaks are assigned to the metallic Pd. Furthermore, there are two other weaker recorded peaks at 336.28 and 341.54 eV, which are attributed to 3d5/2 and 3d3/2 binding energies of the PdO species.
Table 5 reports a comparison between the XPS binding energies of the 0.25 wt% Pd–TiO2 photocatalyst before and after 60 min. of near-UV photoreduction. One can see that there is a significantly increased in Pd° content ranging from 49.8% up to 81.7% after UV irradiation, with the reaming PdO amounting to 18.3% only. This increased Pd° is the result of near-UV electrons diminishing the oxidized Pd species.
Furthermore, the XPS analysis of the 0.25%wt Pd-TiO2 photocatalysts, also showed Ti 2p and O 1 s bands at peaks of 454 and 526 eV, respectively (not displayed). These bands were assigned to the titanium oxide.

2.2. Macroscopic Irradiation Energy Balance (MIEB)

The MIEB were developed in a carefully selected 6000 cm3 Photo-CREC Water-II Reactor control volume. This control volume as described in Figure 6, contained uniformly suspended Pd-TiO2 photocatalysts.
On this basis, the rate of photon absorption was calculated as [27]:
P a = P i P b s P t
where, Pa = the rate of absorbed photons, Pi = the rate of photons reaching the reactor inner surface, Pbs = the rate of backscattered photons exiting the system, and Pt = the rate of transmitted photons in Einstein/s. It is desirable to obtain a high rate of absorbed photons for photocatalytic processes. For detailed calculations of the MIEB refer to Appendix A.
The Pa (rate of absorbed photons) was calculated as shown in Table 6. One can thus see, that when a 0.25 wt% Pd-TiO2 is used, there is a significant Pa increase versus the Pa value when undoped TiO2 is utilized. However, one can also observe that larger loadings than 0.25 wt% Pd on TiO2 yield a mildly increased Pa with 1.00 wt% Pd on TiO2 giving a Pa maximum. Larger than 1.00 wt% Pd on TiO2 yield a modest Pa decrease, with a consequently diminishing absorption efficiency.
In summary, TiO2 photocatalysts doped with Pd considerably augment the absorbed visible irradiation photons, and this is the case when compared to the undoped TiO2.

2.3. Hydrogen Production

The Pd-TiO2 photocatalysts of the present study were evaluated in the PCW-II with respect to their ability to enhance hydrogen production, under the following conditions: (a) by utilizing a photocatalyst concentration of 0.15 g/L, (b) by using a 2.0 v/v% of ethanol as organic scavenger and (c) by utilizing a pH = 4 ± 0.05.
Figure 7 reports the cumulative hydrogen volume produced, using TiO2 doped with different Pd loadings under visible light. It was shown that Pd-TiO2 semiconductors consistently enhance hydrogen production, with the best performance obtained with the 0.25 wt% Pd. This mesoporous semiconductor may have an increased scavenging effect of photogenerated electrons and therefore, prevent electron–hole pairs recombination [53].
Furthermore, it was also observed that when the metal loading of the Pd-TiO2, was augmented to 0.5 wt% Pd and above, a decreased rate of hydrogen production was obtained. One should notice as well, that the MIEB as reported in Table 6, showed relatively stabilized visible light absorption for various Pd doped TiO2 photocatalysts. Thus, one can conclude that hydrogen production differences cannot be assigned to changes in electron and hole pairs generation [54], but rather to a most effective trapping of electrons in 0.25 wt%Pd-TiO2 than in photocatalysts with larger Pd loadings.
Figure 6 also displays that after only 6 h of visible light irradiation, a maximum volume of 4.7 cm3 STP (standard temperature and pressure) of hydrogen is obtained when using the 0.25 wt% Pd on TiO2. This volume is approximately 4 times higher than the volume produced with undoped mesoporous TiO2.

2.3.1. Precursor Near UV-Light Photoreduction

Palladium is present in a metallic state during the sol-gel photocatalyst preparation. However, palladium can be oxidized during the photocatalyst precursor calcination preparation step. Confirmation of this is given by the X-ray Diffraction Analysis where at 34° (111) of the 2θ angle scale, there is indication of the presence of PdO. As well, the XPS also shows that 50.2% of palladium is present as PdO after photocatalyst precursor calcination as reported in Table 5. Thus and on this basis, one can conclude that palladium species on the semiconductor requires further reduction, to ensure that a substantial amount of palladium species is present as Pd°.
Therefore, a special and additional photocatalyst pretreatment was implemented in the present study, to ensure that most palladium was appropriately reduced to Pd°. Pd° promotes the high photocatalytic activity of TiO2, by generating a Schottky junction between the metal and the photocatalyst. The metal particles trap and store the photogenerated electrons, reducing the rate of the electron hole recombination [55].
With this end and as described in Figure 8, a 15W BLB UV-Lamp was employed to irradiate the prepared semiconductor during 1 h.
This PdO photoreduction using near UV light can be described as per Equation (2), with the resulting palladium being present as Pd° on the TiO2 structure.
PdO +2 e → Pd° E° = 0.915 V
It is speculated that photoreduction as per Equation (2), is a very efficient process with photogenerated electrons migrating from the outer TiO2 particle surface to the TiO2 mesoporous inner surface. Formed electrons can reduce the PdO into Pd° [56].
Regarding the present studies, following photoreduction, the near-UV lamp was replaced by a visible light lamp. It was then observed that when the photocatalyst was photoreduced prior to its utilization, this led to an important increase in hydrogen production, under visible light irradiation.
Figure 9 displays an enhanced cumulative hydrogen production under visible light in the Photo-CREC Water Reactor II at different loadings (0.25, 0.50, 1.00, 2.50 and 5.00 wt%). It is interesting to see that the same consistent trends were observed in previous studies of our research team, using near-UV light [57]. It has to be noted as well, that the lower the Pd loadings, the higher the hydrogen production.
Furthermore, when comparing Figure 7 and Figure 9, it can be observed that the photoreduced Pd-TiO2 phtocatalyst displayed significantly increased hydrogen production rates. Particularly for the 0.25 wt% Pd-TiO2 after photoreduction, the maximum hydrogen volume produced was 8.0 cm3 STP. This is equivalent to a 1.7 times increased hydrogen production rate. At this low palladium loading, a good metal dispersion of 75% was also observed, according to the chemisorption studies, with a slightly decreased surface area and average pore size [58].
In contrast, when using the 0.50 and 5.00 wt% Pd-TiO2, lower metal dispersions were observed, with larger metal crystallite sizes being detected. This was in line with their diminished photocatalytic hydrogen production activity [59]. To explain these results, one can consider that under visible light and using a Pd-TiO2 photocatalyst, photons are both absorbed and scattered. The MIEB as reported in Section 2.2, showed that higher Pd loadings (2.5 and 5.0 wt% Pd) do not enhance the absorption of visible light significantly. This phenomenon can be assigned to the presence of larger metal crystallites and TiO2 particle agglomerates. This limits photons from reaching the Pd° active metallic sites and from being absorbed [60]. The opposite of this was observed at lower than 1.00 wt% Pd loadings, where the photon absorption increases, positively impacting the semiconductor photoactivity.
On the other hand, it can be hypothesized as well, that the less effective photoreduction of palladium may occur for 2.5 and 5.0 wt% Pd on TiO2, due to the oversupply of noble metal. In this case, layers of PdO could still be present on the TiO2, shielding the TiO2 from light absorption. The formation of such sites could increase the photocatalyst reflectivity leading to visible light scattering [61]. As well, this phenomenon could also be attributed to the partial blocking of semiconductor pores, which may decrease the TiO2 specific surface area, as reported in Table 1.
Given that hydrogen is produced under visible light, the photogenerated holes created by the noble metal react with the organic scavenger ethanol forming byproducts such as acetaldehyde, ethane, CO2 and methane, in a progressive increment as is shown in Figure 10.

2.4. Quantum Yield (QY) Evaluation

For the photoconversion of organic species, one can consider a Quantum Yield (QY) parameter. The QY calculation methodology is reported in Appendix D.

Effect of Pd Addition on Quantum Yields

The QY calculation for the TiO2 photocatalysts calculation of the Pt transmitted photons, the Pi incident photons, and the Pbs backscattered photons, as described in Section 2.2.
Table 7 and Figure 11 report the QY% for the mesoporous photocatalysts doped with palladium at different metal loadings (0.25, 0.50, 1.00, 2.50, and 5.00 wt%) under the following conditions: (a) Photocatalyst slurry concentrations of 0.15 g/L, (b) 2.0 v/v% ethanol, (c) pH = 4 ± 0.05 and (d) Visible light.
One can observe in Table 6 that for all the prepared Pd-TiO2 photocatalysts, the QYs% obtained while being irradiated with visible light were in the 0.10–1.13% low range. These low QYs% were assigned to the lack of ability of the Pd-TiO2 photocatalysts to produce hydrogen under visible light with only 49.8 wt% of the loaded palladium as Pd°.
However, Table 6 also shows that when the photocatalysts were photoreduced with near-UV irradiation, the QYs% increased, reaching QY% values as high as 1.58%. These results demonstrate the importance of the noble metal photoreduction using near-UV irradiation (photoreduction). One should note that photoreduction was critical to making Pd-TiO2 photocatalysts active under visible light for hydrogen production.
Few papers in the technical literature report QYs of a comparable magnitude. Some authors have used Pd-TiO2 (0.3 wt%) and 3NbTi/Pt-Pd as photocatalysts, achieving maximum QYs between 0.43 and 1.2% [62,63,64]. As well, the absorbed photon rates were measured and calculated using a numerical solution of a radiation equation. These modeled absorption rates may involve significant errors. In this case, as stated before in the present study, the QYs% were determined by using an experimentally evaluated Pa, determined from MIEB macroscopic balances, as reported in Section 2.2.
Furthermore, Figure 12 reports the consistent QY% trends observed for Pd-doped TiO2 photocatalysts as follows: (a) During the first hour of irradiation, the QY% increased progressively until they reached a stable value; and (b) During the six hours of irradiation that followed, the QY% remained unchanged, with the photocatalysts under study exhibiting a stable performance. It can also be observed that there is a significant increase of QY% when using 0.25 and 0.50 wt% Pd-TiO2, whereas higher Pd loadings led to a decrease in the QY%.
It is also interesting to see in Figure 12, that the photoreduced semiconductors of the present study display good and stable QYs%, showing their significant ability to produce hydrogen. This photocatalyst stability was also established with 4 consecutive hydrogen production photocatalytic runs, each lasting 6 h or the equivalent of 24 h under visible light irradiation.
Based on the reported results, it is thus anticipated, that further research with mesoporous TiO2 photocatalysts doped with noble metals will be valuable, particularly in the case of Pd-doped TiO2 using ethanol as a scavenger. This will likely lead to stable and efficient photocatalytic processes for hydrogen production via water splitting. These Pd-TiO2 photocatalysts may also present significant cost advantages versus other noble metal dopants such as Ta, Nb or Pt.

3. Experimental Methods

The photocatalyst of the present study was synthesized using a sol-gel methodology [53]. Different techniques were utilized for its characterization as follows: (a) Specific Aurface Area (BET), (b) Chemisorption, (c) X-ray Diffraction, (d) TPR Analysis, (e) XPS and (f) UV-Vis Absorption. These techniques allowed the determination of the surface area, the pore size distribution and the pore size, the phase composition, the band gap, the metal dispersion, the temperature-programmed reduction and the Pd° crystallite size of the photocatalyst.

3.1. Photocatalyst Synthesis

The synthesis of the photocatalyst of the present study via the sol–gel method requires the following: (a) ethanol USP (C2H5OH) obtained from commercial alcohols, (b) hydrochloric acid (HCl, 37% purity), (c) pluronic F-127, (d) anhydrous citric acid, (e) titanium (IV) isopropoxide, and (f) palladium (II) chloride (PdCl2, 99.9% purity). All the reagents were purchased from Sigma Aldrich (Oakville, Ontario, Canada). Detailed information about the photocatalyst synthesis methodology is reported in Appendix B.
According to the methodology proposed by Das et al. [65], 20 mL ethanol was acidified with 1.65 g of hydrochloric acid. This was followed by the addition of 1 g of copolymer pluronic F-127. The mixture was stirred for 30 min until complete dissolution was reached. 0.315 g of anhydrous citric acid was mixed with 1 mL of water and added to the initial solution. The resultant mixture was stirred for 2 h. Following this, 1.42 g of titanium IV isopropoxide in ethanol added dropwise [29].
Furthermore, and after this, PdCl2 was added to the resulting sol-gel suspension, given its solubility under the conditions of the sol-gel suspension. PdCl2 was incorporated at different concentrations to achieve 0.25–5.00 wt% Pd loadings. The resulting sol–gel suspension was stirred for 24–48 h and then calcined under an air atmosphere at 500 °C for 6 h [66]. This allowed the copolymer to be removed, with an ordered mesoporous titanium framework being formed [67,68].
However, and considering that palladium oxidizes during the calcination step, the resulting photocatalyst had to be reduced in a subsequent step. To accomplish this, the synthesized semiconductor was placed in a flow reactor unit under an atmosphere of 1 cm3/s of Ar/H2 (g) (90/10%, Praxair) at 500 °C for 3 h [69]. Given that this reduction with hydrogen was incomplete, a further and critical Pd-TiO2 photoreduction step was implemented in the Photo-CREC Water-II (PCW-II), by exposing the photocatalyst to near-UV light at room temperature for 60 min.

3.2. Equipment

The Photo-CREC Water-II (PCW-II) Reactor is an innovative unit for hydrogen production. It is a slurry batch reactor with a total volume of 6 L. As seen in Figure 13, it is composed of: (1) A 15-W fluorescent visible light lamp, (2) A Pyrex glass inner tube where the lamp is placed, (3) A black polyethylene outer tube, (4) A centrifugal pump, (5) Two sampling ports where the photocatalyst suspension is always kept sealed under agitation, one for the liquid phase and the other one for the gas phase, (6) A hydrogen storage tank, and (7) Silica windows for irradiation measurements [70]. Refer to Appendix C for a detailed lamp characterization.

3.3. Photocatalyst Characterization

Nitrogen adsorption and desorption isotherms at −195 °C were measured using a Micrometrics ASAP 2010 Surface Area and Porosity analyzer (Norcross, GA, U.S.A). The samples were initially degassed in a vacuum at 300 °C for 3 h. The BET-surface area was determined from the BET plot. The pore volume was measured at a relative pressure of P/Po = 0.99. To estimate the pore size distribution, the N2 physisorption BJH (Barrett–Joyner–Halenda) method was used [71].
Pulse chemisorption with the Micromeritics AutoChem II Analyzer (Norcross, GA U.S.A) was utilized to determine the fraction of dispersed metal. On the other hand, the XRD spectrum for each material was measured in a Rigaku Rotating Anode X-ray Diffractometer (Rigaku, Auburn Hills, MI, United States) rated at 45 kV and 160 mA, to identify the crystalline phases of the materials. The diffractograms were taken in the 2θ angle scale (20°–80° range), with a step size of 0.02° and a dwell time of 2 s/step.
The H2 Temperature Programmed Reduction (TPR) analyses of the Pd-TiO2 photocatalysts were evaluated in a Micromeritics AutoChemII Analyzer. 250 mg of the photocatalyst were placed in the U-tube with a gas reduction mixture of Ar/H2 (g) (90/10%). The reaction temperature was kept within a 0° to 600 °C range with a flow rate of 50 mL min−1. The amount of H2 consumed during the reduction was measured by a thermal conductivity detector (TCD) [42].
To determine the band gap, a UV-VIS-NIR Spectrophotometer (Shimadzu UV-3600, Nakagyo-ku, Kyoto, Japan) was used, with the BaSO4 as a reference [48]. By using the Kubelka–Munk (K–M) methodology, Tauc plots were developed, in order to establish the corresponding band gaps for each photocatalyst [72]. Furthermore, to ascertain the composition and the oxidation/reduction state of palladium, an X-ray Photoelectron Spectroscopy (XPS) analysis was utilized [73].
Furthermore, to measure the irradiation of the visible light lamp inside the Photo-CREC Water-II, the Stellar Net EPP2000-25 Spectrometer (StellarNet Inc., StellarNet, Inc.,Tampa, Florida, U.S.A.) was used. The light source used was a fluorescent mercury Philips lamp (15 W) with an output power of 1.48 W and an average emitted photon energy of 274.5 kJ/photon mole.

3.4. Hydrogen Production

The Photo-CREC Water-II Reactor was used to evaluate the Pd-TiO2 photocatalyst utilized during the water splitting runs. 6000 mL runs of water were used as the main reagent. As well, ethanol served as an organic scavenger. The pH was set to 4 ± 0.05 with H2SO4 [2M]. A BLB UV lamp was turned on for 60 min to photoreduce the PdO present in the semiconductor. After the first hour, the UV lamp was replaced by the Phillips Visible light lamp to initiate the reaction. Then, the water splitting run continued during (6) hours of irradiation [69].
The photocatalyst was loaded at a concentration of 0.15 g/L and sonicated for 10 min to avoid particle agglomeration. This ensured a homogeneous distribution of the photocatalyst throughout the reactor. To avoid any undesirable oxidation reaction, an argon flow was circulated for 10 min to keep an inert atmosphere.
The gases produced were evaluated by making use of a Shimadzu GC2010 Gas Chromatograph (Nakagyo-ku, Kyoto, Japan), which took consecutive gas samples every hour for 6 h. Argon (Praxair 99.999%) was used as a gas carrier. The GC possessed 2 detectors: a Flame Ionization Detector (Nakagyo-ku, Kyoto, Japan) (FID) and a Thermal Conductivity Detector (TCD). This unit also included a HayeSepD 100/120 mesh packed column (9.1 m × 2 mm × 2 μm nominal SS) (Kenilworth, NJ, U.S.A.). This equipment is capable of detecting hydrogen (H2), carbon monoxide (CO), carbon dioxide (CO2), and methane (CH4) among other hydrocarbon organic species.

4. Conclusions

Palladium-doped TiO2 photocatalysts, photoreduced with near-UV irradiation prior to their utilization, are shown to favor hydrogen production under visible light. The structural properties of the best performing photocatalysts (0.25 wt%Pd-TiO2) were established using X-ray Diffraction (XRD), BET Surface Area (SBET), Temperature Programmed Reduction (TPR), Chemisorption and X-ray Photoelectron Spectroscopy (XPS), and UV-Vis Reflectance Spectroscopy. It was shown that the near-UV photoreduced 0.25 wt%Pd-TiO2 enhances hydrogen formation, reaching in 6 h, 8.0 cm3 STP of produced hydrogen. The 0.25 wt% Pd on TiO2 also displays in all cases, a zero-order kinetics, with a highest QY% of 1.58%. These positive findings set favorable prospects for further research, including the development of a kinetic model which’s availability is considered critical for the scaling up of a Photo-CREC-Water II unit for hydrogen production.

Author Contributions

Conceptualization, investigation and supervision, H.d.L.; Proposed methodology and supervision, S.E.; Validation, formal analysis and writing, B.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Sciences and Engineering Research Council of Canada (NSERC) and the University of Western Ontario, through grants awarded to Hugo de Lasa.

Acknowledgments

We would like to gratefully thank Florencia de Lasa who assisted with the editing of this paper and the drafting of the graphical abstract of the present article.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

CO2Carbon dioxide
CH4Methane
C2H6Ethane
C2H4OAcetaldehyde
cSpeed of light (3.0 × 108 m/s)
DpPore diameter (cm)
e−Electron
h+Hole
hPlanck’s constant (6.63 × 1034 J/s)
EbgEnergy band gap (eV)
EavAverage energy of a photon (kJ/mol photon)
F-127Poly (ethylene oxide)/poly (propylene oxide)/poly (ethylene oxide)
H•Hydrogen radical
H2OWater
I(λ)Intensity of light (W/cm2)
OHHydroxide ions
OHHydroxide radicals
P-123Poly (ethylene glycol)-block-poly (propylene glycol)-block-poly (ethylene glycol)
P0Rate of photons emitted by the BLB lamp (einstein/s)
PaRate of absorbed photons (einstein/s)
Pa-wallRate of photons absorbed by the inner pyrex glass (einstein/s)
PbsRate of backscattered photons exiting the system (einstein/s)
PdPalladium
PdCl2Palladium II chloride
PEOPoly (ethylene oxide)
PfsRate of forward-scattered radiation (einstein/s)
PiRate of photons reaching the reactor inner surface (einstein/s)
PnsRate of transmitted non-scattered radiation (einstein/s)
PPOPoly (propylene oxide)
PtRate of transmitted photons (einstein/s)
PtPlatinum
q (θ, z, λ, t)Net radiative flux over the lamp emission spectrum (μW/cm2)
tTime (h)
TiO2 Titanium dioxide
VTotal volume of the gas chamber (5716 cm3)
VPpore volume
WWeight (g)
Wt%Weight percent (% m/m)
Greek symbols
θDiffraction angle, also scattering angular angle (o)
λWave length (nm)
φQuantum Yield Efficiency (%)
Acronyms
BgBand Gap
BJHBarrett–Joyner–Halenda Model
BLBBlack Light Blue Lamp
BETBrunauer–Emmett–Teller Surface Area Method
CBConduction Band
DP25Degussa P25 (TiO2)
EISAEvaporation-Induced-Self-Assembly
FIDFlame Ionization Detector
JCPDSInternational Centre for Diffraction Data
K-MKubelka-Munk
MIEBMacroscopic Irradiation Energy Balance
PCW-IIPhoto-CREC Water II Reactor
PCPhotocatalyst Concentration
STPStandard Temperature and Pressure (273 K and 1 atm)
TPRTemperature Programmed Reduction
TCDThermal Conductivity Detector
UVUltraviolet
VBValence Band
XPSX-ray Photoelectron Spectroscopy
XRDX-ray Diffraction

Appendix A. Macroscopic Irradiation Energy Balance (MIEB)

The various terms in Equation (1) can be calculated as follows:
(a)
Pi is the rate of photons reaching the slurry suspension:
P i = P 0 P a w a l l   ( Einstein / s )
with P0, the rate of photons emitted by the lamp being estimated from radiometric measurements; where q (θ, z, λ) is the radiative flux expressed in J s−1 m−3, where λ is the photon wavelength expressed in nm, where r represents the radial coordinate expressed in m, where z stands for the axial coordinate expressed in m, where h represents the Planck’s constant expressed in J s, and where c denotes the speed of light expressed in m s−1.
P 0 = λ 1 λ 2 λ 0 L 0 2 π q ( θ , Z , λ ) r   d   θ   d   z   d   λ
And Pa-wall is the rate of photons absorbed by the inner Pyrex glass surface.
(b)
Pbs represents the difference between Pi and Pt/c→0+. It is the rate of photons transmitted at a photocatalyst concentration approaching zero [28] as follows:
P b s = P i P t c 0 +
(c)
Pt accounts for the difference between the transmitted non-scattered radiation (Pns) and the forward-scattered radiation (Pfs):
P t = P n s P f s
The described macroscopic balances were originally developed for near-UV irradiation in photocatalytic reactors [70]. However, given that these macroscopic balances are not photon wavelength dependant, they were extended to visible light irradiation.

Appendix B. Photocatalyst Synthesis—EISA Method

In this research, the photocatalyst synthesis uses the Evaporation-Induced-Self-Assembly (EISA) Method. This approach has been utilized in the preparation of semiconductors when adding a template to control the material pore size [74]. This methodology was applied in the present study, given that: (a) It is easy to use, (b) it allows close control of the pore size distribution with the selected copolymer template, (c) it enhances the specific surface area, (d) It is relatively inexpensive, and e) it does not require special equipment [75]. Figure A1 describes the photocatalyst synthesis based on a sol-gel method with the use of a polymeric template [pluronic F-127 (PEO106PPO70PEO106)], a metallic precursor [titanium IV isopropoxide] and an active metal [PdCl2] precursor. The mixing of these components leads to the formation of micelles and subsequently to a mesoporous structure. Subsequent calcination and reduction yield: (a) An adequate template removal, and (b) The formation of mesopores with highly dispersed palladium nanoparticles [76].
Figure A1. Description of the Four Consecutive Steps during Pd-Doped TiO2 Photocatalyst Preparation: (a) Ethanol enters the interface between the hydrophilic (PEO) and hydrophobic (PPO) chains and attaches to the hydrophobic core, (b) Ahydrophobic block (PPO) is placed in a central location surrounded by the PEO tails forming micelles, (c) The PEO tails become attached to the TiO2, leaving a hybrid outer layer with a dominant TiO2 composition, and (d) Calcination of the resulting photocatalyst precursor yields TiO2 with a 3D mesoporous structure [57].
Figure A1. Description of the Four Consecutive Steps during Pd-Doped TiO2 Photocatalyst Preparation: (a) Ethanol enters the interface between the hydrophilic (PEO) and hydrophobic (PPO) chains and attaches to the hydrophobic core, (b) Ahydrophobic block (PPO) is placed in a central location surrounded by the PEO tails forming micelles, (c) The PEO tails become attached to the TiO2, leaving a hybrid outer layer with a dominant TiO2 composition, and (d) Calcination of the resulting photocatalyst precursor yields TiO2 with a 3D mesoporous structure [57].
Catalysts 10 00074 g0a1

Appendix C. Lamp Characterization

Figure A2 reports the spectrum of the polychromatic BLB Ushio near-UV lamp, with an observed output power of 1.61 W and a 325.1 kJ/mole of photons average.
Figure A2. Near-UV Lamp Irradiation Spectrum.
Figure A2. Near-UV Lamp Irradiation Spectrum.
Catalysts 10 00074 g0a2
Figure A3 reports the spectrum of the mercury Philips visible light lamp. It has an output power of 1.48 W and an average emitted photon energy of 274.5 kJ/photon mole.
Figure A3. Visible Lamp Irradiation Spectrum.
Figure A3. Visible Lamp Irradiation Spectrum.
Catalysts 10 00074 g0a3
The axial distribution of the radiative flux was determined using the Stellar Ney EPP2000-25 Spectrometer. Figure A4 displays the observed axial visible lamp radiation distribution with the radiation significantly decreasing, in the lower and upper axial lamp positions [77].
Figure A4. Visible Lamp Axial Distribution.
Figure A4. Visible Lamp Axial Distribution.
Catalysts 10 00074 g0a4

Appendix D. Quantum Yield Calculation

The QY is defined as the formed H· molar rate over the absorbed photons molar rate. Thus, it describes the extent of the photochemical utilization for hydrogen production of absorbed photons [34]. The QY is calculated according to the following equation:
Q Y H = m o l e s   o f   H / s m o l e s   o f   p h o t o n s   a b s o r b e d   b y   t h e   p h o t o c a t a l y s t / s
Equation (A4) is equal to:
% Q Y = d N H d t P a × 100
where, d N H d t is the rate of moles of hydrogen radicals formed at any time during the photocatalyst irradiation.
To calculate the QY, the Pa or the moles of absorbed photons is needed. The Pa was determined by using the MIEB in the Photo-CREC Water-II Reactor [69], as described in Section 2.2.

References

  1. Chiarello, G.L.; Dozzi, M.V.; Selli, E. TiO2-based materials for photocatalytic hydrogen production. J. Energy Chem. 2017, 26, 250–258. [Google Scholar] [CrossRef]
  2. Higashi, M.; Domen, K.; Abe, R. Highly Stable Water Splitting on Oxynitride TaON Photoanode System under Visible Light Irradiation. J. Am. Chem. Soc. 2012, 134, 6968–6971. [Google Scholar] [CrossRef] [PubMed]
  3. Ishikawa, A.; Takata, T.; Kondo, J.N.; Hara, M.; Kobayashi, H.; Domen, K. Oxysulfide Sm2Ti2S2O5 as a Stable Photocatalyst for Water Oxidation and Reduction under Visible Light Irradiation (λ ≤ 650 nm). J. Am. Chem. Soc. 2002, 124, 13547–13553. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, D.; Pierre, A.; Kibria, M.G.; Cui, K.; Han, X.; Bevan, K.H.; Guo, H.; Paradis, S.; Hakima, A.-R.; Mi, Z. Wafer-Level Photocatalytic Water Splitting on GaN Nanowire Arrays Grown by Molecular Beam Epitaxy. Nano Lett. 2011, 11, 2353–2357. [Google Scholar] [CrossRef]
  5. Abe, R.; Sayama, K.; Sugihara, H. Development of New Photocatalytic Water Splitting into H2 and O2 using Two Different Semiconductor Photocatalysts and a Shuttle Redox Mediator IO3-/I-. J. Phys. Chem. B 2005, 109, 16052–16061. [Google Scholar] [CrossRef]
  6. Zou, Z.; Ye, J.; Sayama, K.; Arakawa, H. Direct splitting of water under visible light irradiation with an oxide semiconductor photocatalyst. Nature 2001, 414, 625–627. [Google Scholar] [CrossRef]
  7. Zuo, F.; Wang, L.; Wu, T.; Zhang, Z.; Borchardt, D.; Feng, P. Self-Doped Ti 3+ Enhanced Photocatalyst for Hydrogen Production under Visible Light. J. Am. Chem. Soc. 2010, 132, 11856–11857. [Google Scholar] [CrossRef]
  8. Jensen, S.H.; Larsen, P.H.; Mogensen, M. Hydrogen and synthetic fuel production from renewable energy sources. Int. J. Hydrogen Energy 2007, 32, 3253–3257. [Google Scholar] [CrossRef]
  9. Mazloomi, K.; Gomes, C. Hydrogen as an energy carrier: Prospects and challenges. Renew. Sustain. Energy Rev. 2012, 16, 3024–3033. [Google Scholar] [CrossRef]
  10. Ipsakis, D.; Voutetakis, S.; Seferlis, P.; Stergiopoulos, F.; Elmasides, C. Power management strategies for a stand-alone power system using renewable energy sources and hydrogen storage. Int. J. Hydrogen Energy 2009, 34, 7081–7095. [Google Scholar] [CrossRef]
  11. Dozzi, M.V.; Selli, E. Doping TiO2 with p-block elements: Effects on photocatalytic activity. J. Photochem. Photobiol. C Photochem. Rev. 2013, 14, 13–28. [Google Scholar] [CrossRef]
  12. Liu, G.; Wang, L.; Yang, H.G.; Cheng, H.-M.; Lu, G.Q.M. Titania-based photocatalysts—Crystal growth, doping and heterostructuring. J. Mater. Chem. 2010, 20, 831–843. [Google Scholar] [CrossRef]
  13. Mills, A. An overview of semiconductor photocatalysis. J. Photochem. Photobiol. A Chem. 1997, 108, 1–35. [Google Scholar] [CrossRef]
  14. Abe, R. Significant effect of iodide addition on water splitting into H2 and O2 over Pt-loaded TiO2 photocatalyst: Suppression of backward reaction. Chem. Phys. Lett. 2003, 371, 360–364. [Google Scholar] [CrossRef]
  15. Mills, A. Photosensitised dissociation of water using dispersed suspensions of n-type semiconductors. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1982, 12, 3659–3669. [Google Scholar] [CrossRef]
  16. López, C.R.; Melián, E.P.; Ortega Méndez, J.A.; Santiago, D.E.; Rodríguez, J.M.; González Díaz, O. Comparative study of alcohols as sacrificial agents in H2production by heterogeneous photocatalysis using Pt/TiO2 catalysts. J. Photochem. Photobiol. A Chem. 2015, 312, 45–54. [Google Scholar] [CrossRef]
  17. Galińska, A. Photocatalytic Water Splitting over Pt−TiO2 in the Presence of Sacrificial Reagents. Energy Fuels 2005, 19, 1143–1147. [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. Acar, C. Review of photocatalytic water-splitting methods for sustainable hydrogen production. Energy Res. 2016, 40, 1449–1473. [Google Scholar] [CrossRef]
  20. Cushing, S.K.; Li, J.; Meng, F.; Senty, T.R.; Suri, S.; Zhi, M.; Li, M.; Bristow, A.D.; Wu, N. Photocatalytic activity enhanced by plasmonic resonant energy transfer from metal to semiconductor. J. Am. Chem. Soc. 2012, 134, 15033–15041. [Google Scholar] [CrossRef]
  21. Melián, E.P.; López, C.R.; Méndez, A.O.; Díaz, O.G.; Suárez, M.N.; Rodríguez, J.M.D.; Navío, J.A.; Heviaac, D.F. Hydrogen production using Pt-loaded TiO2 photocatalysts. Int. J. Hydrogen Energy 2013, 38, 11737–11748. [Google Scholar] [CrossRef]
  22. Olivo, A.; Ghedini, E.; Signoretto, M.; Compagnoni, M.; Rossetti, I. Liquid vs. Gas Phase CO2 Photoreduction Process: Which Is the Effect of the Reaction Medium? Energies 2017, 10, 1394. [Google Scholar] [CrossRef] [Green Version]
  23. Castro, A.L. Synthesis of anatase TiO2 nanoparticles with high temperature stability and photocatalytic activity. Solid State Sci. 2008, 10, 602–606. [Google Scholar] [CrossRef]
  24. Holloway, P.; McGuire, G. Handbook of Compound Semiconductor; Noyes Publications: Park Ridge, NJ, USA, 1995. [Google Scholar]
  25. Ahmad, H.; Kamarudin, S.K.; Minggu, L.J.; Kassim, M. Hydrogen from photo-catalytic water splitting process: A review. Renew. Sustain. Energy Rev. 2015, 43, 599–610. [Google Scholar] [CrossRef]
  26. Yang, J. Roles of Cocatalysts in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef]
  27. Hou, L.; Zhang, M.; Guan, Z.; Li, Q.; Yang, J. Effect of platinum dispersion on photocatalytic performance of Pt-TiO2. J. Nanopart. Res. 2018, 20, 60. [Google Scholar] [CrossRef]
  28. Escobedo, S.; Serrano, B.; Calzada, A.; Moreira, J.; De Lasa, H. Hydrogen production using a platinum modified TiO2 photocatalyst and an organic scavenger. Kinetic modeling. Fuel 2016, 181, 438–449. [Google Scholar] [CrossRef]
  29. Guayaquil-Sosa, J.F.; Serrano-Rosales, B.; Valadés-Pelayo, P.J.; de Lasa, H. Photocatalytic hydrogen production using mesoporous TiO2 doped with Pt. Appl. Catal. B Environ. 2017, 211, 337–348. [Google Scholar] [CrossRef]
  30. Etacheri, V.; di Valentin, C.; Schneider, J.; Bahnemann, D.; Pillai, S.C. Visible-light activation of TiO2 photocatalysts: Advances in theory and experiments. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 1–29. [Google Scholar] [CrossRef] [Green Version]
  31. Abdelaal, M.Y.; Mohamed, R.M. Novel Pd/TiO2 nanocomposite prepared by modified sol–gel method for photocatalytic degradation of methylene blue dye under visible light irradiation. J. Alloys Compd. 2013, 576, 201–207. [Google Scholar] [CrossRef]
  32. Espino-Estévez, M.R.; Fernández-Rodríguez, C.; González-Díaz, O.M.; Araña, J.; Espinós, J.P.; Ortega-Méndez, J.A.; Doña-Rodríguez, J.M. Effect of TiO2–Pd and TiO2–Ag on the photocatalytic oxidation of diclofenac, isoproturon and phenol. Chem. Eng. J. 2016, 298, 82–95. [Google Scholar] [CrossRef]
  33. Wahyuni, E.T.; Kuncaka, A.; Sutarno, S. Application of Photocatalytic Reduction Method with TiO2 for Gold Recovery. Chemistry 2015, 3, 207–211. [Google Scholar]
  34. Lasa, H.D.; Rosales, B.S.; Moreira, J.; Valades‐Pelayo, P. Efficiency Factors in Photocatalytic Reactors: Quantum Yield and Photochemical Thermodynamic Efficiency Factor. Chem. Eng. Technol. 2016, 39, 51–65. [Google Scholar] [CrossRef]
  35. Endang, H.A.; Nurul, T.W. Photoreduction Processes over TiO2 Photocatalyst. In Photocatalysts—Applications and Attributes; IntechOpen: London, UK, 2018; p. 17. [Google Scholar]
  36. Pan, X.; Xu, Y.J. Defect-mediated growth of noble-metal (Ag, Pt, and Pd) nanoparticles on TiO2 with oxygen vacancies for photocatalytic redox reactions under visible light. J. Phys. Chem. C 2013, 117, 17996–18005. [Google Scholar] [CrossRef]
  37. AutoChem 2920 Automated Catalyst Characterization System. In Operator’s Manual; Micromeritics Instrument Corporation: Norcross, GA, USA, 2014.
  38. Treacy, J.P.W.; Hussain, H.; Torrelles, X.; Grinter, D.; Cabailh, G.; Bikondoa, O.; Nicklin, C.; Selcuk, S.; Selloni, A.; Lindsay, R.; et al. Geometric structure of anatase TiO2(101). Phys. Rev. B 2017, 95, 1–7. [Google Scholar] [CrossRef] [Green Version]
  39. Deshmane, V.G.; Owen, S.L.; Abrokwah, R.Y.; Kuila, D. Mesoporous nanocrystalline TiO2 supported metal (Cu, Co, Ni, Pd, Zn, and Sn) catalysts: Effect of metal-support interactions on steam reforming of methanol. J. Mol. Catal. A Chem. 2015, 408, 202–213. [Google Scholar] [CrossRef] [Green Version]
  40. Mendez, C.M.; Olivero, H.; Damiani, D.E.; Volpe, M.A. On the role of Pd β-hydride in the reduction of nitrate over Pd based catalyst. Appl. Catal. B Environ. 2008, 84, 156–161. [Google Scholar] [CrossRef]
  41. Baylet, A.; Marécot, P.; Duprez, D.; Castellazzi, P.; Groppi, G.; Forzatti, P. In situ Raman and in situ XRD analysis of PdO reduction and Pd° oxidation supported on γ-Al2O3 catalyst under different atmospheres. Phys. Chem. Chem. Phys. 2011, 13, 4607. [Google Scholar] [CrossRef]
  42. Bratan, V.; Munteanu, C.; Hornoiu, C.; Vasile, A.; Papa, F.; State, R.; Preda, S.; Culita, D.; Ionescu, N.I. CO oxidation over Pd supported catalysts—In situ study of the electric and catalytic properties. Appl. Catal. B Environ. 2017, 207, 166–173. [Google Scholar] [CrossRef]
  43. González, C.A.; Ardila, A.N.; de Correa, C.M.; Martínez, M.A.; Fuentes-Zurita, G. Pd/TiO2 Washcoated Cordierite Minimonoliths for Hydrodechlorination of Light Organochlorinated Compounds. Ind. Eng. Chem. Res. 2007, 46, 7961–7969. [Google Scholar] [CrossRef]
  44. Tauc, A.; Grigorovici, J.; Vancu, R. Optical Properties and Electronic Structure of Amorphous Germanium. Basic Solid State Phys. 1966, 15, 627–637. [Google Scholar] [CrossRef]
  45. Santara, B.; Pal, B.; Giri, P.K. Signature of strong ferromagnetism and optical properties of Co doped TiO2 nanoparticles. J. Appl. Phys. 2011, 110, 114322. [Google Scholar] [CrossRef]
  46. Khairy, W.; Zakaria, M. Effect of metal-doping of TiO2 nanoparticles on their photocatalytic activities toward removal of organic dyes. Egypt. J. Pet. 2014, 23, 419–426. [Google Scholar] [CrossRef] [Green Version]
  47. Ola, O.; Maroto-Valer, M.M. Transition metal oxide based TiO2 nanoparticles for visible light induced CO2 photoreduction. Appl. Catal. A Gen. 2015, 502, 114–121. [Google Scholar] [CrossRef]
  48. Sobana, N.; Muruganadham, M.; Swaminathan, M. Nano-Ag particles doped TiO2 for efficient photodegradation of Direct azo dyes. J. Mol. Catal. A Chem. 2006, 258, 124–132. [Google Scholar] [CrossRef]
  49. Leong, K.H.; Chu, H.Y.; Ibrahim, S.; Saravanan, P. Palladium nanoparticles anchored to anatase TiO2 for enhanced surface plasmon resonance-stimulated, visible-light-driven photocatalytic activity. Beilstein J. Nanotechnol. 2015, 6, 428–437. [Google Scholar] [CrossRef] [Green Version]
  50. Yang, X.; Ma, F.; Li, K.; Guo, Y.; Hu, J.; Li, W.; Huo, M.; Guo, Y. Mixed phase titania nanocomposite codoped with metallic silver and vanadium oxide: New efficient photocatalyst for dye degradation. J. Hazard. Mater. 2010, 175, 429–438. [Google Scholar] [CrossRef]
  51. Kuvarega, A.T.; Krause, R.W.M.; Mamba, B.B. Nitrogen / Palladium-Codoped TiO2 for Efficient Visible Light Photocatalytic Dye Degradation. J. Phys. Chem. C 2011, 115, 22110–22120. [Google Scholar] [CrossRef]
  52. Lee, Y.; Kim, E.; Park, Y.; Kim, J.; Ryu, W.H.; Rho, J.; Kim, K. Photodeposited metal-semiconductor nanocomposites and their applications. J. Mater. 2018, 4, 83–94. [Google Scholar] [CrossRef]
  53. Peng, J.; Wang, S. Performance and characterization of supported metal catalysts for complete oxidation of formaldehyde at low temperatures. Appl. Catal. B Environ. 2007, 73, 282–291. [Google Scholar] [CrossRef]
  54. Gomes, J.F.; Leal, I.; Bednarczyk, K.; Gmurek, M.; Stelmachowski, M.; Diak, M.; Emília Quinta-Ferreira, M.; Costa, R.; Quinta-Ferreira, R.M.; Martins, R.C. Photocatalytic ozonation using doped TiO2 catalysts for the removal of parabens in water. Sci. Total Environ. 2017, 609, 329–340. [Google Scholar] [CrossRef] [PubMed]
  55. Maicu, M.; Hidalgo, M.C.; Colón, G.; Navío, J.A. Comparative study of the photodeposition of Pt, Au and Pd on pre-sulphated TiO2 for the photocatalytic decomposition of phenol. J. Photochem. Photobiol. A Chem. 2011, 217, 275–283. [Google Scholar] [CrossRef]
  56. Bahruji, H.; Bowker, M.; Davies, P.R.; Morgan, D.J.; Morton, C.A.; Egerton, T.A.; Kennedy, J.; Jones, W. Rutile TiO2–Pd Photocatalysts for Hydrogen Gas Production from Methanol Reforming. Top. Catal. 2015, 58, 70–76. [Google Scholar] [CrossRef]
  57. Rusinque, B.; Escobedo, S.; de Lasa, H. Photocatalytic hydrogen production under near-UV using Pd-doped mesoporous TiO2 and ethanol as organic scavenger. Catalysts 2019, 9, 33. [Google Scholar] [CrossRef] [Green Version]
  58. Riyapan, S.; Boonyongmaneerat, Y.; Mekasuwandumrong, O.; Yoshida, H.; Fujita, S.-I.; Arai, M.; Panpranot, J. Improved catalytic performance of Pd/TiO2 in the selective hydrogenation of acetylene by using H2-treated sol-gel TiO2. J. Mol. Catal. A Chem. 2014, 383–384, 182–187. [Google Scholar] [CrossRef]
  59. Akbayrak, S.; Tonbul, Y.; Özkar, S. Nanoceria supported palladium (0) nanoparticles: Superb catalyst in dehydrogenation of formic acid at room temperature. Appl. Catal. B Environ. 2017, 206, 384–392. [Google Scholar] [CrossRef]
  60. Zhang, H.; Sun, J.; Dagle, V.L.; Halevi, B.; Datye, A.K.; Wang, Y. Influence of ZnO facets on Pd/ZnO catalysts for methanol steam reforming. ACS Catal. 2014, 4, 2379–2386. [Google Scholar] [CrossRef]
  61. Onderwaater, W.G.; Taranovskyy, A.; van Baarle, G.C.; Frenken, J.W.M.; Groot, I.M.N. In Situ Optical Reflectance Difference Observations of CO Oxidation over Pd (100). J. Phys. Chem. C 2017, 121, 11407–11415. [Google Scholar] [CrossRef] [Green Version]
  62. Ravishankar, T.N.; Vaz, M.D.O.; Ramakrishnappa, T.; Teixeira, S.R.; Dupont, J.; Pai, R.K.; Banuprakash, G. The heterojunction effect of Pd on TiO2 for visible light photocatalytic hydrogen generation via water splitting reaction and photodecolorization of trypan blue dye. J. Mater. Sci. Mater. Electron. 2018, 29, 11132–11143. [Google Scholar] [CrossRef]
  63. Caudillo-Flores, U.; Muñoz-Batista, M.J.; Fernández-García, M.; Kubacka, A. Bimetallic Pt-Pd co-catalyst Nb-doped TiO2 materials for H2 photo-production under UV and Visible light illumination. Appl. Catal. B Environ. 2018, 238, 533–545. [Google Scholar] [CrossRef]
  64. Sokolov, S.; Ortel, E.; Kraehnert, R. Mesoporous titania films with adjustable pore size coated on stainless steel substrates. Mater. Res. Bull. 2009, 44, 2222–2227. [Google Scholar] [CrossRef]
  65. Das, S.K.; Bhunia, M.K.; Bhaumik, A. Self-assembled TiO2 nanoparticles: Mesoporosity, optical and catalytic properties. Dalt. Trans. 2010, 39, 4382–4390. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, S.-H.; Syu, H.-R. One-step fabrication of N-doped mesoporous TiO2 nanoparticles by self-assembly for photocatalytic water splitting under visible light. Appl. Energy 2012, 100, 148–154. [Google Scholar] [CrossRef]
  67. Yu, J.C.; Wang, X.; Fu, X. Pore-Wall Chemistry and Photocatalytic Activity of Mesoporous Titania Molecular Sieve Films. Chem. Mater. 2004, 16, 1523–1530. [Google Scholar] [CrossRef]
  68. Wang, J.; Li, H.; Li, H.; Zou, C.; Wang, H.; Li, D. Mesoporous TiO2 thin films exhibiting enhanced thermal stability and controllable pore size: Preparation and photocatalyzed destruction of Cationic Dyes. ACS Appl. Mater. Interfaces 2014, 6, 1623–1631. [Google Scholar] [CrossRef]
  69. Salas, S.E.; de Lasa, H. Photocatalytic Water Splitting using a Modified Pt-TiO2. Kinetic Modeling and Hydrogen Production Efficiency. Ph.D. Thesis, The University of Western Ontario, London, ON, Canada, 27 August 2013. [Google Scholar]
  70. De Lasa, H.; Serrano, B.; Salaices, M. Photocatalytic Reaction Engineering; Springer Science: New York, NY, USA, 2005. [Google Scholar]
  71. Fagerlund, G. Determination of specific surface by the BET method. Matér. Constr. 1973, 6, 239–245. [Google Scholar] [CrossRef]
  72. Slav, A. Optical characterization of TiO2-Ge nanocomposite films obtained by reactive magnetron sputtering. Dig. J. Nanomater. Biostruct. 2011, 6, 915–920. [Google Scholar]
  73. Guo, S.P.; Li, J.C.; Xu, Q.T.; Ma, Z.; Xue, H.G. Recent achievements on polyanion-type compounds for sodium-ion batteries: Syntheses, crystal chemistry and electrochemical performance. J. Power Sources 2017, 361, 285–299. [Google Scholar] [CrossRef]
  74. Colmenares, J.C.; Lisowski, P.; Łomot, D.; Chernyayeva, O.; Lisovytskiy, D. Sonophotodeposition of Bimetallic Photocatalysts Pd-Au/TiO2: Application to Selective Oxidation of Methanol to Methyl Formate. ChemSusChem 2015, 8, 1676–1685. [Google Scholar] [CrossRef]
  75. Sanchez, C.; Boissière, C.; Grosso, D.; Laberty, C.; Nicole, L. Design, Synthesis, and Properties of Inorganic and Hybrid Thin Films Having Periodically Organized Nanoporosity. Chem. Mater. 2008, 20, 682–737. [Google Scholar] [CrossRef]
  76. Ortel, E.; Polte, J.; Bernsmeier, D.; Eckhardt, B.; Paul, B.; Bergmann, A.; Strasser, P.; Emmerling, F.; Kraehnert, R. Pd/TiO2 coatings with template-controlled mesopore structure as highly active hydrogenation catalyst. Appl. Catal. A Gen. 2015, 493, 25–32. [Google Scholar] [CrossRef]
  77. Salaices, M.; Serrano, B.; de Lasa, H. Photocatalytic Conversion of Organic Pollutants Extinction Coefficients and Quantum Efficiencies. Ind. Eng. Chem. Res. 2001, 40, 5455–5464. [Google Scholar] [CrossRef]
Figure 1. X-ray Diffractograms for Pd-Doped TiO2 Photocatalysts. XRDs for A = anatase and Pd = palladium are shown as a reference.
Figure 1. X-ray Diffractograms for Pd-Doped TiO2 Photocatalysts. XRDs for A = anatase and Pd = palladium are shown as a reference.
Catalysts 10 00074 g001
Figure 2. Comparative Analysis of 0.25 wt% Pd-TiO2 Photocatalysts Before and After Reduction. A = anatase, PdO = palladium oxide and Pd° = metallic palladium.
Figure 2. Comparative Analysis of 0.25 wt% Pd-TiO2 Photocatalysts Before and After Reduction. A = anatase, PdO = palladium oxide and Pd° = metallic palladium.
Catalysts 10 00074 g002
Figure 3. Temperature Programmed Reduction (TPR) of TiO2 and 0.25 wt% Pd-TiO2.
Figure 3. Temperature Programmed Reduction (TPR) of TiO2 and 0.25 wt% Pd-TiO2.
Catalysts 10 00074 g003
Figure 4. Band Gap Calculation Using the Tauc Plot Methodology. A straight-line extrapolation is shown for the 0.25 wt% Pd–TiO2.
Figure 4. Band Gap Calculation Using the Tauc Plot Methodology. A straight-line extrapolation is shown for the 0.25 wt% Pd–TiO2.
Catalysts 10 00074 g004
Figure 5. High-Resolution X-ray Photoelectron Spectroscopy (XPS) Spectra for 0.25 wt% Pd–TiO2: (a) Before photoreduction and (b) After photoreduction. Note: Continuous lines represent Pd° at (i) 3d5/2 and (ii) 3d3/2. Broken lines represent PdO at (iii) 3d5/2 and (iv) 3d3/2.
Figure 5. High-Resolution X-ray Photoelectron Spectroscopy (XPS) Spectra for 0.25 wt% Pd–TiO2: (a) Before photoreduction and (b) After photoreduction. Note: Continuous lines represent Pd° at (i) 3d5/2 and (ii) 3d3/2. Broken lines represent PdO at (iii) 3d5/2 and (iv) 3d3/2.
Catalysts 10 00074 g005
Figure 6. Schematic Representation of the 6000 cm3 Slurry Control Volume Involved in the Macroscopic Radiation Balances.
Figure 6. Schematic Representation of the 6000 cm3 Slurry Control Volume Involved in the Macroscopic Radiation Balances.
Catalysts 10 00074 g006
Figure 7. Cumulative Hydrogen Volume Using Pd-TiO2 photocatalysts with Different Metal Loadings (0.25, 1.50, 1.00, 2.50 and 5.00 wt%). Conditions: photocatalyst concentration 0.15 g/L, 2.0 v/v% ethanol, pH = 4 ± 0.05 and visible light. Note: The experimental data reported is the result of 4 consecutive runs, each lasting 6 h. Standard deviation for repeats were +/−4.3%.
Figure 7. Cumulative Hydrogen Volume Using Pd-TiO2 photocatalysts with Different Metal Loadings (0.25, 1.50, 1.00, 2.50 and 5.00 wt%). Conditions: photocatalyst concentration 0.15 g/L, 2.0 v/v% ethanol, pH = 4 ± 0.05 and visible light. Note: The experimental data reported is the result of 4 consecutive runs, each lasting 6 h. Standard deviation for repeats were +/−4.3%.
Catalysts 10 00074 g007
Figure 8. Schematic Representation of: (a) The synthesized photocatalysts following calcination at 500 °C with most of the Pd being present as PdO, (b) The photoreduction of the PdO to Pd° using a UV-Lamp for 1 h, (c) The H2 production using a photoreduced Pd-TiO2, with molecular H2 being generated on the semiconductor.
Figure 8. Schematic Representation of: (a) The synthesized photocatalysts following calcination at 500 °C with most of the Pd being present as PdO, (b) The photoreduction of the PdO to Pd° using a UV-Lamp for 1 h, (c) The H2 production using a photoreduced Pd-TiO2, with molecular H2 being generated on the semiconductor.
Catalysts 10 00074 g008
Figure 9. Cumulative Hydrogen Volume at Different 0.25, 0.50, 1.00, 2.50 and 5.00 wt% Pd Loadings utilizing a photoreduced photocatalyst after UV light exposure. Conditions: Photocatalyst concentration: 0.15g/L, 2.0 v/v% ethanol, pH = 4 ± 0.05. Experimental data is the result of 4 consecutive runs for each photocatalyst.
Figure 9. Cumulative Hydrogen Volume at Different 0.25, 0.50, 1.00, 2.50 and 5.00 wt% Pd Loadings utilizing a photoreduced photocatalyst after UV light exposure. Conditions: Photocatalyst concentration: 0.15g/L, 2.0 v/v% ethanol, pH = 4 ± 0.05. Experimental data is the result of 4 consecutive runs for each photocatalyst.
Catalysts 10 00074 g009
Figure 10. Hydrocarbon Profiles of (a) Carbon dioxide (CO2), (b) Methane (CH4), (c) Acetaldehyde (C2H4O) and (d) Ethane (C2H6) at 0.25 wt% Pd. Conditions: Photocatalyst concentration 0.15 g/L, 2.0 v/v% ethanol, argon atmosphere, pH = 4 ± 0.05 after UV light photoreduction.
Figure 10. Hydrocarbon Profiles of (a) Carbon dioxide (CO2), (b) Methane (CH4), (c) Acetaldehyde (C2H4O) and (d) Ethane (C2H6) at 0.25 wt% Pd. Conditions: Photocatalyst concentration 0.15 g/L, 2.0 v/v% ethanol, argon atmosphere, pH = 4 ± 0.05 after UV light photoreduction.
Catalysts 10 00074 g010
Figure 11. QY% at Various Irradiation Times under Visible Light using a 0.15 g/L of Photocatalyst Concentration. Note: Pd-TiO2 photocatalyst with different palladium loadings: 0.25, 0.50, 1.00, 2.50, and 5.00 wt%.
Figure 11. QY% at Various Irradiation Times under Visible Light using a 0.15 g/L of Photocatalyst Concentration. Note: Pd-TiO2 photocatalyst with different palladium loadings: 0.25, 0.50, 1.00, 2.50, and 5.00 wt%.
Catalysts 10 00074 g011
Figure 12. QY% at Various Irradiation Times, under Visible Light, using a 0.15 g/L of Photocatalyst Concentration. Notes: (a) Pd on TiO2 photocatalyst with palladium loadings: 0.25, 0.5, 1.0, 2.5 and 5.0 wt%) and (b) Photocatalysts photoreduced using Near-UV irradiation for 1 h.
Figure 12. QY% at Various Irradiation Times, under Visible Light, using a 0.15 g/L of Photocatalyst Concentration. Notes: (a) Pd on TiO2 photocatalyst with palladium loadings: 0.25, 0.5, 1.0, 2.5 and 5.0 wt%) and (b) Photocatalysts photoreduced using Near-UV irradiation for 1 h.
Catalysts 10 00074 g012
Figure 13. Schematics of the Photo-CREC Water II Reactor. Components: (a) Opaque polyethylene tube, (b) Fused silica windows, (c) Flow distributor, (d) Gas sampling port, (e) Jet driving mixing port, (f) Self-driven mixing impeller, (g) Centrifugal pump, (h) Pyrex tube, (i) Draining gas valve, (j) Purging gas injector, and (k) Slurry sampling port.
Figure 13. Schematics of the Photo-CREC Water II Reactor. Components: (a) Opaque polyethylene tube, (b) Fused silica windows, (c) Flow distributor, (d) Gas sampling port, (e) Jet driving mixing port, (f) Self-driven mixing impeller, (g) Centrifugal pump, (h) Pyrex tube, (i) Draining gas valve, (j) Purging gas injector, and (k) Slurry sampling port.
Catalysts 10 00074 g013
Table 1. Surface Area, Pore Diameter, and Pore Volume of the Prepared Photocatalysts.
Table 1. Surface Area, Pore Diameter, and Pore Volume of the Prepared Photocatalysts.
PhotocatalystSBET
(m2 g−1)
DpBJH (4 VpBJH/SBET)
(nm)
VpBJH
(cm3g−1)
DP-25 597.50.11
TiO2 14017.50.61
0.25 wt% Pd–TiO2 13116.50.53
0.50 wt% Pd–TiO2 12416.80.52
1.00 wt% Pd–TiO212321.20.65
2.50 wt% Pd–TiO212219.90.60
5.00 wt% Pd–TiO211918.90.56
Table 2. Metal Dispersion on TiO2 Photocatalysts.
Table 2. Metal Dispersion on TiO2 Photocatalysts.
PhotocatalystMetal Dispersion (%)
0.25 wt% Pd–TiO275
0.50 wt% Pd–TiO227
1.00 wt% Pd–TiO226
2.50 wt% Pd–TiO212
5.00 wt% Pd–TiO28
Table 3. Photocatalyst Crystallite Sizes.
Table 3. Photocatalyst Crystallite Sizes.
PhotocatalystCrystallite Size (nm)
DP 2521
TiO29
0.25 wt% Pd -TiO2 11
0.50 wt% Pd -TiO2 11
1.00 wt% Pd -TiO2 11
2.50 wt% Pd -TiO2 13
5.00 wt% Pd -TiO2 14
Table 4. Photocatalysts Band Gaps.
Table 4. Photocatalysts Band Gaps.
PhotocatalystBand Gap (eV)
DP-253.10
TiO22.99
0.25 wt% Pd–TiO22.51
0.50 wt% Pd–TiO22.55
1.00 wt% Pd–TiO22.60
2.50 wt% Pd–TiO22.67
5.00 wt% Pd–TiO22.67
Table 5. High-Resolution X-ray Photoelectron Spectroscopy (XPS) Spectra Binding Energies and Peak Areas for the 0.25 wt% Pd–TiO2.
Table 5. High-Resolution X-ray Photoelectron Spectroscopy (XPS) Spectra Binding Energies and Peak Areas for the 0.25 wt% Pd–TiO2.
Peak NameBefore PhotoreductionAfter 60 min of Photoreduction Using Near-UV Irradiation
Binding EnergyFWHM% AreaPosFWHM% Area
Pd 3d3/2 PdO341.542.0050.2341.492.0018.3
Pd 3d3/2 Pd°339.691.1349.8339.561.2981.7
Pd 3d5/2 PdO336.282.0050.2336.232.0018.3
Pd 3d5/2 Pd°334.431.1349.8334.301.2981.7
Table 6. Rates of Absorbed Photons on the Photocatalysts of this study using a Consistent 0.15 g/L Photocatalyst Loading. All photocatalysts doped with Pd were photoreduced for 60 min. prior to their utilization. Pi is estimated to be 9.54 × 10−6 Einsten/s.
Table 6. Rates of Absorbed Photons on the Photocatalysts of this study using a Consistent 0.15 g/L Photocatalyst Loading. All photocatalysts doped with Pd were photoreduced for 60 min. prior to their utilization. Pi is estimated to be 9.54 × 10−6 Einsten/s.
Catalyst LoadingPa (Einstein/s)
TiO22.23 × 10−6
0.25 wt% Pd - TiO24.37 × 10−6
0.50 wt% Pd- TiO24.45 × 10−6
1.00 wt% Pd- TiO25.62 × 10−6
2.50 wt% Pd- TiO24.87 × 10−6
5.00 wt% Pd- TiO24.81 × 10−6
Table 7. QYs% for Pd-TiO2 Photocatalysts at Different Metal Loadings (0.25, 0.50, 1.00, 2.50, and 5.00 wt%) under: (a) Visible light irradiation only, (b) Using near-UV light followed by visible light irradiation.
Table 7. QYs% for Pd-TiO2 Photocatalysts at Different Metal Loadings (0.25, 0.50, 1.00, 2.50, and 5.00 wt%) under: (a) Visible light irradiation only, (b) Using near-UV light followed by visible light irradiation.
PhotocatalystQY (%) (a)QY (%) (b)
TiO20.23-
0.25 wt% Pd TiO21.131.58
0.50 wt% Pd TiO20.341.07
1.00 wt% Pd TiO20.300.80
2.50 wt% Pd TiO20.100.79
5.00 wt% Pd TiO20.100.78

Share and Cite

MDPI and ACS Style

Rusinque, B.; Escobedo, S.; de Lasa, H. Photoreduction of a Pd-Doped Mesoporous TiO2 Photocatalyst for Hydrogen Production under Visible Light. Catalysts 2020, 10, 74. https://doi.org/10.3390/catal10010074

AMA Style

Rusinque B, Escobedo S, de Lasa H. Photoreduction of a Pd-Doped Mesoporous TiO2 Photocatalyst for Hydrogen Production under Visible Light. Catalysts. 2020; 10(1):74. https://doi.org/10.3390/catal10010074

Chicago/Turabian Style

Rusinque, Bianca, Salvador Escobedo, and Hugo de Lasa. 2020. "Photoreduction of a Pd-Doped Mesoporous TiO2 Photocatalyst for Hydrogen Production under Visible Light" Catalysts 10, no. 1: 74. https://doi.org/10.3390/catal10010074

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