Next Article in Journal
Aqueous Dehydration, Hydrogenation, and Hydrodeoxygenation Reactions of Bio-Based Mucic Acid over Ni, NiMo, Pt, Rh, and Ru on Neutral or Acidic Catalyst Supports
Previous Article in Journal
Cyanosilylation of Aldehydes Catalyzed by Ag(I)- and Cu(II)-Arylhydrazone Coordination Polymers in Conventional and in Ionic Liquid Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Hydrogen Evolution Using Bi-Metallic (Ni/Pt) Na2Ti3O7 Whiskers: Effect of the Deposition Order

by
Luis F. Garay-Rodríguez
1,
S. Murcia-López
2,*,
T. Andreu
2,
Edgar Moctezuma
3,
Leticia M. Torres-Martínez
1,* and
J. R. Morante
2
1
Facultad de Ingeniería Civil-Departamento de Ecomateriales y Energía, Universidad Autónoma de Nuevo León, Cd. Universitaria, 66455 San Nicolás de los Garza, N.L., Mexico
2
Catalonia Institute for Energy Research (IREC), Jardins de les Dones de Negre 1, 08930 Sant Adrià de Besós, Spain
3
Facultad de Ciencias Químicas, Universidad Autónoma de San Luis Potosí, Av. Manuel Nava #6, 78290 San Luis Potosí, S.L.P., Mexico
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(3), 285; https://doi.org/10.3390/catal9030285
Submission received: 15 February 2019 / Revised: 7 March 2019 / Accepted: 15 March 2019 / Published: 20 March 2019
(This article belongs to the Section Photocatalysis)

Abstract

:
Photocatalytic hydrogen production through ethanol photo-reforming using Na2Ti3O7 whiskers increases if the sodium titanate is decorated with well-known metallic catalysts such as Ni and Pt. Whereas wet impregnation with nickel gives only a slight increase in the activity, photo-deposition of Pt increased the H2 production by more than one order of magnitude. Through the combination of both co-catalysts (Ni and Pt) a superior performance in terms of H2 production is further observed. However, hydrogen yield is largely enhanced (almost three-fold), up to 778 μmol·g−1·h−1, if the Pt is photo-deposited on the surface of the catalyst before wet impregnation with Ni species (NTO/Pt/Ni) compared to H2 yield (283 μmol·g−1·h−1) achieved with the catalyst prepared in the reverse order (NTO/Ni/Pt). Structural, morphological, optical, and chemical characterization was carried out in order to correlate physicochemical properties with their photocatalytic activity. The X-ray photoelectron spectroscopy (XPS) results show a higher concentration of Pt2+ species if this metallic layer is under the nickel oxide layer. Moreover, X-ray diffraction patterns (XRD) show that Na2Ti3O7 surface is modified for both metal decoration processes.

Graphical Abstract

1. Introduction

In recent years, hydrogen has attracted a lot of attention as one of the most promising substitute of fossil fuels, as a clean energy carrier because its calorific power per mass unit (142 MJ/kg) which is higher than those of such fuels as CH4 (55.5 MJ/kg), carbon (~30 MJ/kg), and gasoline (46.4 MJ/kg) [1,2]. Some strategies have been studied for its production, the heterogeneous photocatalysis being one alternative of great interest considering the possibility of obtaining H2 from water or organic compounds, such as alcohols, by simply using solar radiation. In this sense, photo-reforming of organic molecules, such as alcohols, for hydrogen evolution in both liquid and gas phase, has attracted a lot of attention recently mainly because the organic molecules act as hole scavengers and undergo relatively fast and irreversible oxidation [3].
Among several semiconductor metal oxides, TiO2 has been one of the most reported photocatalysts for degradation of organic molecules, hydrogen production by the water splitting reaction and CO2 reduction under UV and visible light irradiation [4,5,6,7]. To overcome the TiO2-based system limitations, several strategies have been developed, including the use of co-catalysts to accelerate the charge transfer and avoid recombination or the modification of the semiconductor chemical or crystalline structure to improve the light absorption processes.
Regarding the use of co-catalysts, Ni has been one of the most reported metallic oxides to improve photocatalytic activity of TiO2 for hydrogen production [8,9]. Some recent studies have also shown an enhanced photocatalytic activity of semiconductor catalysts partially covered or a bi-metallic catalyst which gives rise to an improved charge transfer rate [10,11].
In addition to TiO2, materials based on the family of inorganic alkali titanates (with the general formula of A2TinO2n+1) have also shown to be promising alternatives for hydrogen generation through photocatalytic water splitting reactions without sacrificial agents. These materials possess a structure consisting of edge and corner TiO6 octahedral in units of (Ti3O7)2− layers held together by different ions forming layered tunnels [12]. More importantly, their higher efficiencies in the photo-induced processes [13,14], lithium batteries [15,16], chemical adsorption [17,18], and biomedics [19,20], have been attributed to this crystalline arrangement. Some of the most reported alkali titanates based on Na, K, and Rb have been evaluated in the photocatalytic hydrogen evolution reaction under UV–Vis irradiation, being the Na titanates the most efficient photocatalysts under different reaction conditions. This is mainly attributed to the lower distortion of its crystalline structure as a result of the use of a cation with a lower atomic radius [21,22].
Therefore, sodium tri-titanate (Na2Ti3O7) has been prepared by different synthesis methods, leading to nanostructures with varied morphologies such as particles, rods, tubes, belts, and ribbons [23,24], which have an important effect on their properties for specific applications. For instance, Na2Ti3O7 belts have shown high activity for the photocatalytic degradation of organic pollutants and hydrogen evolution in liquid phase batch reactors. In particular, the superior photocatalytic activity of sodium titanates doped with hydrogen or impregnated with different co-catalysts, such as noble metals (Ag, Au, Pt) or In2S3, has been attributed to the low charge recombination rate and to the capacity of absorbing more energy under UV–Vis illumination [25,26]. In fact, despite the higher band gap of Na2Ti3O7 (3.6 eV) [26], the conduction band potential is thermodynamically negative enough for a better performance in the hydrogen evolution reaction [27].
In this scenario, this work presents the evaluation of the photocatalytic activity of Na2Ti3O7 whiskers prepared by solid-state reaction in the ethanol photo-reforming process to produce hydrogen in a continuous gas phase reactor under UV radiation. For this purpose, special interest has been paid on the improvement achieved by the use of nickel and platinum co-catalysts. Therefore, titanate surface was decorated with Ni by the wet impregnation method and with Pt by photo-deposition in order to obtain nickel oxide and metallic platinum species [28,29] that will modify the photocatalytic properties of the semiconductor.

2. Results and Discussion

2.1. Characterization

Figure 1a presents the XRD patterns of the bare Na2Ti3O7 (NTO) and catalysts loaded with Ni or Pt (NTO/Ni and NTO/Pt), where sodium titanate was the main observed phase. Due to the low concentration of co-catalysts, there is an absence of Pt or Ni species reflections in the patterns.
The diffractogram of the nickel loaded catalyst shows some extra small spikes in the baseline at around 10.9° and 24.2°; these reflections are more evident on the diffractogram of the platinum-loaded catalyst. These extra reflections correspond to some of the main reflections of the H2Ti3O7 phase (JCPDS 00-036-0654), suggesting the protonation of the synthesized Na2Ti3O7 as a result of impregnation (Ni) or photo-deposition (Pt) conditions. Furthermore, the Pt-loaded sample presents very small reflections around 14.5° and 18.4° which correspond to the NaTi8O13 phase; however, as a result of their imperceptible intensity, they can be omitted. The majority phase in the patterns (Na2Ti3O7) in Figure 1a displays a monoclinic crystalline structure (JCPDS 00-031-1329) with an average crystallite size, calculated by the Scherrer equation using the (0 0 1) reflection of around 75 nm.
More complex features were observed in the bimetallic samples as shown in the XRD patterns in Figure 1b. As it can be noticed, both samples present more extra reflections with higher intensities than the single metal loaded ones, which are representative of a modification of the pristine phases. They may be due to Na deficiency in the crystalline structure of sodium titanate and these reflections may indicate the formation of NaTi8O13 and H2Ti3O7 phases (JCPDS 00-048-0523 and 00-036-0654, respectively), being the first one representative of a family of sodium titanates with Ti3+/Ti4+ mixed-valence [30]. In addition, it is evident that the (0 0 1) reflection moved to slightly higher 2θ values, which also suggests the partial replacement of sodium atoms to form Na2−xHxTi3O7. According to some studies, these phases are intermediate products of the transformation of Na2Ti3O7 to Na2Ti6O13 under extreme conditions [31]. In this context, we can assume that the bi-metallic deposition on the surface of sodium titanate under the performed reaction conditions favors the partial reduction of titanium and the substitution of the sodium atoms by hydrogen atoms.
It is evident that the intensity of the (0 0 1) reflection (≈10.5°) is different comparing bare and bi-metallic samples. The most noticeable decrease in this pattern is presented by NTO/Ni/Pt; however, NTO/Pt/Ni is also less intense compared to the bare NTO (Figure S1). This reduction in the (0 0 1) titanate main intensity is related with a decrease in the interlayer distance between TiO6 octahedrons [32] suggesting a major replacement of Na by smaller cations such as Pt, Ni or H in the NTO/Ni/Pt sample and reducing the interlayer distance; furthermore, this sample presents more intense peaks of the main H2Ti3O7 phase (24°, 28° and 48°) compared to the NTO/Pt/Ni, corroborating this interlayer reduction and suggesting the major presence of this phase instead of the NaTi8O13. A contrary behavior is observed in NTO/Pt/Ni sample, where H2Ti3O7 phase peaks are less intense compared to the NaTi8O13 (17.8° and 27.8°), suggesting the higher presence of the Ti3+/Ti4+ mixed-valence on this catalyst.
A less marked behavior is observed in single metal-loaded samples, as it is observed, also in Figure S1, Ni- and Pt-loaded NTO present a reduction in the (0 0 1) reflection intensity, as a result of the decrease in the interlayer distance and more evident in the NTO/Ni sample, which suggests the possible introduction of Ni species between the TiO6 octahedrons.
In this context, it is important to highlight that the metal deposition order produces a different development in the formed faces from NTO.
Bi-metallic photocatalysts were analyzed by XPS to obtain more information about their chemical composition.
Figure S2 shows the XPS survey of NTO, NTO/Ni/Pt, and NTO/Pt/Ni samples. As it is evident, some peaks come from the covered substrate, Na and Ti, while others come from the deposited metals (Pt and Ni) being much more intense the Pt signal in NTO/Ni/Pt sample compared to the NTO/Pt/Ni one, whereas the Ni signal is less intense. Carbon corresponds to some surface contamination whereas the signal of oxygen comes from the NTO substrate, possible oxygen bond with the surface carbon contamination, as well as from the oxidized state of the Ni and Pt.
Figure 2 shows the de-convoluted spectra of Na 1s, Ti 2p, and O 1s species in the bare Na2Ti3O7 sample; as it can be noticed, Ti 2p spectrum was resolved in three single peaks characteristic of Ti2p3/2 and Ti 2p1/2 orbitals and their respective satellite, all of them located at 458.1, 463.8, and 471.3 eV, respectively, and representative of Ti4+ species [33]. On the contrary, Na 1s and O 1s present both a single peak located at 1071.1 and 529.7 eV, respectively, which are related to the presence of Na+ and O2− [34,35] species in the sample. All of these peaks corroborates the presence of a single Na2Ti3O7 phase.
A different behavior is observed in the NTO/Ni/Pt sample, which is evidenced on its de-convoluted XPS spectra presented in Figure 3. In this context, the Ti 2p region was resolved on its representative Ti 2p3/2 and 2p1/2 peaks and its satellite at 472.8 eV. Both Ti 2p3/2 and 2p1/2 peaks were fitted in two doublet ones centered at 459, 460, 464, and 465.7 eV, respectively. Both peaks located at 459 and 465.7 are representative of the Ti4+ species [33]; meanwhile, the 460 and 464 peaks correspond to the Ti3+ species [36]. The presence of these peaks corroborates the Ti4+/Ti3+ mixture in the intermediate phase formed during bi-metallic synthesis (NaTi8O13). A peak of Na 1s was found at 1071.95 eV (Figure 3) as evidence of Na+ species present in the sample [34]. The O 1s spectrum was resolved in a very intensive peak at 530 eV, corresponding to O2− species [35] related to the formation of O–M bonds, where M can be any of the present metals. A peak at 534 eV was resolved suggesting the presence of water in the surface, as a result of the lack of thermal treatment after Pt photo-deposition process. This step was omitted in order to avoid the Pt oxidation by temperature.
Ni 2p and Pt 4f spectra are also shown in Figure 3. The Ni 2p spectrum was resolved in characteristic Ni 2p3/2 and 2p1/2 peaks, located at around 855.5 and 873 eV; furthermore, the respective satellites of these peaks were found at around 860.5 and 880 eV. These signals are associated to Ni2+ oxidation state, [37] which might exist in the form of Ni(OH)2 over the titanate surface, as expected by the impregnation conditions (calcination at 400 °C after impregnation and further Pt photo-deposition without calcination). On the other hand, the Pt 4f spectrum was de-convoluted into four peaks, with the most intense one located at 76.1 eV, characteristic of Pt 4f7/2 and attributed to Pt4+ species. A second peak was resolved at 79.8 eV, which is also associated with Pt4+ (Pt 4f5/2) [38]. Other peaks with lower intensities were resolved at 74.1 and 77.2 eV associated with Pt 4f7/2 and Pt 4f5/2 (Pt2+). The very high intensity of Pt4+/Pt2+ peaks is indicative of partially oxidized states as a consequence of an incomplete photo-deposition process, which has also been reported for some photo-deposition methods under certain conditions [39]. The XPS quantitative analyses of this sample for each element peak are collected in Table S1.
The XPS survey spectrum of NTO/Pt/Ni sample is presented also in Figure S2. As in the NTO/Ni/Pt, representative peaks of the expected elements were observed. The de-convoluted spectra obtained from the NTO/Pt/Ni sample are shown in Figure 4. In this case, as Ni was impregnated in the outer layer, the intensities of its peaks are relatively higher than those of Pt. In addition, Na 1s, Ti 2p, and O 1s peaks were found also at similar binding energies than the previous sample, thus confirming that these elements are in the same oxidation states. In addition, Ti 2p spectra also presented the characteristic peak of Ti3+ species conforming again the presence of a mixture of Ti4+/Ti3+ in the intermediate phases formed. Compared to the NTO/Ni/Pt sample, the O 1s spectrum mainly corresponds to the O–M bonds. The Ni 2p spectrum was de-convoluted with the Ni 2p3/2 and 2p1/2 signals located at around 855.39 and 873.02 eV with their respective satellites at around 861.11 and 879.36 eV, which are associated with Ni2+ oxidation state. With this information, we can conclude that the co-catalyst deposition order does not affect the oxidation state of Ni because the catalysts are calcined under air atmosphere and the metallic particles impregnated over titanate are apparently oxidized to Ni2+ species (NiO or Ni(OH)2) [37]. On the other hand, the Pt 4f spectra, unlike the previous sample, now present an oxidation state Pt2+ as a predominant platinum contribution. The peaks found at 73.6 and 76.5 eV are characteristic of Pt 4f5/2 and Pt 4f7/2 attributed to Pt2+ species. Furthermore, the other peaks at 75.5 and 79.5 eV are representative of Pt 4f7/2 and Pt 4f5/2 (Pt4+) species. This mayor difference in oxidation states could suggest the presence of several phases such as PtO2, or PtO species under the photo-deposition. Likewise, the posterior nickel impregnation seems to prevent full oxidation of the platinum.
Table 1 shows the calculated ratios between Ti4+/Na, Ti3+/Ti4+, and Pt2+/Pt4+ from both samples. The slightly higher value of Ti3+/Ti4+ ratio in NTO/Pt/Ni sample suggests the higher number of oxygen vacancies in this sample compared to NTO/Ni/Pt (0.46 vs. 0.36) as a result of the formation of the intermediate phase NaTi8O13; this information confirms the results obtained by XRD, with more intense peaks associated with this phase on this sample. The lower value of the Ti4+/Na ratio in the NTO/Ni/Pt sample gives information about the higher Na loss on this sample because of the possible Ni substitution between the TiO6 octahedrons or its fast transition to the H2Ti3O7 phase. This also agrees with the XRD patterns, as the NTO/Ni/Pt sample presents more intense diffraction peaks corresponding to the H2Ti3O7 protonated phase compared to the other one (NTO/Ni/Pt). Finally, the higher Pt2+/Pt4+ ratio on the NTO/Pt/Ni sample (1.45 vs. 0.32) suggests a higher Pt reduction efficiency than over bare NTO and over Ni-loaded NTO, which can be related to the phase transition to the mixed valence phase and the formation of oxygen vacancies.
The C 1s spectra of both materials also present differences in the surface contamination. Due to the Pt photo-deposition after Ni in the NTO/Ni/Pt sample (Figure 3), an extra peak at 286.3 eV is observed; this peak could suggest the presence of carbonates on the catalyst surface (as confirmed also by the 531.7 eV peak in O 1s spectrum) associated to the use of isopropanol during the photo-deposition reaction which was not completely removed by washing. However, both peaks disappear completely in the NTO/Pt/Ni sample as a result of thermal treatment at 400 °C, carried out after Ni impregnation.
SEM images of Na2Ti3O7 samples are shown in Figure 5a,b. As it can be seen, this material exhibits a whisker-like morphology, which is characteristic of some alkali metal titanates [40]. In general, the samples have an apparent average whisker length of around 5 μm, resulting from the high temperatures used during solid-state reaction synthesis.
Micrographs of the samples loaded with Ni/Pt and Pt/Ni are shown in Figure 5c,d, respectively. It is evident that metal impregnation or photo-deposition did not modify the morphology of the support. EDS analyses confirmed the presence of Ni and Pt in both samples. The quantitative EDS analyses are presented in Table 2. As it can be noticed, the Na loss during both co-catalyst depositions as a result of the transition to the protonated phase is evident. More notorious changes are presented when bi-metallic samples are evaluated; in concordance to XRD and XPS analysis, more Na deficiency was observed in the NTO/Pt/Ni sample compared to NTO/Ni/Pt, suggesting the transition to the mixed valence phase and the protonated one. Pt and Ni as first co-catalyst (samples NTO/Ni and NTO/Pt) seem not to show a significant change even with the second metal deposition process; for that reason, their concentration was similar in NTO/Ni/Pt and NTO/Pt/Ni.
N2 adsorption was used for determining the BET surface area of the samples. Due to the high-temperature thermal treatment applied during the synthesis method, the bare titanate exhibits a relatively low surface area (<10 m2·g−1), which practically remains unmodified after single Ni or Pt decoration. However, in the case of bimetallic samples, surface areas had a significate variation, which can be attributed to the secondary process. The NTO/Ni/Pt sample exhibited a BET area increase, although the value remains lower than 10 m2·g−1. On the contrary, the NTO/Pt/Ni sample presented a six-fold increase (28 m2·g−1) compared with sample only loaded with Pt (4 m2·g−1). In addition, pore size distributions of both samples are presented in Figure 6. As it is evident, higher pore volume was obtained in sample NTO/Pt/Ni (0.029 cm3·g−1) compared with NTO/Ni/Pt (0.009 cm3·g−1). This difference in pore volume between samples can be attributed to the sodium loss during the Pt photo-deposition and Ni impregnation processes forming some intermediate phases with higher pore volume than bare titanate. The results from BET analysis of all materials synthesized are summarized in Table 3.
In order to perform the optical characterization, diffuse reflectance UV–VIS analysis of the samples was carried out. Figure 7 shows the diffuse reflectance spectra of NTO metallic and bi-metallic samples. A sharp decrease in reflectance is observed around 365 nm, which makes evident the higher absorption of these materials under UV radiation. As it is evident, a decrease in reflectance is observed with the Ni impregnation. In addition, the NTO/Pt/Ni sample presents a transition around of 550 nm, which is representative of Ni(OH)2 species [41] confirming the previous assumption made with XPS results. On the contrary, this transition is less visible on the NTO/Ni/Pt sample, probably due to the Pt species in the outer layer. No Ni(OH)2 transitions were observed in the NTO/Ni sample as a result of the thermal treatment performed after impregnation suggesting the presence of only NiO particles. Furthermore, some other transitions less marked are observed, mainly in the Pt loaded samples between 500–800 nm; some of them can be considered representative of PtO2 or Pt(OH)2 species [42,43], confirming the presence of these phases as a result of the incomplete Pt reduction over catalysts.
With this data and by using the Kubelka–Munk transformation, Tauc plots were obtained, as shown in Figure S3. As seen, a minor variation in the bandgap values (Table 3) is observed after single metal or bi-metallic deposition. In this context, the bare NTO presents a bandgap value of 3.4 eV, which is similar to the previously reported in the literature [26]. The lowest value was calculated for the NTO/Ni sample (3.25 eV); this reduction in Eg can be related to the presence of Ni2+ species between TiO6 octahedrons as a result of its exchange during the impregnation conditions, creating impurity states within the band gap and resulting in a reduction of itself [25,44].
In general, no significant variation is found in the modified samples as a result of the very low concentration of loaded metals (1 wt %).
As a result of the further characterization of the prepared catalysts, it is possible to propose a reaction mechanism of the phase transition during both processes (impregnation and photo-deposition). Figure 8 shows the proposed mechanism of the deposition of a single co-catalyst. As it can be observed, during the Ni impregnation, as a result of the performed conditions, it is possible to exchange some Na+ species by Ni2+ ones into the TiO6 layers [45], reducing the interlayer distance between them. Furthermore, NiO species were deposited over the NTO surface.
A different behavior was observed as a result of the Pt photo-deposition conditions. In this context, thanks to the catalyst illumination, electrons (e) and holes (h+) are produced in the NTO surface. These species are responsible for the isopropanol degradation (h+) and Pt precursor reduction (e) [39,46]; however, electrons are also capable of the Ti4+ species reduction producing Ti3+ ones [47]. This reduction is responsible for the oxygen vacancies generation and the formation of the NaTi8O13 phase (Ti3+/Ti4+ mixed valence). From the isopropanol degradation, protons (H+) are generated and this species can be exchanged by Na into the layered tunnels resulting in the formation of the protonated phase (Na2−xHxTi3O7).
A similar development is presented in the case of bi-metallic samples. In this context, it is important to remind that the starting material for the second deposition is the obtained from the single metal loaded catalyst explained previously. The proposed mechanism for the synthesis of the bi-metallic catalyst is presented in Figure 9.
As it can be observed, for the Pt photo-deposition in the second layer (NTO/Ni/Pt sample), a Na+ exchanged by Ni2+ sample is photo-activated. In this context, as a result of this exchange, a favored transition to the H2Ti3O7 phase is observed, probably associated to a higher generation of protons. Unfortunately, electrons are not enough for completely reducing the Pt precursors and more PtO2 species were detected according to the XPS data; in addition, this e deficiency was not enough for the oxygen vacancies generation producing a lower transition to the Ti3+/Ti4+ mixed phase (NaTi8O13) formation.
In the case of the NTO/Pt/Ni sample, a Pt-loaded protonated phase mixed with a lower quantity of NaTi8O13 is used as starting material. As it was explained before, during the Ni impregnation it is possible the Na+ exchanges with Ni2+ into the TiO6 layers; in this context, as a result of the Na loss during this process it is able to grow the Na-deficient phase (NaTi8O13) thanks to the posterior thermal treatment.

2.2. Photocatalytic Evaluation

Figure 10 summarizes the hydrogen evolution reaction rate of the experiments carried out under UV irradiation (λmax = 365 nm) in a continuous gas phase reactor. It can be seen that the bare sample (NTO) has a low hydrogen production reaction rate under these conditions (13 μmol·g−1·h−1). Hydrogen was not formed when the reaction experiment was carried out with bare glass substrates.
The use of a single co-catalyst (Ni or Pt) on this NTO material leads to an increase in the hydrogen production, moderate in the case of nickel and higher in the case of the Pt, as shown in the same figure. This behavior can be mainly due to the presence of Pt on the NTO surface, which improves the carrier transfer step. Comparison on the effect of Pt and Ni deposition shows more than one order of magnitude improvement (factor 18) for Pt, while Ni deposition hardly favors the hydrogen evolution in relation to the bare NTO, with a two-fold increase [26,48].
The effect of the order in the metal deposition on the photo-reforming reaction efficiency is also clearly observed. As shown in Figure 10, the NTO/Ni/Pt sample (Ni impregnation before Pt photo-deposition) exhibits lower hydrogen reaction rate (283 vs. 778 μmol·g−1·h−1) compared to NTO/Pt/Ni sample (Pt deposition before Ni). This fact can be explained through the combination of several factors. First at all, as we have Pt deposited onto the NTO surface, which allowed the formation of the more reactive H2Ti3O7 or Na2−xHxTi3O7 phases [49], as demonstrated by the XRD, XPS, and EDS analysis. Secondly, the NTO/Pt/Ni catalyst has a higher surface area and pore volume than the other samples, implying a higher number of active sites for adsorption of ethanol-water molecules. Third, according to XPS, calculated Ti3+/Ti4+ ratios of the NTO/Ni/Pt catalyst present a higher quantity of oxygen vacancies which also correspond to reactive sites for the reaction [50]. Fourth, NTO/Pt/Ni apparently presents more Pt2+ species than Pt4+ ones, the Pt being the reduced species that is more efficient for photocatalytic hydrogen production [51,52,53].
The time-dependent rate of hydrogen evolution over the most efficient semiconductor catalyst (NTO/Pt/Ni) is presented in Figure S4. As seen, hydrogen production increases continuously during the first 5 h of reaction and remains stable over time (16 h), which is indicative of initial pre-activation and further catalyst stability. In addition, products of the oxidation reactions of ethanol were not detected during the experiments. These species may be adsorbed on the surface of the catalysts or they may be present in very low concentrations and could not be detected during the GC analysis.
Furthermore, in order to corroborate the catalyst stability, XPS and SEM characterization was performed on the recovered sample after the photocatalytic reaction experiments. The XPS results are shown in Figure S5. As seen, the formation of this Ni 2p spectra did not present significant variations in the position of peaks, which evidences the apparent stability in the oxidation state of Ni species impregnated over titanate surface. In case of Pt spectrum, no peaks were practically detected due to the high intensity of C 1s species present in the surface, masking the low Pt 4f signals as well as those of the other buried layers. This increase in intensity in the C 1s spectrum can be related to the adsorption in the surface of some products not evolved as a result of the oxidative process.
Additional EDS mapping of the post-reaction NTO/Pt/Ni sample in Figure S6 shows the mixed and individual maps of Na, Ti, Ni, and O, clearly indicating that the whiskers retain the Na, Ti, and O composition, with a very homogeneous distribution of Ni.

3. Materials and Methods

3.1. Catalysts Synthesis

Na2Ti3O7 synthesis (NTO) was performed by the solid-state reaction; for this purpose, stoichiometric proportions of anhydrous Na2CO3 (99% DEQ Químicos, Monterrey, México) and TiO2 (Evonik P25, Essen, Germany) were perfectly mixed in an agate mortar using acetone as dispersant media. The wet slurry was transferred into a platinum crucible and thermally treated in air at 800 °C for 12 h.
Nickel oxide was deposited on the surface of the sodium titanate (NTO/Ni) by the wet impregnation method. The appropriate amount of nickel acetate (Fermont, Monterrey, México), calculated for 1% in weight, was dissolved in 40 mL of anhydrous ethanol. Then, Na2Ti3O7 was added to the solution and mixed under vigorous agitation for 2 h. Then, the solvent was evaporated by heating at 110 °C. The dry solid was thermally treated at 400 °C for 2 h in air in order to transform the nickel organic salt deposited on the surface of the support to nickel oxide species.
Platinum was deposited by using a photo-deposition method (NTO/Pt) [54]. For this purpose, chloroplatinic acid hydrate (≥99.9%, Sigma Aldrich, Darmstadt, Germany), calculated for 1% in weight of Pt, was dissolved in a water-isopropanol (0.3 M) mixture. Na2Ti3O7 was added to the solution and the suspension was sonicated for 15 min. Photo-deposition was carried out by using an Osram lamp (300 W, I = 100 mW·cm−2) under a continuous N2 flow (0.5 mL·min−1) for 2 h. Under this reaction conditions, only metallic platinum is expected to be deposited on the surface of the support. Finally, the suspensions were centrifuged and washed with Milli-Q water in order to eliminate remaining chlorine species. The wet solids were dried at 110 °C.
Bi-metallic catalysts were prepared by a similar procedure. The first batch of catalysts was labeled NTO/Ni/Pt indicating that nickel oxide species form the first layer of co-catalyst and the metallic platinum form the second layer. The second batch of bi-metallic catalyst was labeled NTO/Pt/Ni indicating that the metals were deposited in the reverse order.

3.2. Characterization

The structural characterization was performed with a Bruker D8 Advance X-Ray (Bruker Corporation, Billerica, MA, USA) diffractometer (CuKα radiation) equipped with an LYNXEYE super speed detector and a Ni filter over the 2θ collection range of 10–70° with a scan rate of 0.05° s−1.
The morphology of the photocatalysts was analyzed with a Field Emission Scanning Electron Microscope (FESEM, Zeiss Auriga, Madrid, Spain) equipped with an Electron Dispersive X-Ray analyzer. XPS analyses were carried out in PHI 5500 Multitechnique equipment (Physical Electronics, Chanhassen, MN, USA) with Al Kα radiation.
Diffuse reflectance UV–VIS spectra of all the photocatalysts were obtained in a UV–VIS NIR spectrophotometer (Cary 5000, Agilent Technologies, Santa Clara, CA, USA) coupled with an integrating sphere. The band gap of the samples was calculated with the Kubelka–Munk function.
BET surface area measurements were performed by N2 adsorption–desorption isotherms using a Micromeritics TriStar II instrument (Micrometrics Instrument Corp, Norcross, GA, USA).

3.3. Photocatalytic Reactions

In order to evaluate the photocatalytic activity of the Na2Ti3O7 and Ni-Pt-Na2Ti3O7 catalysts for hydrogen evolution through ethanol reforming, several gas-phase experiments were carried out in a continuous reactor.
Catalysts were deposited on glass substrates previously cleaned sonicated with a mixture of acetone, isopropanol, and water during 15 min. Then, 0.1 g of powder catalysts were suspended in 6 mL of Mili-Q water and sonicated 15 min to form a homogenous slurry that was deposited by drop-casting over the glass substrates and naturally dried at room temperature for about 12 h. The remaining moisture was eliminated by heating at 80 °C during 15 min under air flow.
The glass substrates are placed inside a circular reaction cell illuminated with a Hamamatsu Lightningcure Spot LC8 UV-lamp (I = 21.8 mW·cm−2 at 365 nm, Hamamatsu Photonics K. K., Bridgewater, NJ, USA) located at 15 cm above the quartz window. An argon stream (14 mL·min−1) is bubbled through a water/ethanol mixture (50:50 v/v), kept at constant temperature of 35 °C, and continuously fed to the photocatalytic reactor.
The outlet of the reactor system is connected to a Varian 490-Micro GC (Agilent Technologies, Santa Clara, CA, USA) equipped with MS5A and PPQ columns. A sample of the gas stream was injected to the GC every 15 min to determine H2 concentration.
A schematic representation of the reaction system is presented in Figure 11.

4. Conclusions

Na2Ti3O7 whiskers were successfully synthesized by the solid state reaction. Ni and Pt were deposited over the titanate surface with different methodologies (wet impregnation for Ni and photo-deposition for Pt) resulting in the formation of NiO, Ni(OH)2, and Pt (in different oxidation states, Pt2+ or Pt4+). Both metals gave raise to NTO modification observed in the XRD spectra.
Whereas Ni impregnation did not lead to a relevant increase in catalytic activity, Pt did, achieving an increment higher than one order of magnitude. The combination of both catalysts had a synergic effect, and its order of deposition was a very important factor. The previous nickel impregnation (NTO/Ni/Pt) did not lead to a significant increase in photocatalytic activity; on the contrary, the previous photo-deposition of platinum (NTO/Pt/Ni) led to almost three-fold higher productivity in the NTO/Ni/Pt sample (i.e., 778 versus 283 μmol·g−1·h−1, respectively). Aside from this difference in the photocatalytic performance, there are also other complementary differences between these NTO/Pt/Ni and NTO/Ni/Pt samples. First, in both cases, the samples present modifications in their XRD spectra, likely attributed to the formation of the reduced phase of the NTO. The experimental data evidence that the longer bi-metallic deposition times give rise to a higher partial reduction of bare material creating phases with a mixture of Ti4+/Ti3+ valences and leading to the formation of protonated H2Ti3O7 phases that origin these extra peaks in the XRD spectra. Second, the sample ending with nickel shows a higher surface area and larger porous volume pointing out a larger effective surface to interact with the alcohol molecules. Third, the oxidation state of the platinum is mainly Pt2+ in the NTO/Pt/Ni sample whereas it is predominantly Pt4+ in the NTO/Ni/Pt one. In all the case, nickel corresponds to Ni2+ oxidation state related to the presence of NiO and/or Ni(OH)2. All these features facilitate an increase in photocatalytic activity for the hydrogen evolution proving the synergic effect in the combination of these two catalyst pointing out the more significant role played by the platinum in direct contact with the NTO substrate and the complementary role played by the nickel and its higher effective area to facilitate the final charge transfer to the water-ethanol molecules.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/3/285/s1, Figure S1: Close view of (0 0 1) main reflection on the prepared catalysts, Figure S2: XPS survey spectra of NTO, NTO/Ni/Pt and NTO/Pt/Ni samples, Figure S3: Kubelka–Munk spectra of Na2Ti3O7 bare and deposited with Ni and Pt metals, Figure S4: Time dependent hydrogen evolution over NTO/Pt/Ni sample, Figure S5: XPS characterization of the sample with the best performance (NTO/Pt/Ni) post reaction, Figure S6: EDX mapping analysis of the NTO/Pt/Ni sample after photocatalytic test, Table S1: Semi-quantitative analysis of XPS peaks of bi-metallic samples.

Author Contributions

Methodology: L.F.G.-R. and S.M.-L.; supervision: T.A., L.M.T.-M. and J.R.M.; validation: S.M.-L.; writing–original draft: L.F.G.-R. and S.M.-L.; writing–review and editing: T.A., E.M., L.M.T.-M. and J.R.M.

Funding

This research was founded by Generalitat de Catalunya through the CERCA Program and M2E (2017SGR1246); CONACYT through projects CB-2014-237049, PDCPN-2015-487, Ph. D. scholarship 635249, and “Becas Mixtas 2017 Movilidad en el extranjero 291212”; SEP through PROFIDESS-PRODEP-25292. IREC also acknowledges support by the European Regional Development Funds (ERDF, FEDER) and by MINECO project ENE2017-85087-C3-2-R. S.M.-L. thanks European Union’s Horizon 2020 and the Agency for Business Competitiveness of the Government of Catalonia for funding under the Marie Sklodowska-Curie grant agreement no. 712939 (TECNIOspring PLUS).

Acknowledgments

Authors thank to David Avellaneda from FIME-UANL for his valuable help with the XPS measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ahmed, A.; Al-Amin, A.Q.; Ambrose, A.F.; Saidur, R. Hydrogen fuel and transport system: A sustainable and environmental future. Int. J. Hydrogen Energy 2016, 41, 1369–1380. [Google Scholar] [CrossRef]
  2. Singh, S.; Jain, S.; Ps, V.; Tiwari, A.K.; Nouni, M.R.; Pandey, J.K.; Goel, S. Hydrogen: A sustainable fuel for future of the transport sector. Renew. Sustain. Energy Rev. 2015, 51, 623–633. [Google Scholar] [CrossRef]
  3. Kawai, T.; Sakata, T. Photocatalytic hydrogen production from liquid methanol and water. J. Chem. Soc. Chem. Commun. 1980, 15, 694–695. [Google Scholar] [CrossRef]
  4. Lin, W.-C.; Yang, W.-D.; Huang, I.-L.; Wu, T.-S.; Chung, Z.-J. Hydrogen production from methanol/water photocatalytic decomposition using Pt/TiO2−xNx catalyst. Energy Fuels 2009, 23, 2192–2196. [Google Scholar] [CrossRef]
  5. Strataki, N.; Bekiari, V.; Kondarides, D.I.; Lianos, P. Hydrogen production by photocatalytic alcohol reforming employing highly efficient nanocrystalline titania films. Appl. Catal. B Environ. 2007, 77, 184–189. [Google Scholar] [CrossRef]
  6. Petala, A.; Ioannidou, E.; Georgaka, A.; Bourikas, K.; Kondarides, D.I. Hysteresis phenomena and rate fluctuations under conditions of glycerol photo-reforming reaction over CuOx/TiO2 catalysts. Appl. Catal. B Environ. 2015, 178, 201–209. [Google Scholar] [CrossRef]
  7. Gombac, V.; Sordelli, L.; Montini, T.; Delgado, J.J.; Adamski, A.; Adami, G.; Cargnello, M.; Bernal, S.; Fornasiero, P. CuOx-TiO2 photocatalysts for H2 production from ethanol and glycerol solutions. J. Phys. Chem. A 2010, 114, 3916–3925. [Google Scholar] [CrossRef] [PubMed]
  8. Chen, W.-T.; Chan, A.; Sun-Waterhouse, D.; Moriga, T.; Idriss, H.; Waterhouse, G.I.N. Ni-TiO2: A promising low-cost photocatalytic system for solar H2 production from ethanol-water mixtures. J. Catal. 2015, 326, 43–53. [Google Scholar] [CrossRef]
  9. Liu, R.; Yoshida, H.; Fujita, S.-I.; Arai, M. Photocatalytic hydrogen production from glycerol and water with NiOx/TiO2 catalysts. Appl. Catal. B Environ. 2014, 144, 41–45. [Google Scholar] [CrossRef]
  10. Luna, A.L.; Novoseltceva, E.; Louarn, E.; Beaunier, P.; Kowalsa, E.; Ohtani, B.; Valenzuela, M.A.; Remita, H.; Colbeau-Justin, C. Synergetic effect of Ni and Au nanoparticles synthesized on titania particles for efficient hydrogen production. Appl. Catal. B Environ. 2016, 191, 18–28. [Google Scholar] [CrossRef]
  11. Luna, A.L.; Dragoe, D.; Wang, K.; Beaunier, P.; Kowalska, E.; Ohtani, B.; Uribe, D.B.; Valenzuela, M.A.; Remita, H.; Colbeau-Justin, C. Photocatalytic hydrogen evolution using Ni-Pd/TiO2: Correlation of light absorption, charge-carrier dynamics, and quantum efficiency. J. Phys. Chem. C 2017, 121, 14302–14311. [Google Scholar] [CrossRef]
  12. Ramírez-Salgado, J.; Djurado, E.; Fabry, P. Synthesis of sodium titanate composites by sol-gel method for use in gas potentiometric sensors. J. Eur. Ceram. Soc. 2004, 24, 2477–2483. [Google Scholar] [CrossRef]
  13. Ogura, S.; Kohno, M.; Sato, K.; Inoue, Y. Effects of RuO2 on activity for water decomposition of a RuO2/Na2Ti3O7 photocatalysts with a zigzag layer structure. J. Mater. Chem. 1998, 8, 2335–2337. [Google Scholar] [CrossRef]
  14. Ogura, S.; Kohno, M.; Sato, K.; Inoue, Y. Photocatalytic properties of M2Ti6O13 (M = Na, K, Rb, Cs) with rectangular tunnel and layer structures: Behavior of a surface radical produced by UV irradiation and photocatalytic activity for water decomposition. Phys. Chem. Chem. Phys. 1999, 1, 179–183. [Google Scholar] [CrossRef]
  15. Cech, O.; Castkova, K.; Chladil, L.; Dohnal, P.; Cudek, P.; Libich, J.; Vanysek, P. Synthesis and characterization of Na2Ti6O13 and Na2Ti6O13/Na2Ti3O7 sodium titanates with nanorod-like structure as negative electrode materials for sodium-ion batteries. J. Energy Storage 2017, 14, 391–398. [Google Scholar] [CrossRef]
  16. Pérez-Flores, J.C.; García-Alvarado, F.; Hoelzel, M.; Sobrados, I.; Sanz, J.; Kuhn, A. Insight into the channel ion distribution and influence on the lithium insertion properties of hexatitanates A2Ti6O13 (A = Na, Li, H) as candidates for anode materials in lithium-ion batteries. Dalt. Trans. 2012, 41, 14633–14642. [Google Scholar] [CrossRef] [PubMed]
  17. Feng, M.; You, W.; Wu, Z.; Chen, Q.; Zhan, H. Mildly alkaline preparation and methylene blue adsorption capacity of hierarchical flower-like sodium titanate. ACS Appl. Mater. Interfaces 2013, 5, 12654–12662. [Google Scholar] [CrossRef]
  18. Wang, Y.; Liu, H. Preparation and characterizations of Na2Ti3O7, H2Ti3O7 and TiO2 nanobelts. Adv. Mater. Res. 2011, 306–307, 1233–1237. [Google Scholar] [CrossRef]
  19. Ma, X.W.; Shen, J.X.; Zhang, K.C.; Kong, L.K.; le Sun, J.; Zhang, J.Y. Self-toughness porous sodium titanate bioceramics prepared via in situ grown Na2Ti6O13 rods. Key Eng. Mater 2017, 727, 806–814. [Google Scholar] [CrossRef]
  20. Becker, I.; Hofmann, I.; Müller, F.A. Preparation of bioactive sodium titanate ceramics. J. Eur. Ceram. Soc. 2007, 27, 4547–4553. [Google Scholar] [CrossRef]
  21. Ogura, S.; Sato, K.; Inoue, Y. Effects of RuO2 dispersion on photocatalytic activity for water decomposition of BaTi4O9 with a pentagonal prism tunnel and K2Ti4O9 with zigzag layer structure. Phys. Chem. Chem. Phys. 2000, 2, 2449–2454. [Google Scholar] [CrossRef]
  22. Inoue, Y.; Kubokawa, T.; Sato, K. Photocatalytic activity of alkali-metal titanates combined with Ru in the decomposition of water. J. Phys. Chem. 1991, 95, 4059–4063. [Google Scholar] [CrossRef]
  23. Mancic, L.T.; Marinkovic, B.A.; Jardim, P.M.; Milosevic, O.B.; Rizzo, F. Precursor particle size as the key parameter for isothermal tuning of morphology from nanofibers to nanotubes in the Na2−xHxTinO2n+1 system through hydrothermal alkali treatment of rutile mineral sand. Cryst. Growth Des. 2009, 9, 2152–2158. [Google Scholar] [CrossRef]
  24. Zhang, X.Y.; Zhang, Y.J.; Gong, H.Y.; Song, Y.N. Preparation and characterization of sodium hexatitanate whiskers. Adv. Mater. Res. 2013, 602–604, 1339–1343. [Google Scholar] [CrossRef]
  25. Vithal, M.; Krishna, S.R.; Ravi, G.; Palla, S.; Velchuri, R.; Pola, S. Synthesis of Cu2+ and Ag+ doped Na2Ti3O7 by a facile ion-exchange method as visible-light driven photocatalysts. Ceram. Int. 2013, 39, 8429–8439. [Google Scholar] [CrossRef]
  26. Liu, J.; Ding, T.; Li, Z.; Zhao, J.; Li, S.; Liu, J. Photocatalytic hydrogen production over In2S3-Pt-Na2Ti3O7 nanotube films under visible light irradiation. Ceram. Int. 2013, 39, 8059–8063. [Google Scholar] [CrossRef]
  27. Araújo-Filho, A.A.; Silva, F.L.R.; Righi, A.; da Silva, M.B.; Silva, B.P.; Caetano, E.W.S.; Freire, V.N. Structural, electronic and optical properties of monoclinic Na2Ti3O7 from density functional theory calculations: A comparison with XDR and optical absorption measurements. J. Solid State Chem. 2017, 250, 68–74. [Google Scholar] [CrossRef]
  28. Murcia, J.J.; Navío, J.A.; Hidalgo, M.C. Insights towards the influence of Pt features on the photocatalytic activity improvement of TiO2 platinization. Appl. Catal. B Environ. 2012, 126, 76–85. [Google Scholar] [CrossRef]
  29. Prasad, M.S.; Chen, R.; Ni, H.; Kumar, K.K. Directly grown of 3D-nickel oxide nano flowers on TiO2 nanowire arrays by hydrothermal route for electrochemical determination of naringenin flavonoid in vegetable samples. Arab. J. Chem. 2018. [Google Scholar] [CrossRef]
  30. Akimoto, J. Synthesis and crystal structure of NaTi8O13. J. Solid State Chem. 1991, 90, 147–154. [Google Scholar] [CrossRef]
  31. Xu, H.; Li, C.; He, D.; Jinag, Y. Stability and structure changes of Na-titanate nanotubes at high temperature and high pressure. Powder Diffr. 2014, 29, 147–150. [Google Scholar] [CrossRef]
  32. Lai, S.-W.; Park, J.-W.; Yoo, S.-H.; Ha, J.-M.; Song, E.-H.; Cho, S.-O. Surface synergism of Pd/H2Ti3O7 composite nanowires for catalytic and photocatalytic hydrogen production from ammonia borane. Int. J. Hydrogen Energy 2016, 41, 3428–3435. [Google Scholar] [CrossRef]
  33. Xing, J.-H.; Xia, Z.-B.; Hu, J.-F.; Zhang, Y.-H.; Zhong, L. Growth and crystallization of titanium oxide films at different anodization modes. J. Elctrochem. Soc. 2013, 160, C239–C246. [Google Scholar] [CrossRef]
  34. Murai, S.; Fujita, K.; Kawase, S.; Ukon, S.; Tanaka, K. Formation of silver nanoparticles under anodic surface of tellurite glass via thermal poling-assisted ion implantation across solid-solid interface. J. Appl. Phys. 2007, 102, 073515. [Google Scholar] [CrossRef]
  35. Chia-Ching, W.; Cheng-Fu, Y. Investigation of the properties of nanostructured Li-doped NiO films using the modified spray pyrolysis method. Nanoscale Res. Lett. 2013, 8, 33. [Google Scholar] [CrossRef]
  36. Bharti, B.; Kumar, S.; Lee, H.-N.; Kumar, R. Formation of oxygen vacancies and Ti3+ state in TiO2 thin film and enhanced optical properties by air plasma treatment. Sci. Rep. 2016, 6, 32355–32367. [Google Scholar] [CrossRef]
  37. George, G.; Anandhan, S. Synthesis and characterization of nickel oxide nanofiber webs with alcohol sensing characteristics. RSC Adv. 2014, 4, 62009–62020. [Google Scholar] [CrossRef]
  38. Matin, M.A.; Lee, E.; Kim, H.; Yoon, W.-S.; Kwon, Y.-U. Rational syntheses of core-shell Fe@(PtRu) nanoparticle electrocatalysts for the methanol oxidation reaction with complete suppression of CO-poisoning and highly enhanced activity. J. Mater. Chem. A 2015, 3, 17154–17164. [Google Scholar] [CrossRef]
  39. Wenderich, K.; Mul, G. Methods, mechanism, and applications of the photodeposition in photocatalysis: A review. Chem. Rev. 2016, 116, 14587–14619. [Google Scholar] [CrossRef]
  40. Wang, W.; Yu, C.; Liu, Y.; Hou, J.; Zhu, H.; Jiao, S. Single crystalline Na2Ti3O7 rods as an anode material for sodium-ion batteries. RSC Adv. 2013, 3, 1041–1044. [Google Scholar] [CrossRef]
  41. Hall, D.S.; Lockwood, D.J.; Bock, C.; MacDougall, B.R. Nickel hydroxides and related materials: A review of their structures, synthesis and properties. Proc. Math. Phys. Eng. Sci. 2015, 471, 20140792. [Google Scholar] [CrossRef] [PubMed]
  42. Vorontsov, A.V.; Savinov, E.N.; Zhensheng, J. Influence of the form of photodeposited platinum on titania upon its photocatalytic activity in CO and acetone oxidation. J. Photochem. Photobiol. A Chem. 1999, 125, 113–117. [Google Scholar] [CrossRef]
  43. Li, F.B.; Li, X.Z. The enhancement of photodegradation efficiency using Pt-TiO2 catalyst. Chemosphere 2002, 48, 1103–1111. [Google Scholar] [CrossRef]
  44. Machida, M.; Ma, X.W.; Taniguchi, H.; Yabunaka, J.-I.; Kijima, T. Pillaring and photocatalytic properties of partially substituted layered titanates, Na2Ti3−xMxO7 and K2Ti4−xMxO9 (M = Mn, Fe, Co, Ni, Cu). J. Mol. Catal. A Chem. 2000, 155, 131–142. [Google Scholar] [CrossRef]
  45. Li, N.; Zhang, L.; Chen, Y.; Fang, M.; Zhang, J.; Wang, H. Highly efficient, irreversible and selective ion exchange property of layered titanate nanostructures. Adv. Funct. Mater. 2012, 22, 835–841. [Google Scholar] [CrossRef]
  46. Wenderich, K.; Han, K.; Mul, G. The effect of methanol on the photodeposition of Pt nanoparticles on tungsten oxide. Part. Part. Syst. Charact. 2018, 35, 1700250. [Google Scholar] [CrossRef]
  47. Xiong, L.-B.; Li, J.-L.; Yang, B.; Yu, Y. Ti3+ in the surface of titanium dioxide: Generation, properties and photocatalytic application. J. Nanomater. 2012, 2012, 831524. [Google Scholar] [CrossRef]
  48. Townsend, T.K.; Browning, D.; Osterloh, F.E. Overall photocatalytic water splitting with NiOx-SrTiO3—A revised mechanism. Energy Environ. Sci. 2012, 5, 9543–9550. [Google Scholar] [CrossRef]
  49. Du, Q.; Lu, G. Controllable synthesis and photocatalytic properties study of Na2Ti3O7 and H2Ti3O7 nanotubes with high exposed facet (0 1 0). J. Nanosci. Nanotechnol. 2015, 15, 4385–4391. [Google Scholar] [CrossRef]
  50. Liu, J.; Ma, X.; Yang, L.; Liu, X.; Han, A.; Lv, H.; Zhang, C.; Xu, S. In situ green oxidation synthesis of Ti3+ and N self-doped SrTiOxNy nanoparticles with enhanced photocatalytic activity under visible light. RSC Adv. 2018, 8, 7142–7158. [Google Scholar] [CrossRef]
  51. Parayil, S.K.; Kibombo, H.S.; Wu, C.-M.; Peng, R.; Kindle, T.; Mishra, S.; Ahrenkiel, S.P.; Baltrusaitis, J.; Dimitrijevic, N.M.; Rajh, T.; et al. Synthesis-dependent oxidation state of platinum on TiO2 and their influences on the solar simulated photocatalytic hydrogen production from water. J. Phys. Chem. C 2013, 33, 16850–16862. [Google Scholar] [CrossRef]
  52. Wang, Y.; Wang, Y.; Xu, R. Photochemical deposition of Pt on CdS for H2 evolution from water: Markedly enhanced activity by controlling Pt reduction environment. J. Phys. Chem. C 2013, 117, 783–790. [Google Scholar] [CrossRef]
  53. Majrik, K.; Pászti, Z.; Korecz, L.; Trif, L.; Domján, A.; Bonura, G.; Cannilla, C.; Frusteri, F.; Tompos, A.; Tálas, E. Study of PtOx/TiO2 photocatalysts in the photocatalytic reforming of glycerol: The role of co-catalyst formation. Materials 2018, 11, 1927. [Google Scholar] [CrossRef]
  54. Murcia-López, S.; Vaiano, V.; Hidalgo, M.C.; Navío, J.A.; Sannino, D. Photocatalytic reduction of CO2 over platinized Bi2WO6-based materials. Photochem. Photobiol. Sci. 2015, 14, 678–685. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Na2Ti3O7 pure and loaded with Ni and Pt (a), and NTO/Ni/Pt and NTO/Pt/Ni (b).
Figure 1. XRD patterns of Na2Ti3O7 pure and loaded with Ni and Pt (a), and NTO/Ni/Pt and NTO/Pt/Ni (b).
Catalysts 09 00285 g001
Figure 2. XPS spectra of Na 1s, Ti 2p, and O 1s in the NTO sample.
Figure 2. XPS spectra of Na 1s, Ti 2p, and O 1s in the NTO sample.
Catalysts 09 00285 g002
Figure 3. XPS spectra in C 1s, Ti 2p, Na 1s, O 1s, Ni 2p, and Pt 4f zones of the NTO/Ni/Pt sample.
Figure 3. XPS spectra in C 1s, Ti 2p, Na 1s, O 1s, Ni 2p, and Pt 4f zones of the NTO/Ni/Pt sample.
Catalysts 09 00285 g003
Figure 4. XPS spectra in C 1s, Ti 2p, Na 1s, O 1s, Ni 2p, and Pt 4f zones of the NTO/Pt/Ni sample.
Figure 4. XPS spectra in C 1s, Ti 2p, Na 1s, O 1s, Ni 2p, and Pt 4f zones of the NTO/Pt/Ni sample.
Catalysts 09 00285 g004
Figure 5. SEM images of NTO (a,b), NTO/Ni/Pt (c), and NTO/Pt/Ni (d), and EDX analysis of NTO/Ni/Pt (e) and NTO/Pt/Ni (f).
Figure 5. SEM images of NTO (a,b), NTO/Ni/Pt (c), and NTO/Pt/Ni (d), and EDX analysis of NTO/Ni/Pt (e) and NTO/Pt/Ni (f).
Catalysts 09 00285 g005
Figure 6. Pore size distribution of (a) NTO/Ni/Pt and (b) NTO/Pt/Ni samples.
Figure 6. Pore size distribution of (a) NTO/Ni/Pt and (b) NTO/Pt/Ni samples.
Catalysts 09 00285 g006
Figure 7. Diffuse reflectance spectra of Na2Ti3O7 bare and deposited with Ni and Pt metals.
Figure 7. Diffuse reflectance spectra of Na2Ti3O7 bare and deposited with Ni and Pt metals.
Catalysts 09 00285 g007
Figure 8. Proposed reaction mechanism of the phase transition during the Na2Ti3O7 single Ni impregnation and Pt photo-deposition.
Figure 8. Proposed reaction mechanism of the phase transition during the Na2Ti3O7 single Ni impregnation and Pt photo-deposition.
Catalysts 09 00285 g008
Figure 9. Proposed reaction mechanism of the phase transition during the bi-metallic Na2Ti3O7 synthesis.
Figure 9. Proposed reaction mechanism of the phase transition during the bi-metallic Na2Ti3O7 synthesis.
Catalysts 09 00285 g009
Figure 10. H2 evolution rates from photocatalytic photo-reforming reaction under metallic and bi-metallic NTO samples.
Figure 10. H2 evolution rates from photocatalytic photo-reforming reaction under metallic and bi-metallic NTO samples.
Catalysts 09 00285 g010
Figure 11. Schematic representation of the photocatalytic gas phase system used.
Figure 11. Schematic representation of the photocatalytic gas phase system used.
Catalysts 09 00285 g011
Table 1. XPS ratios obtained from NTO/Ni/Pt and NTO/Pt/Ni samples.
Table 1. XPS ratios obtained from NTO/Ni/Pt and NTO/Pt/Ni samples.
SampleTi4+/NaTi3+/Ti4+Pt2+/Pt4+
NTO/Ni/Pt0.780.360.32
NTO/Pt/Ni0.280.461.45
Table 2. Elemental quantification of the synthesized catalysts.
Table 2. Elemental quantification of the synthesized catalysts.
MaterialAtomic (%) QuantificationWeight (%) Quantification
NaPtNiNaPtNi
NTO8.5----9.2----
NTO/Ni8.4--0.78.9--1.2
NTO/Pt8.10.1--8.40.8--
NT/Ni/Pt7.40.20.57.91.01.1
NT/Pt/Ni3.40.10.94.80.71.0
Table 3. Textural and optical characterization of the synthesized catalysts.
Table 3. Textural and optical characterization of the synthesized catalysts.
Surface Area (m2/g)Pore Volume (cm3/g)Bandgap (eV)
20.0023.40
30.0033.25
40.0043.43
70.0093.45
280.0293.36

Share and Cite

MDPI and ACS Style

Garay-Rodríguez, L.F.; Murcia-López, S.; Andreu, T.; Moctezuma, E.; Torres-Martínez, L.M.; Morante, J.R. Photocatalytic Hydrogen Evolution Using Bi-Metallic (Ni/Pt) Na2Ti3O7 Whiskers: Effect of the Deposition Order. Catalysts 2019, 9, 285. https://doi.org/10.3390/catal9030285

AMA Style

Garay-Rodríguez LF, Murcia-López S, Andreu T, Moctezuma E, Torres-Martínez LM, Morante JR. Photocatalytic Hydrogen Evolution Using Bi-Metallic (Ni/Pt) Na2Ti3O7 Whiskers: Effect of the Deposition Order. Catalysts. 2019; 9(3):285. https://doi.org/10.3390/catal9030285

Chicago/Turabian Style

Garay-Rodríguez, Luis F., S. Murcia-López, T. Andreu, Edgar Moctezuma, Leticia M. Torres-Martínez, and J. R. Morante. 2019. "Photocatalytic Hydrogen Evolution Using Bi-Metallic (Ni/Pt) Na2Ti3O7 Whiskers: Effect of the Deposition Order" Catalysts 9, no. 3: 285. https://doi.org/10.3390/catal9030285

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