Next Article in Journal
Study of the Influence of the Copper Component’s Shape on the Properties of the Friction Material Used in Brakes—Part One, Tribological Properties
Previous Article in Journal
Analysis and Modeling of Aged SAC-Bi Solder Joints Subjected to Varying Stress Cycling Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CeO2-rGO Composites for Photocatalytic H2 Evolution by Glycerol Photoreforming

by
Stefano Andrea Balsamo
1,*,
Eleonora La Greca
2,
Marta Calà Pizzapilo
1,
Salvatore Sciré
1 and
Roberto Fiorenza
1
1
Department of Chemical Sciences, University of Catania, Viale A. Doria 6, 95125 Catania, Italy
2
Institute for The Study of Nanostructured Materials (ISMN)-CNR, Via Ugo La Malfa 153, 90146 Palermo, Italy
*
Author to whom correspondence should be addressed.
Materials 2023, 16(2), 747; https://doi.org/10.3390/ma16020747
Submission received: 22 December 2022 / Revised: 6 January 2023 / Accepted: 10 January 2023 / Published: 12 January 2023
(This article belongs to the Section Catalytic Materials)

Abstract

:
The interaction between CeO2-GO or CeO2-rGO and gold as co-catalysts were here investigated for solar H2 production by photoreforming of glycerol. The materials were prepared by a solar photoreduction/deposition method, where in addition to the activation of CeO2 the excited electrons were able to reduce the gold precursor to metallic gold and the GO into rGO. The presence of gold was fundamental to boost the H2 production, whereas the GO or the rGO extended the visible-light activity of cerium oxide (as confirmed by UV-DRS). Furthermore, the strong interaction between CeO2 and Au (verified by XPS and TEM) led to good stability of the CeO2-rGO-Au sample with the evolved H2 that increased during five consecutive runs of glycerol photoreforming. This catalytic behaviour was ascribed to the progressive reduction of GO into rGO, as shown by Raman measurements of the photocatalytic runs. The good charge carrier separation obtained with the CeO2-rGO-Au system allowed the simultaneous production of H2 and reduction of GO in the course of the photoreforming reaction. These peculiar features exhibited by these unconventional photocatalysts are promising to propose new solar-light-driven photocatalysts for green hydrogen production.

Graphical Abstract

1. Introduction

The development of the hydrogen economy is currently a hot topic highlighted by the recent politic and energy scenario and promoted by increasing interest in the circular economy. In this view, H2 is the ideal energy vector because it can be produced with several sustainable technologies and obtaining, as the main by-products of its combustion, water and CO2 that can be easily integrated in circular approaches (as the “blue hydrogen” where the obtained CO2 is captured and stored [1]). In the vision of an even more decarbonized economy, the production of hydrogen in fully sustainable ways (i.e., “green hydrogen”) is mandatory, together with the use of uncritical materials [2,3]. The solar photoreforming of sustainable organic substrates can be considered a well-performing and green process that allows the production of hydrogen from biomass or waste, valorising them as raw materials and using the solar energy as a renewable energy source [2]. To increase the hydrogen evolution from photoreforming, the suitable selection of photocatalysts that are able to absorb the solar radiation and efficiently oxidize the organic substrates acting as hole scavengers is necessary.
Therefore, the employed photocatalysts (usually semiconductors) should own the conduction band more negative of the H+/H2 potential and the valence band more positive of the H2O/O2 couple and of the oxidation potentials of organic compounds [4]. For the photoreforming reaction, TiO2-based photocatalysts are the most investigated materials [5]. However, just two years ago the EU added titanium to the critical raw materials list [6]. In this contest for the overall sustainability of the process, the investigation of new photocatalysts is necessary.
Among these, in recent years cerium oxide (CeO2, ceria) has emerged as an interesting alternative to TiO2 or ZnO due to its redox properties, suitable band position for the photocatalytic H2 evolution and the lower bandgap (2.7–2.9 eV) with respect to ZnO and TiO2 [7,8,9,10]. To enhance the H2 production with the CeO2-based materials, chemical and/or structural modifications are necessary. The addition of carbonaceous materials, such as graphene oxide (GO) or reduced graphene oxide (rGO), can increase the separation between the photoproduced e/h+ charge carriers and the absorption of the visible part of the solar radiation [11,12,13,14].
To the best of our knowledge, the study of the interaction between the CeO2 and the GO and/or rGO structures is poorly investigated in the literature for the solar photoreforming reaction. For this reason, we examined CeO2-GO and CeO2-rGO composites prepared with easy and green procedures also employing an in situ photoreduction step that allowed the progressive GO reduction into rGO with the contextual evolution of hydrogen. These interesting strategies allowed us to obtain versatile and unconventional photocatalysts that can be proposed as new and green materials for the sustainable H2 production. The main objectives of the work are (i) the investigation of the performance in the photocatalytic H2 production of unconventional CeO2-Au-GO and CeO2-Au-rGo composites; (ii) the study of the modification of the chemico-physical features of CeO2 due to the addition of gold and/or GO and rGO and (iii) the application of an in situ one-step solar photoreduction method for CeO2-based materials.
As the renewable organic substrate for the photoreforming reaction, we chose glycerol due to the growing interest in its valorisation, being obtained in large amounts as secondary product of biodiesel [15,16]. In this contest, the development of sustainable processes, such as H2 production by glycerol photoreforming, can have a double positive effect: (a) the increase in the viability of the biodiesel market with the contextual production of hydrogen and (b) the development of green technologies to exploit the biomass, the glycerol being a biomass derivate. Furthermore, compared to the other organic substrates produced from biomass and used as sacrificial agents for the photoreforming reaction, glycerol leads to easier hydrogen formation, with seven moles of H2 evolved for each glycerol molecule (reaction (1)) [17,18]:
C3H8O3 + 3H2O → 7H2 + 3CO2

2. Materials and Methods

2.1. Sample Preparation

Bare CeO2 was prepared starting from Ce(NO3)3·6H2O (Fluka, Buchs, Switzerland) solubilized in a 0.02 M CTAB solution. The pH was kept above 8, slowly dropping to KOH 1 M. The obtained suspension was stirred at 80 °C for 3 h and then left to digest at room temperature for 1 day. Then, the slurry was centrifuged at 8000 rpm for 15 min several times, washed with deionized water and then dried overnight in oven at 80 °C. Finally, the powders were calcined in air at 350 °C for 3 h. The same method was followed for the synthesis of another CeO2 sample synthetized without any templating agent, coded as CeO2-noCTAB.
CeO2-rGO and CeO2-rGO-Au samples were obtained following a slight modified method proposed in our previous works [12]. Briefly, 250 mg of the as-prepared ceria with CTAB was added to 37.5 mL of absolute ethanol, together with the stoichiometric volume of HAuCl4 in aqueous solution in order to obtain the 1 wt% amount of gold and adding previously sonicated (1 h at room temperature) aliquots of GO in aqueous solution. This slurry was prepared in a batch reactor and purged with nitrogen for 30 min in order to remove all of the oxygen. Then, the slurry was irradiated for 2 h with a solar simulator (Newport, equipped with a 150 W xenon lamp, photon flux ca. equal to 1.9 × 10−7 Einstein s−1 measured with ferrioxalate actinometer). Finally, the powders were dried overnight under vacuum at 60 °C. This temperature was chosen in order to not alter the reduction degree of the GO. The sample CeO2-GO was prepared in the same way described before but without any irradiation and without the addition of the gold precursor. The sample CeO2-Au was prepared without adding the GO in the reaction slurry, whereas the CeO2-GO-Au catalysts were obtained starting from the as-prepared CeO2-Au and adding without irradiation the GO. In addition, these powders were dried overnight under vacuum at 60 °C.

2.2. Photocatalytic Tests

The photocatalytic hydrogen production was conducted under solar irradiation, with a suspension of 1 mg mL−1 of catalyst in 10% (v/v) solution of glycerol in MilliQ water. In detail, a jacketed two-side neck batch reactor with an inner free volume of 68 mL was filled with 5 mg of catalyst and 5 mL of the glycerol solution. The reaction mixture was then purged for 30 min with N2. The solar irradiation was performed with a Newport solar simulator equipped with a 150 W xenon lamp (photon flux ca. equal to 1.9 × 10−7 Einstein s−1) thanks to an optical fibre enclosed in a cylindrical tube inside the reactor through the central neck. After 3 h of irradiation, 1 mL of the reaction gases was collected with a gastight syringe and analysed by gas chromatography with a thermal conductivity detector (GC-TCD) using an Agilent 7890A equipped with a Supelco Carboxen 1010 column.

2.3. Characterization Techniques

UV-vis DRS measurements were carried out with a Perkin Elmer Lambda 650 S UV/VIS spectrophotometer equipped with a 60 mm integrating sphere and using BaSO4 as reference material. The optical band gap was estimated by the TAUC PLOT method applying the Kubelka–Munk function.
FTIR was carried out with a Perkin-Elmer Spectrum Two FT-IR Spectrometer.
Raman spectra were acquired by a JASCO NRS-5500 Raman instrument equipped with a green laser (532 nm, power 1.7 mW) at 10s exposure and acquiring with a CCD detector (temperature 203 K) an L1800 grating with slit setting at 100 × 1000 µm (resolution 4 cm−1).
XPS data were recorded on pellets (4 mm diameter, 0.5 mm thick) after outgassing the samples to a pressure below 2 × 10−8 Torr at 150 °C. A Leibold–Heraeus LHS10 spectrometer (SPECS, Berlin, Germany) was adopted with an Al Kα X-ray source (hν = 1486.6 eV) at 120 W and 30 mA, using C (1s) or Au (4f7/2) as the binding energy reference (284.8 eV and 84.0 eV, respectively) and analysing the spectra with CasaXPS software.
Textural properties were characterized by N2 adsorption–desorption measurements carried out by a Micromeritics Tristar II Plus 3020 (Micromeritics Instrument Corp. Norcross, GA, USA) applying the Brunauer–Emmet–Teller (BET) and the Barret, Joyner and Halenda (BJH) methods in order to estimate the surface area and the pore size distribution, respectively.
Finally, microstructural analysis was accomplished by TEM using a Thermo Scientific Talos F200i S/TEM microscope operating at 200 kV. Both conventional TEM and STEM mode using high angle annular dark field (HAADF) imaging were performed. Moreover, elemental mapping was carried out using energy-dispersive X-ray spectroscopy (EDS) to study the elemental distribution throughout different areas of the sample. EDS maps were acquired using the VELOX software with a dwell time of 100 µs and 256 × 256 frame size.

3. Results

3.1. Photocatalytic Hydrogen Evolution

The CeO2 prepared with the CTAB was used in all the photocatalytic tests. Indeed, it was not possible to determine the H2 produced with the CeO2-noCTAB sample due to the very low amount (lower than the sensitivity of GC), whereas the sample prepared with the surfactant showed a production of 5 μmol gcat−1 h−1. Then, preliminary tests were carried out in order to evaluate the best rGO content added to the cerium dioxide as a photocatalyst.
Figure 1 reports the H2 evolution results over CeO2-based samples with 0.5 wt%, 1 wt% and 2 wt% of rGO.
The samples with 0.5 wt% and 1 wt% of rGO show similar activity, with production rates of around 225 μmol gcat−1 h−1 and 210 μmol gcat−1 h−1, respectively. Slightly lower performances were obtained with the 2 wt% sample. Considering the small difference within the error bars between the 0.5 wt% and the 1 wt% samples, the latter was chosen as the reference amount for this work in order to facilitate the catalysts characterization.
In Figure 2, the photocatalytic activity data for the other examined samples are reported. Bare CeO2 shows very low activity, with a production rate of 5 μmol gcat−1 h−1, slightly enhanced by the addition of GO or rGO (up to 20 μmol gcat−1 h−1 for both the modified samples). As reported by Christoforidis and Fornasiero [4], the activity of a semiconductor without any co-catalyst is very low, especially for materials like CeO2 that have a charge recombination rate faster than TiO2. In this context, the addition of GO on CeO2 allowed us to obtain only a slight increase in the photocatalytic activity. However, as expected, the addition of a metallic co-catalyst increases the activity up to 90 μmol gcat−1 h−1 for the CeO2-Au sample and to 210 μmol gcat−1 h−1 in the presence of GO or rGO.
It is evident how the electron-sink effect of the carbonaceous material is helped by the beneficial effects of gold as a co-catalyst. This behaviour suggests an important role of the gold in the electron transport mechanism. This is clear because the hydrogen production is not affected by the status of the graphene oxide (GO or rGO), with the activities of CeO2-GO and CeO2-rGO being similar, as well as those of CeO2-GO-Au and CeO2-rGO-Au. This can be ascribed to the weak photoreduction degree of GO caused by the poor photocatalytic features of CeO2 under solar irradiation.
However, the photocatalytic properties of ceria can be better exploited after progressive runs. Figure 3 shows the results of five subsequent runs of the CeO2-rGO-Au catalyst. Interestingly, an increment of 60 μmol gcat−1 h−1 was observed after five runs, attributed to the progressive rGO reduction promoted by the increase in the time of solar irradiation (3 h for every run).

3.2. Photocatalyst Characterizations

Figure 4 shows the UV-vis DRS spectra of the examined samples and their relative TAUC plots for the estimation of the optical band gaps. The absorption spectrum of the bare ceria prepared without CTAB as the templating agent is reported in panel (a). It is possible to note, according to the literature, that the presence of CTAB induces a slight bathochromic shift in the absorption band (i.e., decrease in the optical band gap) [19], and this is another positive feature of the use of the templating agent in the CeO2 synthesis. In addition, the presence of graphene oxide does not give evident variation in the band onset. In the gold-containing samples, it is possible to distinguish the signal at 556 nm due to the plasmon of the metal. According to the surface plasmon resonance (SPR) effect theory, due to the lower dielectric constant of CeO2 compared to TiO2, the position of this signal is shifted to higher wavelengths with respect to the TiO2-based samples [20,21,22]. The figure in panel (b) illustrates the TAUC plots for the previously discussed samples. The smallest band gaps are obtained with the CeO2-rGO-Au and CeO2-GO-Au samples (2.56 eV). Analogously, no substantial differences were detected between the CeO2-GO and the CeO2-rGO samples (2.72 eV and 2.71 eV, respectively).
It must be noted that in order to verify the removal of the CTAB templating agent after the synthesis, a calcination temperature of 350 °C was chosen for all the samples [7]. Figure S1 reports the FTIR spectra for the bare cerium dioxide before and after the calcination. It is evident how the signals related to the CTAB (the asymmetric and symmetric C-CH2 stretching vibrations of the methylene chains at 2919 cm−1 and 2851 cm−1, and the weak signal at 3011 cm−1 assigned to the N-CH3 stretching of the methyl group and their wagging vibrations in the intense and sharp peak at 1390 cm−1) disappear after the calcining treatment, confirming the elimination of the surfactant [19].
Figure 5 shows the Raman spectra of the investigated samples. The band at around 460 cm−1 is the F2g signal related to the fluorite-phase skeletal vibrations. No significant variations in its position were detected, thus the ceria structure was not modified due to the addition of GO and/or gold. In all the spectra the weak signal at 600 cm−1 due to Frenkel defects on the bulk is evident. For this reason, a modest quantity of Ce3+ sites are expected [23].
To investigate the occurrence of the reduction process from GO to rGO, we compared the bands of the rGO-containing samples with the bare GO spectrum (showed in Figure S2). All the samples show the typical bands of this material. The most important are surely those located at 1344 cm−1 and 1600 cm−1 (D and G band, respectively). These two signals are important because thanks to them it is possible to estimate the reduction degree of the carbonaceous material: the higher their ratio is (ID/IG), the higher the amount of disordered carbons [12,24]. The parameters related to these bands are reported in Table 1.
The ID/IG ratio increases from 0.88 to 0.99 with the addition of the ceria. As reported in the literature, this is due to the interaction between the terminal oxygens of ceria and the graphene oxide layers [25]. Nevertheless, this value does not change after a photoreduction treatment in the presence or absence of gold. After five runs, it reaches a value of 1.02. However, a slight reduction trend can be found looking at the FWHM values that slightly go down in the reduced samples. Indeed, the reduction degree can also be evaluated as a separation between the two bands, often approximated with the FWHM values: the lower they are, the higher the reduction degree [26]. A slight decrease can be noted after the reduction treatment (136-76 vs. 188-90 in the CeO2-rGO and CeO2-GO, respectively), in addition to an increase in the presence of gold (159-84 for CeO2-rGO-Au) and once again a decrease after a long irradiation (136-70 for the CeO2-rGO-Au after 5 runs). Considering these behaviours, it can be concluded that the photocatalytic features of cerium dioxide under solar irradiation are not enough to obtain an efficient GO reduction, which starts to be progressively reduced after long irradiation steps and in situ during the photocatalytic hydrogen evolution. Finally, at higher Raman shift the second-order bands due to the carbon are present at 2660 cm−1 (2D band), at 2930 cm−1 (D + D’ band) and at 3192 cm−1 (2D’ band). The 2D band is normally forbidden and becomes visible with defective structures. The D + D’ band is ascribed to the combination of phonons with different momenta, so the presence of defects is necessary to allow the transition according to the selection rules [27]. Instead, the 2D’ band is the only one permitted by the selection rules [26].
The textural properties of the samples were investigated by N2 adsorption–desorption isotherms, evaluating the BET surface area and the BJH pore size distribution (Figure S3). All the isotherms are of type IV, confirming the mesoporous nature of the CeO2-based samples. Each curve shows an H2-type hysteresis typical of ink bottle pores, as widely reported for metal oxides [28]. The BET surface area and the mean pores size are summarized in Table 2.
There is a general decrease in surface area compared to the bare CeO2 (about 10 m2 g−1) after the photoreduction of gold and GO. This can be reasonably ascribed to the occurrence of a partial agglomeration of the CeO2 particles due to the further thermal treatment (drying at 60 °C) and the 2 h of solar irradiation [29,30]. Moreover, the intercalation of the CeO2-Au nanoparticles with the graphene oxide sheets induced pores modification [12]. Indeed, the mean pore size was 8 nm in the CeO2-Au sample, whereas it was 17 and 18 nm for the CeO2-GO-Au and CeO2-rGO-Au samples, respectively. This can be due to the presence of interstitial spaces between the nanoparticles and the graphene oxide sheets [31].
To analyse the morphology of the prepared samples, we performed TEM microstructural measures of the CeO2-rGO-Au as the representative sample (Figure 6).
In Figure 6a it is possible to note the morphology of the nanoparticles modified with rGO. In this image, the elemental analyses were carried out to examine the elements’ configurations. The most remarkable image (Figure 6c) shows that the gold is homogeneously dispersed on the catalyst, whereas in Figure 6e it is possible to note that the CeO2-Au nanoparticles are preferentially deposited on the rGO sheets, distinguishable from the delineated edges.
Finally, the Ce 3d and O 1s XPS analyses for the CeO2-rGO-Au sample are reported in Figure 7, while the XPS analyses of the other samples are reported in Figure S4 and Table S1.
Panel (a) reports the Ce analysis of the CeO2-rGO-Au sample. In this region, all the characteristic peaks of the cerium are present, including the Ce3+-related signals at 885 and 903 eV [29,32]. The Ce3+/Ce4+ ratio is a parameter that gives us an idea of the reduction processes and/or defectivity of the material. This value passes from 0.42 of bare CeO2 (spectrum showed in the Figure S4a) to 0.33 of the CeO2-rGO-Au sample shown in panel (a), indicating a slight overall oxidation of the material. This behaviour can be fairly understood if we assume that the photoreduction process rearranges the chemical structure, eliminating the defective Ce3+ sites. Indeed, it is reasonably assumed that a reduction of the GO corresponds to an overall oxidation of the Ce3+ species into Ce4+. Panel (b) shows the O 1s spectrum of the same sample. Here the signals are assigned to five different components: CeIV-O and CeIV-OH (529.5 eV and 531 eV), adsorbed carbonates on the surface (532 eV) and CeIII-O and CeIII-OH (533 eV and 534 eV) [33]. The same signals are found on the bare CeO2 sample (Figure S4b). Regarding the other features, in the C 1s region (Figure S4c), the three principal signals are related to the C-C, the C-O and the C=O of the reduced graphene oxide [32]. In the Au 4f region (Figure S4d), there are the two signals attributed to the Au 4f7/2 and 4f5/2 at 84.0 eV and 87.5 eV, respectively, typical of metallic gold. This pointed to that during the photoreduction process an efficient photoreduction of the gold precursors occurred, explaining the increase in the photocatalytic activity verified for all the gold-based samples.

4. Discussion

From the data reported in the previous sections it is clear that all the components of the proposed photocatalytic composites (CeO2, GO or rGO and Au) contributed to both the photoreduction/deposition preparation method adopted and the H2 evolution by glycerol photoreforming. Indeed, the ceria was the photoactive semiconductor that, due to the solar irradiation, allowed the formation of the e/h+ pairs, able to start the involved reactions. The addition of GO permitted, instead, to extend the photo-response to a larger part of the visible portion of the solar incident light, as confirmed by the optical band gap decrease from 2.75 eV of bare ceria (λexcitation ≤ 451 nm) to 2.56 eV of the CeO2-GO-Au sample (λexcitation ≤ 484 nm). A further exploitation of the visible light was endorsed by the presence of gold due to its surface plasmon resonance effect (activated at 556 nm, as detected by the UV-DRS), which induced the gold electrons to participate in the photoreforming reactions, thus increasing the photocatalytic activity. Notwithstanding that the photocatalytic features of CeO2 were poor compared to TiO2 or ZnO [34,35], the strong interaction between gold and ceria [36,37], also favoured by the presence of the GO sheets, granted a good photodeposition/reduction of Au (confirmed by TEM and XPS).
It is important to note that the proposed photoreduction preparation required the occurrence of similar reactions of the photoreforming process. The first reaction (2) is the formation of the charge carriers, and subsequently the holes (h+) in the valence band (VB) of ceria react with the sacrificial agent (ethanol in the case of the photoreduction method, glycerol in the photoreforming reaction) to promote the charge carrier separation (3). In this way the excited electrons in the conduction band (CB) of ceria are able to reduce the gold precursor (4) and/or the GO to rGO (5) (Figure 8).
CeO2 + hν(solar) → CeO2(e + h+)
CeO2(e + h+) + s.agent → CeO2(e) + s.agent+
CeO2(e) + Au3+ → CeO2 + Au
CeO2(e) + GO → CeO2+ rGO
In the case of photoreforming after reactions (2) and (3), the evolution of H2 was obtained by means of the same excited electrons of the CB of ceria and by the presence of glycerol (6) and (7).
C3H8O3 + 3H2O + 2h+CeO2 → 3CO2 + 6H2+2H+
CeO2(e) + 2H+ → CeO2+ H2
Interestingly, while for the photoreforming the presence of gold resulted in an important increase in H2 evolution, the addition of GO or rGO had the same effect (Figure 2). As detected by the Raman measurements, reaction (5) was favoured after subsequent runs of continuous solar photoreforming reaction, therefore the reduction of GO into rGO further proceeded together to the reactions (6) and (7). As a result, the activity of the CeO2-rGO-Au increased (Figure 3) from 210 μmol gcat−1 h−1 in the first run to 270 μmol gcat−1 h−1 in the fifth run. This behaviour was different compared to that reported for conventional photocatalysts [38,39], which after different consecutive runs maintained the activity or were deactivated due to the occurrence of the e/h+ recombination.
This is a fascinating result, with the obtained H2 that was higher than other CeO2-based systems reported in the literature for solar photoreforming (Table 3), even though the different experimental set-ups employed by the various research groups must be considered.
This one-pot material synthesis with a progressive performance improvement during the photocatalytic reaction can open new ways of designing solar-light-driven photocatalysts that can enhance their photoactivity in situ through the reaction. This versatile approach can also be useful for possible scale-up applications.

5. Conclusions

A solar one-pot photoreduction method was investigated in this work for the synthesis of cerium dioxide-graphene oxide-based systems. The obtained materials were used to catalyse the solar glycerol photoreforming for green H2 production. The photocatalytic activity data highlighted an increase in the H2 evolution due to the addition of graphene oxide or reduced graphene oxide to ceria, further boosted by the presence of gold as a co-catalyst. Furthermore, the strong interaction between this metal and CeO2 led to excellent stability performances of the CeO2-rGO-Au sample, with a progressive increase in the H2 production over the runs ascribed to a contextual enhancement of the amount of GO reduced into rGO. The presence of gold is fundamental for the enhancement of the H2 production due to the exploitation of the surface plasmon resonance effect with a contextual better use of the visible light portion of the solar irradiation, whereas the presence of GO and rGO sheets promoted a good dispersion of the Au nanoparticles. In conclusion, in the glycerol photoreforming, the synergism between the photocatalytic features of CeO2, the visible-light-induced properties of gold and the presence of GO and rGO allowed to obtain versatile photocatalysts activated by solar irradiation with an improved photoactivity induced by the continuous light absorption and by the subsequent photocatalytic runs. This can be a starting point to propose new solar-driven photocatalysts for sustainable hydrogen production.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma16020747/s1, Figure S1: FTIR spectra of the CeO2 samples before and after the calcination; Figure S2: Raman signal deconvolution of D and G bands in their five different components for the bare GO as representative sample; Figure S3: N2 adsorption–desorption isotherms and pore size distribution of (a) CeO2, (b) CeO2-Au, (c) CeO2-GO-Au and (d) CeO2-rGO-Au; Figure S4: XPS spectra of (a) Ce 3d of bare CeO2, (b) O 1s of bare CeO2, (c) C 1s of CeO2-rGO-Au and (d) Au 4f of CeO2-rGO-Au; Table S1: XPS signal deconvolution. The values are in binding energy (eV).

Author Contributions

Conceptualization, S.A.B.; investigation, S.A.B., E.L.G. and M.C.P.; writing—original draft preparation S.A.B. and R.F.; writing—review and editing, S.A.B., R.F. and S.S.; supervision, R.F. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available on request to the corresponding author.

Acknowledgments

R.F. thanks the STARTING GRANT (PIACERI 2020–2022) support of the University of Catania. S.A.B. thanks the research group of “Sustainable Organic Chemistry” of the University of Córdoba (Spain). S.S. thanks the C.I.R.C.C. (University Consortium in Chemical Reactivity and Catalysis) for the support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pruvost, F.; Cloete, S.; Arnaiz del Pozo, C.; Zaabout, A. Blue, Green, and Turquoise Pathways for Minimizing Hydrogen Production Costs from Steam Methane Reforming with CO2 Capture. Energy Convers. Manag. 2022, 274, 116458. [Google Scholar] [CrossRef]
  2. Banerjee, D.; Kushwaha, N.; Shetti, N.P.; Aminabhavi, T.M.; Ahmad, E. Green Hydrogen Production via Photo-Reforming of Bio-Renewable Resources. Renew. Sustain. Energy Rev. 2022, 167, 112827. [Google Scholar] [CrossRef]
  3. Bhakta, A.K.; Fiorenza, R.; Jlassi, K.; Mekhalif, Z.; Ali, A.M.A.; Chehimi, M.M. The Emerging Role of Biochar in the Carbon Materials Family for Hydrogen Production. Chem. Eng. Res. Des. 2022, 188, 209–228. [Google Scholar] [CrossRef]
  4. Christoforidis, K.C.; Fornasiero, P. Photocatalytic Hydrogen Production: A Rift into the Future Energy Supply. ChemCatChem 2017, 9, 1523–1544. [Google Scholar] [CrossRef] [Green Version]
  5. 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]
  6. Lewicka, E.; Guzik, K.; Galos, K. On the Possibilities of Critical Raw Materials Production from the EU’s Primary Sources. Resources 2021, 10, 50. [Google Scholar] [CrossRef]
  7. Fiorenza, R.; Balsamo, S.A.; Condorelli, M.; D’Urso, L.; Compagnini, G.; Scirè, S. Solar Photocatalytic H2 Production over CeO2-Based Catalysts: Influence of Chemical and Structural Modifications. Catal. Today 2021, 380, 187–198. [Google Scholar] [CrossRef]
  8. Sujay, S.G.C.; Alkanad, K.; Alnaggar, G.; Al-Zaqri, N.; Bajiri, M.A.; Thejaswini, B.; Dhileepan, M.D.; Neppolian, B.; Lokanath, N.K. Surface Defect-Engineered CeO2−x by Ultrasound Treatment for Superior Photocatalytic H2 Production and Water Treatment. Catal. Sci. Technol. 2022, 12, 2071–2083. [Google Scholar] [CrossRef]
  9. Hou, H.; Yamada, H.; Nitta, A.; Murakami, Y.; Saito, N. Efficient Separation of Photoexcited Charge at Interface between Pure CeO2 and Y3+-Doped CeO2 with Heterogonous Doping Structure for Photocatalytic Overall Water Splitting. Materials 2021, 14, 350. [Google Scholar] [CrossRef]
  10. Liu, X.; Liu, R.; Li, T.; Liu, Y.; Liu, L.; Lyu, K.; Shah, S.P. Research on the Anticorrosion Properties of CeO2-GO/EP Nanocomposite Coating in Simulated Sea Water. Polymers 2021, 13, 2072. [Google Scholar] [CrossRef]
  11. Liu, F.; Wang, X.; Zhang, L.; Persson, K.M.; Chen, B.-Y.; Hsu, Y.; Chang, C.T. Near-Visible-Light-Driven Noble Metal-Free of Reduced Graphene Oxide Nanosheets over CeO2 Nanowires for Hydrogen Production. J. Taiwan Inst. Chem. Eng. 2020, 107, 139–151. [Google Scholar] [CrossRef]
  12. Balsamo, S.A.; Fiorenza, R.; Condorelli, M.; Pecoraro, R.; Brundo, M.V.; Presti, F.L.; Sciré, S. One-Pot Synthesis of TiO2-RGO Photocatalysts for the Degradation of Groundwater Pollutants. Materials 2021, 14, 5938. [Google Scholar] [CrossRef] [PubMed]
  13. Nemati, F.; Rezaie, M.; Tabesh, H.; Eid, K.; Xu, G.; Ganjali, M.R.; Hosseini, M.; Karaman, C.; Erk, N.; Show, P.-L.; et al. Cerium Functionalized Graphene Nano-Structures and Their Applications; A Review. Environ. Res. 2022, 208, 112685. [Google Scholar] [CrossRef] [PubMed]
  14. Ganguly, S.; Mondal, S.; Das, P.; Bhawal, P.; Das, T.K.; Ghosh, S.; Remanan, S.; Das, N.C. An Insight Into the Physico-Mechanical Signatures of Silylated Graphene Oxide in Poly(Ethylene Methyl Acrylate) Copolymeric Thermoplastic Matrix. Macromol. Res. 2019, 27, 268–281. [Google Scholar] [CrossRef]
  15. Rumayor, M.; Corredor, J.; Rivero, M.J.; Ortiz, I. Prospective Life Cycle Assessment of Hydrogen Production by Waste Photoreforming. J. Clean. Prod. 2022, 336, 130430. [Google Scholar] [CrossRef]
  16. Balsamo, S.A.; Sciré, S.; Condorelli, M.; Fiorenza, R. Photocatalytic H2 Production on Au/TiO2: Effect of Au Photodeposition on Different TiO2 Crystalline Phases. J 2022, 5, 92–104. [Google Scholar] [CrossRef]
  17. Rueda-Navarro, C.M.; Ferrer, B.; Baldoví, H.G.; Navalón, S. Photocatalytic Hydrogen Production from Glycerol Aqueous Solutions as Sustainable Feedstocks Using Zr-Based UiO-66 Materials under Simulated Sunlight Irradiation. Nanomaterials 2022, 12, 3808. [Google Scholar] [CrossRef]
  18. Corredor, J.; Rivero, M.J.; Rangel, C.M.; Gloaguen, F.; Ortiz, I. Comprehensive Review and Future Perspectives on the Photocatalytic Hydrogen Production. J. Chem. Technol. Biotechnol. 2019, 94, 3049–3063. [Google Scholar] [CrossRef] [Green Version]
  19. Wang, G.; Mu, Q.; Chen, T.; Wang, Y. Synthesis, Characterization and Photoluminescence of CeO2 Nanoparticles by a Facile Method at Room Temperature. J. Alloy. Compd. 2010, 493, 202–207. [Google Scholar] [CrossRef]
  20. Ahmad, T.; Shahazad, M.; Ubaidullah, M.; Ahmed, J. Synthesis, Characterization and Dielectric Properties of TiO2-CeO2 Ceramic Nanocomposites at Low Titania Concentration. Bull. Mater. Sci. 2018, 41, 99. [Google Scholar] [CrossRef]
  21. Mahmoud, M.A.; Chamanzar, M.; Adibi, A.; El-Sayed, M.A. Effect of the Dielectric Constant of the Surrounding Medium and the Substrate on the Surface Plasmon Resonance Spectrum and Sensitivity Factors of Highly Symmetric Systems: Silver Nanocubes. J. Am. Chem. Soc. 2012, 134, 6434–6442. [Google Scholar] [CrossRef] [PubMed]
  22. García-López, E.I.; Abbasi, Z.; Parrino, F.; La Parola, V.; Liotta, L.F.; Marcì, G. Au/CeO2 Photocatalyst for the Selective Oxidation of Aromatic Alcohols in Water under Uv, Visible and Solar Irradiation. Catalysts 2021, 11, 1467. [Google Scholar] [CrossRef]
  23. Ma, L.; Wang, D.; Li, J.; Bai, B.; Fu, L.; Li, Y. Ag/CeO2 Nanospheres: Efficient Catalysts for Formaldehyde Oxidation. Appl. Catal. B Environ. 2014, 148–149, 36–43. [Google Scholar] [CrossRef]
  24. Ramesh, K.; Gnanavel, B.; Shkir, M. Enhanced Visible Light Photocatalytic Degradation of Bisphenol A (BPA) by Reduced Graphene Oxide (RGO)–Metal Oxide (TiO2, ZnO and WO3) Based Nanocomposites. Diam. Relat. Mater. 2021, 118, 108514. [Google Scholar] [CrossRef]
  25. Qian, J.; Chen, Z.; Sun, H.; Chen, F.; Xu, X.; Wu, Z.; Li, P.; Ge, W. Enhanced Photocatalytic H2 Production on Three-Dimensional Porous CeO2/Carbon Nanostructure. ACS Sustain. Chem. Eng. 2018, 6, 9691–9698. [Google Scholar] [CrossRef]
  26. Ferrari, A.C.; Meyer, J.C.; Scardaci, V.; Casiraghi, C.; Lazzeri, M.; Mauri, F.; Piscanec, S.; Jiang, D.; Novoselov, K.S.; Roth, S.; et al. Raman Spectrum of Graphene and Graphene Layers. Phys. Rev. Lett. 2006, 97, 187401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Jorio, A.; Ferreira, E.H.M.; Moutinho, M.V.O.; Stavale, F.; Achete, C.A.; Capaz, R.B. Measuring Disorder in Graphene with the G and D Bands. Phys. Status Solidi (B) Basic Res. 2010, 247, 2980–2982. [Google Scholar] [CrossRef]
  28. Sotomayor, F.; Quantatec, A.P.; Sotomayor, F.J.; Cychosz, K.A.; Thommes, M. Characterization of Micro/Mesoporous Materials by Physisorption: Concepts and Case Studies. Acc. Mater. Surf. Res 2018, 3, 34–50. [Google Scholar]
  29. Fiorenza, R.; Balsamo, S.A.; D’urso, L.; Sciré, S.; Brundo, M.V.; Pecoraro, R.; Scalisi, E.M.; Privitera, V.; Impellizzeri, G. CeO2 for Water Remediation: Comparison of Various Advanced Oxidation Processes. Catalysts 2020, 10, 446. [Google Scholar] [CrossRef] [Green Version]
  30. Aslam, M.; Qamar, M.T.; Soomro, M.T.; Ismail, I.M.I.; Salah, N.; Almeelbi, T.; Gondal, M.A.; Hameed, A. The Effect of Sunlight Induced Surface Defects on the Photocatalytic Activity of Nanosized CeO2 for the Degradation of Phenol and Its Derivatives. Appl. Catal. B Environ. 2016, 180, 391–402. [Google Scholar] [CrossRef]
  31. Sing, K.S.W.; Everett, D.H.; Haul, R.A.W.; Moscou, L.; Pierotti, R.A.; Rouquerol, J.; Siemieniewska, T. Reporting Physisorption Data for Gas/Solid Systems with Special Reference to the Determination of Surface Area and Porosity. Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  32. Holgado, J.P.; Alvarez, R.; Munuera, G. Study of CeO2 XPS Spectra by Factor Analysis: Reduction of CeO2. Appl. Surf. Sci. 2000, 161, 301–315. [Google Scholar] [CrossRef]
  33. Natile, M.M.; Glisenti, A. Nanostructured CeO2 Powders by XPS. Surf. Sci. Spectra 2006, 13, 17–30. [Google Scholar] [CrossRef]
  34. Chang, A.; Peng, W.-S.; Tsai, I.-T.; Chiang, L.-F.; Yang, C.-M. Efficient Hydrogen Production by Selective Alcohol Photoreforming on Plasmonic Photocatalyst Comprising Sandwiched Au Nanodisks and TiO2. Appl. Catal. B Environ. 2019, 255, 117773. [Google Scholar] [CrossRef]
  35. Morales, M.V.; Asedegbega-Nieto, E.; Castillejos-López, E.; Bachiller-Baeza, B.; Guerrero-Ruiz, A. Difference in the Deactivation of Au Catalysts during Ethanol Transformation When Supported on ZnO and on TiO2. RSC Adv. 2018, 8, 7473–7485. [Google Scholar] [CrossRef] [Green Version]
  36. Nath, M.P.; Biswas, S.; Nath, P.; Choudhury, B. Synergy of Adsorption and Plasmonic Photocatalysis in the Au–CeO2 Nanosystem: Experimental Validation and Plasmonic Modeling. Langmuir 2022, 38, 7628–7638. [Google Scholar] [CrossRef]
  37. Fiorenza, R.; Bellardita, M.; D’Urso, L.; Compagnini, G.; Palmisano, L.; Scirè, S. Au/TiO2-CeO2 Catalysts for Photocatalytic Water Splitting and VOCs Oxidation Reactions. Catalysts 2016, 6, 121. [Google Scholar] [CrossRef] [Green Version]
  38. Sundaram, I.M.; Kalimuthu, S.; Priya, P.G.; Sekar, K.; Rajendran, S. Hierarchical TiO2 Spheroids Decorated g-C3N4 Nanocomposite for Solar Driven Hydrogen Production and Water Depollution. Int. J. Hydrogen Energy 2022, 47, 3709–3721. [Google Scholar] [CrossRef]
  39. Corredor, J.; Rivero, M.J.; Ortiz, I. New Insights in the Performance and Reuse of RGO/TiO2 Composites for the Photocatalytic Hydrogen Production. Int. J. Hydrogen Energy 2021, 46, 17500–17506. [Google Scholar] [CrossRef]
  40. Tian, B.; Yang, B.; Li, J.; Li, Z.; Zhen, W.; Wu, Y.; Lu, G. Water Splitting by CdS/Pt/WO3-CeOx Photocatalysts with Assisting of Artificial Blood Perfluorodecalin. J. Catal. 2017, 350, 189–196. [Google Scholar] [CrossRef]
  41. Yadav, A.A.; Hunge, Y.M.; Kang, S.-W. Visible Light-Responsive CeO2/MoS2 Composite for Photocatalytic Hydrogen Production. Catalysts 2022, 12, 1185. [Google Scholar] [CrossRef]
  42. Tong, R.; Sun, Z.; Zhong, X.; Wang, X.; Xu, J.; Yang, Y.; Xu, B.; Wang, S.; Pan, H. Enhancement of Visible-Light Photocatalytic Hydrogen Production by CeCO3OH in g-C3N4/CeO2 System. ChemCatChem 2019, 11, 1069–1075. [Google Scholar] [CrossRef]
  43. Dong, B.; Li, L.; Dong, Z.; Xu, R.; Wu, Y. Fabrication of CeO2 Nanorods for Enhanced Solar Photocatalysts. Int. J. Hydrogen Energy 2018, 43, 5275–5282. [Google Scholar] [CrossRef]
Figure 1. H2 evolution considering the different rGO content on CeO2-based samples. All the samples contain gold at 1 wt% as co-catalyst.
Figure 1. H2 evolution considering the different rGO content on CeO2-based samples. All the samples contain gold at 1 wt% as co-catalyst.
Materials 16 00747 g001
Figure 2. H2 production over the different CeO2-based samples.
Figure 2. H2 production over the different CeO2-based samples.
Materials 16 00747 g002
Figure 3. Photocatalytic activity of the CeO2-rGO-Au sample over five subsequent runs.
Figure 3. Photocatalytic activity of the CeO2-rGO-Au sample over five subsequent runs.
Materials 16 00747 g003
Figure 4. (a) UV-DRS of the samples and (b) their TAUC plots with the corresponding linear fitting (dashed lines) for the estimation of the band gap.
Figure 4. (a) UV-DRS of the samples and (b) their TAUC plots with the corresponding linear fitting (dashed lines) for the estimation of the band gap.
Materials 16 00747 g004
Figure 5. Raman spectra of the samples. The discussed signals are highlighted by dashed lines.
Figure 5. Raman spectra of the samples. The discussed signals are highlighted by dashed lines.
Materials 16 00747 g005
Figure 6. STEM micrograph of (a) CeO2-rGO-Au sample and its elemental mapping; (b) Ce; (c) Au and (d) O. Panel (e) is a micrograph in which it is evident how the CeO2-Au nanoparticles are deposited in a GO sheet distinguishable for the delineated edges.
Figure 6. STEM micrograph of (a) CeO2-rGO-Au sample and its elemental mapping; (b) Ce; (c) Au and (d) O. Panel (e) is a micrograph in which it is evident how the CeO2-Au nanoparticles are deposited in a GO sheet distinguishable for the delineated edges.
Materials 16 00747 g006
Figure 7. (a) Ce 3d and (b) O 1s XPS spectra of CeO2-rGO-Au. The deconvolution curves of the single components are depicted by dashed/dotted lines.
Figure 7. (a) Ce 3d and (b) O 1s XPS spectra of CeO2-rGO-Au. The deconvolution curves of the single components are depicted by dashed/dotted lines.
Materials 16 00747 g007
Figure 8. Representative scheme of the interactions in the sample CeO2-rGO-Au.
Figure 8. Representative scheme of the interactions in the sample CeO2-rGO-Au.
Materials 16 00747 g008
Table 1. Raman parameters related to the D and G bands.
Table 1. Raman parameters related to the D and G bands.
SampleID/IGFWHM (cm−1)
GO0.88166D, 83G
CeO2-GO0.99188D, 90G
CeO2-rGO0.99136D, 76G
CeO2-rGO-Au0.99159D, 84G
CeO2-rGO-Au after 5 runs1.02130D, 70G
Table 2. Textural properties: type of N2 isotherms, BET surface area and mean pore size.
Table 2. Textural properties: type of N2 isotherms, BET surface area and mean pore size.
SampleIsothermSBET (m2 g−1)dP (nm)
CeO2 IV; H2 81 ± 2 24
CeO2-Au IV; H2 72 ± 2 8
CeO2-GO-Au IV; H2 69 ± 2 17
CeO2-rGO-Au IV; H2 70 ± 2 18
Table 3. Comparison among several CeO2-based systems used for the solar photocatalytic H2 production.
Table 3. Comparison among several CeO2-based systems used for the solar photocatalytic H2 production.
SampleSynthesisPhotoreforming ConditionsIrradiation SourceH2 Evolution (μmol/gcat h)Ref.
CeO2-rGO-Au
(after 5 runs)
Photoreduction10% (v/v) solution of glycerol in MilliQ water G150 W Xe lamp
solar simulator
270this work
CdS/Pt/WO3-CeOxPrecipitationIn situ material synthesis and H2 evolution300 W Xe lamp, with a 420 nm cutoff filter50[40]
CeO2/C quantum dotsBio-template strategy100 mL of water containing 10 vol% methanol300 W, 50 mW/cm2 Xe lamp60[25]
CeO2/MoS2Hydrothermal100 mL aqueous solution containing 0.3 M Na2SO3/Na2S150 W xenon arc lamp (>400 nm)125[41]
C3N4/CeCO3OH/CeO2Hydrothermal80 mL water including 10 vol% TEOAAM 1.5G solar light40[42]
CeO2 nanorodsHydrothermal100 mL of an aqueous solution containing 0.43 M Na2S mixed with 0.50 M Na2SO3300 W Xe lamp5[43]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Balsamo, S.A.; La Greca, E.; Calà Pizzapilo, M.; Sciré, S.; Fiorenza, R. CeO2-rGO Composites for Photocatalytic H2 Evolution by Glycerol Photoreforming. Materials 2023, 16, 747. https://doi.org/10.3390/ma16020747

AMA Style

Balsamo SA, La Greca E, Calà Pizzapilo M, Sciré S, Fiorenza R. CeO2-rGO Composites for Photocatalytic H2 Evolution by Glycerol Photoreforming. Materials. 2023; 16(2):747. https://doi.org/10.3390/ma16020747

Chicago/Turabian Style

Balsamo, Stefano Andrea, Eleonora La Greca, Marta Calà Pizzapilo, Salvatore Sciré, and Roberto Fiorenza. 2023. "CeO2-rGO Composites for Photocatalytic H2 Evolution by Glycerol Photoreforming" Materials 16, no. 2: 747. https://doi.org/10.3390/ma16020747

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