Next Article in Journal
Prolonged Antibacterial Activity in Tannic Acid–Iron Complexed Chitosan Films for Medical Device Applications
Previous Article in Journal
Nanomedicine: New Frontiers in Fighting Microbial Infections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Tandem Control of La-Doping and CuO-Heterojunction on SrTiO3 Perovskite by Double-Nozzle Flame Spray Pyrolysis: Selective H2 vs. CH4 Photocatalytic Production from H2O/CH3OH

by
Pavlos Psathas
1,
Areti Zindrou
1,
Christina Papachristodoulou
1,
Nikos Boukos
2 and
Yiannis Deligiannakis
1,*
1
Department of Physics, University of Ioannina, 45110 Ioannina, Greece
2
Institute of Nanoscience and Nanotechnology (INN), NCSR Demokritos, 15310 Athens, Greece
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(3), 482; https://doi.org/10.3390/nano13030482
Submission received: 30 December 2022 / Revised: 19 January 2023 / Accepted: 21 January 2023 / Published: 25 January 2023
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)

Abstract

:
ABO3 perovskites offer versatile photoactive nano-templates that can be optimized towards specific technologies, either by means of doping or via heterojunction engineering. SrTiO3 is a well-studied perovskite photocatalyst, with a highly reducing conduction-band edge. Herein we present a Double-Nozzle Flame Spray Pyrolysis (DN-FSP) technology for the synthesis of high crystallinity SrTiO3 nanoparticles with controlled La-doping in tandem with SrTiO3/CuO-heterojunction formation. So-produced La:SrTiO3/CuO nanocatalysts were optimized for photocatalysis of H2O/CH3OH mixtures by varying the La-doping level in the range from 0.25 to 0.9%. We find that, in absence of CuO, the 0.9La:SrTiO3 material achieved maximal efficient photocatalytic H2 production, i.e., 12 mmol g−1 h−1. Introduction of CuO on La:SrTiO3 enhanced selective production of methane CH4. The optimized 0.25La:SrTiO3/0.5%CuO catalyst achieved photocatalytic CH4 production of 1.5 mmol g−1 h−1. Based on XRD, XRF, XPS, BET, and UV-Vis/DRS data, we discuss the photophysical basis of these trends and attribute them to the effect of La atoms in the SrTiO3 lattice regarding the H2-production, plus the effect of interfacial CuO on the promotion of CH4 production. Technology-wise this work is among the first to exemplify the potential of DN-FSP for scalable production of complex nanomaterials such as La:SrTiO3/CuO with a diligent control of doping and heterojunction in a single-step synthesis.

Graphical Abstract

1. Introduction

Photocatalytic storage of light photons in the form of fuels such as H2 or CH4 is globally envisaged as an environmentally benign technology [1,2]. To this front, in the last decades, photocatalytic semiconductors receive great consideration [3].
Strontium Titanate (SrTiO3) has a classical ABO3 perovskite structure, with the ideal cubic lattice [4]. SrTiO3 has received enormous attention as a photocatalyst, oxidative catalyst, or catalyst-support, due to beneficial characteristics such as low price [5], chemical stability, and excellent thermal stability, e.g., melting point as high as 2080 °C, with carbon and sulfur tolerance, adaptability and modifiable oxidative properties [6]. SrTiO3 has a highly reducing conduction-band-edge energy position (ECB) of −1.2 eV vs. NHE [7,8], which makes it a highly efficient photocatalyst for hydrogen production from H2O [9]. SrTiO3 has the disadvantage of a broad 3.2 eV band gap, permitting only the absorption of UV photons, thus to remediate this drawback, a common technique is to apply dopants or heterojunction with other materials [10].
So far, synthesis of nanocrystalline SrTiO3 has been achieved with various methods, each having distinct advantages/disadvantages [4,11], e.g., solid phase [12] that is a low-cost approach, however, produces rather large particles, sol-gel [11] that gives pure phase small nanoparticles at a higher cost, hydrothermal [5] that produces controllable size but requires long synthesis-times with impurity possibility. Moreover, doping of SrTiO3 with an appropriate dopant-cation at low concentration, and the formation of heterojunctions between SrTiO3 and selected co-catalytic particles, pose additional hurdles, complexity, and cost of the synthesis process [13]. In this context, it is of great importance to develop synthesis technologies for high-quality SrTiO3 in scale-up production, at the same time allowing doping of SrTiO3 and heterojunction SrTiO3/MOx formation in one step.
The Flame Spray Pyrolysis (FSP) process is an aerosol combustion technology for nanoparticle synthesis [14,15,16] that is well-established for industrial-scale production [17,18]. As such, FSP is already used in industry to produce several commonly used metal oxides [19]. FSP utilizes high-temperature combustion, i.e., 1000–2000 K [20], of a precursor containing metal atoms dispersed in an organic solvent, that is sprayed in the form of aerosol droplets. Appropriate selection of precursor mixtures and process parameters allows precise control of the crystal phases [19], devoid of chemical leftovers. So far, FSP has been established for the successful production of a wide range of single-metal oxides [19], however, establishing FSP process parameters for the production of ABO3 perovskite nanomaterials is typically more challenging, e.g., such as avoiding the formation of the separate oxides AOn and BOm. The relative importance of process parameters has been exemplified in our recent works on FSP-engineering of NaTaO3 [21], Bi2Fe4O9 and BiFeO3 [22,23], W-, Zr-Doped BiVO4 [24], and by Kudo and Amal for BiVO4 [25], as well as Abe and Laine for La2Ti2O7 [26]. Very recently, the synthesis of SrTiO3 [27,28] by FSP has been achieved and used for catalytic combustion of CO and CH4, however with no reference to photocatalytic evaluation or optimization.
Lanthanum (La) doping of SrTiO3 has been shown to increase the electrochemical efficiency of the perovskite [29,30], and as co-dopant with other atoms [31,32]. Domen [33,34] and Kudo [35,36] have demonstrated that SrTiO3 co-doped with La and Rh exhibits high solar-to-hydrogen energy conversion efficiency for water splitting to H2. Moreover, selected cations, e.g., La3+, Ce3+, or Nitrogen [8,37] allow control of structural/photo/electronic properties of doped-SrTiO3, enhancing the photocatalytic efficiency [38,39]. Thus, SrTiO3 offers a versatile template, that can be optimized towards specific technologies either via doping or heterojunction engineering.
Surficial heterojunctions of SrTiO3 with pertinent metal-oxides allow control of the selectivity of surface reactions. More specifically, Cu-atoms and their particles (CuO, Cu2O, Cu0) are particularly attractive in photocatalytic processes, since they have been shown to control selectivity towards specific products, i.e., Cu2O exhibits selectivity towards CH3OH production and Cu0 towards hydrocarbons and C2 products [40,41]. In particular, studies show that CuO clusters on the TiO2 surface can serve as an efficient co-catalyst to enhance H2 production [42,43]. Meanwhile, Cu2O on TiO2 was also reported to be more suitable for H2 production, which was attributed to the better energy band alignment between TiO2 and Cu2O, which facilitates charge separation [44,45].
Herein, we employ a Single-Nozzle FSP (SN-FSP) process for the synthesis of La:SrTiO3 and a Double-Nozzle FSP (DN-FSP) creating a heterojunction of CuO finely dispersed on La:SrTiO3 in one-step [46,47], see Figure 1. Recently, our group utilized DN-FSP to produce TiO2 loaded with noble metal particles (>2 nm), to significantly enhance photocatalytic hydrogen production [48]. In the present work, our hypothesis was that an in tandem La-doping of SrTiO3 heterojunctions with CuO can allow control of the SrTiO3 electronic/photocatalytic properties, as well as controlled selectivity in photocatalytic products. In this context, our specific aims were: (i) To develop an FSP protocol for the synthesis of La:SrTiO3 with controlled La-doping; (ii) To advance the FSP protocol to allow heterojunction of CuO on La:SrTiO3; (iii) To study the H2 vs. CH4 photocatalytic process in comparison to CuO. For this reason, we have performed photocatalytic studies in an H2O/CH3OH mixture. CH3OH is a well-known hole-scavenger [49], however, most publications do not examine the CH3OH involvement in the reaction path and the final products, e.g., eventually CH4.

2. Materials and Methods

2.1. Flame Spray Pyrolysis (FSP) Synthesis of SrTiO3 Nanoparticles

The precursor solution contains Strontium acetate (STREM), Titanium(VI) isopropoxide (97%, Aldrich) for the synthesis of perovskite SrTiO3. For the deposition of cocatalytic metals, i.e., La atoms, with Lanthanum Acetylacetonate (97%, STREM). The Sr and Ti precursors were dispersed in a mixture of Acetic acid and Xylene (1:1 volume ratio), while for the La precursor a mixture of toluene and 2-EHA (1:1 volume ratio) that consisted of approximately 10% of the total volume of the final precursor solution. We underline that it is the control of FSP parameters that allows Sr- and Ti- to be engaged exclusively in the formation of the SrTiO3 crystals, while La- is introduced as a lattice-dopant.
Double-Nozzle FSP: In the DN-FSP as shown in Figure 1, two FSP-nozzles operate in tandem that are asymmetrically positioned, so that the system generates two different kinds of nanomaterials by controlling the properties of each nozzle independently, where the left FSP-nozzle contained the Sr/Ti/La precursors, while the right FSP-nozzle contained Cu(NO3)2·3H2O (Supelco) in Acetonitrile:Ethyleglygol (1:1 volume ratio). Screening experiments were conducted to find the preferent geometric parameters: α1 nozzle angle at 20o, and α2 = 20o. The internozzle distance was placed at x = 8 cm and the vertical intersection distance of the two flames, above the nozzle, was b = 10, see Figure 1C.
Figure 1. (A) Description of the Double-Nozzle FSP (DN-FSP) set-up, with the SrTiO3 formation in the left-side nozzle and the CuO formation in the right-side nozzle. (B) Schematic depiction of the structure of the La:SrTiO3/CuO particle. (C) The geometric parameters for the two nozzles in DN-FSP.
Figure 1. (A) Description of the Double-Nozzle FSP (DN-FSP) set-up, with the SrTiO3 formation in the left-side nozzle and the CuO formation in the right-side nozzle. (B) Schematic depiction of the structure of the La:SrTiO3/CuO particle. (C) The geometric parameters for the two nozzles in DN-FSP.
Nanomaterials 13 00482 g001
The FSP parameters for the synthesis of the nanoparticles consisted of an oxygen dispersion flow rate of D = 5 L min−1 (Linde 99.999%) and a precursor flow rate of P = 5 mL min−1. These D and P rates were also used for both nozzles in the DN-FSP setup. The pilot flame was ignited by premixed O2 and CH4 (4 L min−1, 2 L min−1). In the Single-Nozzle as well as in the Double-Nozzle FSP, with the assistance of a vacuum pump (BUSCH), the produced particles were deposited on a glass microfiber filter with a binder (Albet Labscience GF_6_257) and collected by scrubbing the nanoparticles from the filter. The nanomaterials were collected in glass vials under an inert Argon atmosphere, until use.
For convenience, herein the produced materials, listed in Table 1, are codenamed as follows: X_La:STO/Y_Cu where STO = SrTiO3, X the nominal % La-content per weight of SrTiO3, Y = the nominal % Cu-content per weight of SrTiO3. In this way, we have prepared 0.9La:STO and 0.35La:STO, respectively, listed in Table 1. For the DN-FSP experiments, we have produced STO/2Cu, STO/1.2Cu, and STO/0.5Cu, respectively. Finally the 0.25La:STO/0.5Cu produced by DN-FSP is codenamed as La:STO/Cu.

2.2. Photocatalytic Evaluation

Photocatalytic experiments were performed in a double-walled photochemical reactor (Toption instrument co. Ltd.) with a total reaction volume of 340 mL, at a temperature of 25 ± 2 °C, controlled by a recirculation chiller cooling system. The UV source was a 250 W Mercury lamp, positioned at the geometrical center of the photoreactor inside the quartz-immersion well. The irradiation power at the experimental mean distance of 3 cm was 0.5 W cm−2 as measured with a power meter (Newport model, 1918-C).
A Gas-Chromatography System combined with a Thermal-Conductivity-Detector (TCD-Shimadzu GC-2014, Carboxen 1000 column, Ar carrier gas) was used to identify and quantify the produced H2 and CH4 gases. In each experiment, 50 mg of the catalyst was suspended in 200 mL Milli-Q water and 50 mL methanol (20% per volume) as a hole-scavenger. Photo-deposition of Pt-cocatalyst was implemented to increase the photocatalytic production, using hydrogen hexachloroplatinate (IV) hydrate, (H2Pt4Cl6·H2O, 99.9%, Alfa Aesar). The error bars of approximately 8%, for the photocatalytic products H2 and CH4, reflect the detection uncertainty and statistical deviation after three catalytic runs.

2.3. Characterization of Materials

XRD measurements were carried out to identify the crystal phase identification and structural properties of the materials, using a Bruker D8-Advance diffractometer with a Cu source (Kα, λ = 1,5418 Å), with operation parameters of 40 KV generator voltage and 40 mA current. The particle crystal size (dXRD) as obtained from the XRD data was calculated with the Scherrer Equation (1)
d XRD = K   λ FWHM ×   cos θ
where K = 0.9, λ = 1,5418 Å, and FWHM is the full-width at half-maximum of the XRD peaks [50].
Transmission Electron Microscopy analysis was used to map the particle morphology. The amount of La-atoms and Cu-atoms of the nanomaterials was calculated by a home-built Energy-Dispersive X-Ray Fluorescence (EDXRF), and an annular 241 Am radioisotopic source was used for sample excitation. The source is fixed coaxially above a CANBERRA SL80175 Si(Li) detector (5 mm crystal thickness, 80 mm2 area), with a 25 μm thick Be window and an energy resolution of 171 eV for the 5.9 keV Mn Kα line. Spectral analysis was carried out using the WinQxas software package (International Atomic Energy Agency, 1997–2002). Quantitative analysis was based on in-house standard samples and the construction of calibration curves.
The morphology and phase composition of the La:STO/Cu particle was studied using an FEI Talos F200i field-emission (scanning) transmission electron microscope (Thermo Fisher Scientific Inc., Waltham, MA, USA) operating at 200 kV, equipped with a windowless energy-dispersive spectroscopy microanalyzer (6T/100 Bruker, Hamburg, Germany).
The Specific Surface Area (SSA) and the pore size of the nanomaterials were measured by a Quantachrome NOVAtouch_LX2 to record the N2 adsorption-desorption isotherms at 77 K. The SSA was calculated using the absorption data points in the range of 0.1−0.3 relative pressure P/Po. While the pore radius analysis was obtained by the BJH method [51] in the range of 0.35–0.99 P/Po.
Diffuse-Reflectance UV-Vis absorption spectra were recorded with a Perkin Elmer Lambda-35 spectrometer with BaSO4 powder used as a background standard, operating at room temperature for the wavelength range of 200–800 nm, while the band gap energy (Eg) was calculated using the Kubelka–Munk method [52].
The oxidation state of the Sr, Ti, O, and Cu atoms was monitored by X-ray photoelectron spectroscopy (XPS), using a SPECS spectrometer equipped with a twin Al-Mg anode X-ray source and a multi-channel hemispherical sector electron analyzer (HSA-Phoibos 100), a monochromatized Mg Kα line at 1253.6eV, analyzer pass-energy at 15 eV, and the base pressure at 2–5 × 10−9 mbar. The binding energies were determined vs. the energy of C1s carbon peak at 284.5 eV. The peak deconvolution was performed employing mixed Gaussian–Lorentzian functions, using WinSpec software, developed at the Laboratoire Interdisciplinaire de Spectroscopie Electronique, University of Namur, Belgium.

3. Results

3.1. SrTiO3-Based Perovskite Nanoparticle Synthesis by FSP

Highly crystalline SrTiO3 has been successfully produced by Single-Nozzle FSP, see XRD data in Figure 2A. The distinct peaks 2θ at 22.71, 32.34, 39.91, 46.43, 52.29, 57.74, 67.78, and 77.13 degrees, confirm the formation of pure cubic perovskite SrTiO3 structure (JCPDS74–1296). All diffraction peaks are assigned to Miller indices of the (100), (110), (111), (200), (210), (211), (220), and (310) planes for the cubic perovskite structure (Pm3m), with no traces of other crystalline byproducts, such as TiO2 or Sr-oxide.
The dXRD values calculated by the Scherrer method [50], listed in Table 1, indicate dXRD sizes in the range of 45–55nm. The TEM images of the pristine SrTiO3, show the formation of quasi-spherical particles with a distribution of particle sizes, see Figure 2B, as commonly observed in FSP-made particles [19]. Specifically, according to TEM, the FSP-made SrTiO3 includes a few large particles with diameters of 40–50nm and many smaller particles with diameters below 20nm. The ensuing particle-size distribution calculated from the TEM images, Figure 2C, shows a mean size of dTEM = 17 ± 0.2nm as obtained from a Gaussian fitting. Comparison of dXRD and dTEM exemplifies the well-known effect of large particles to predominate the diffraction peaks in XRD, thus dXRD overestimates the average particle size.
Table 1. Structural Characteristics of La:SrTiO3, SrTiO3/CuO, and La:SrTiO3/CuO nanoparticles.
Table 1. Structural Characteristics of La:SrTiO3, SrTiO3/CuO, and La:SrTiO3/CuO nanoparticles.
Nanomaterial 1La-, Cu-Content
XRF Analysis (%wt)
dXRD (nm)
(±4)
SSA (m2 g−1)
(± 0.5)
Total Pore Volume (cm3 g−1)
(±0.02 × 10−2)
Band Gap Eg (eV) (± 0.1)
Single-Nozzle FSP
Pristine SrTiO3La:0 /Cu:04532.30.14 × 10−23.2
0.9La:STOLa:0.93 ± 0.05 /Cu:04753.10.39 × 10−23.2
0.35La:STOLa:0.36 ± 0.05 /Cu:05557.50.36 × 10−23.2
Double-Nozzle FSP
STO/2CuLa:0.05 ± 0.05 /Cu:2 ± 0.15334.90.12 × 10−23.2
STO/1.2CuLa:0/Cu:1.2 ± 0.14932.10.11 × 10−23.2
STO/0.5CuLa: 0.05 ± 0.05 /Cu:0.5 ± 0.14532.00.11 × 10−23.2
La:STO/CuLa: 0.26 ± 0.05 /Cu:0.5 ± 0.15237.30.16 × 10−23.2
1 X_La:STO/Y_Cu where STO = SrTiO3, X = the nominal % La-content per weight of SrTiO3, Y = the nominal % Cu-content per weight of SrTiO3.
In Figure 3A, HRTEM images for La:STO/Cu are shown. Distinct Miller planes of CuO (110) are resolved with d = 2.75 Å [53], showing that CuO particles are deposited on the surface of the SrTiO3 particle. These CuO particles are small, <2 nm, therefore they are not detected in the XRD. Figure 3C–F present the scanning-TEM images for the Ti, Sr, La, and Cu atoms mapping for material La:STO/Cu. Importantly, Figure 3E demonstrates that the La atoms are evenly dispersed throughout the whole volume of the SrTiO3 particle. This verifies the atomic distribution of La, i.e., without La-clusters. The STEM for Cu atoms, Figure 3D, verifies the formation of dense/particle structures on SrTiO3. Caution is drawn to the fact that the faint blue hue in Figure 3D is the enhanced emissions from Cu-grid atoms employed for the TEM measurements that are in proximity to the Sr-atoms. These are secondary Cu-electrons enhanced via Strontium excitations. Thus, in Figure 3F, only the particles inside the white circle are actual CuO particles, showing that CuO particles are found only on the surface of the SrTiO3.
La-doped SrTiO3 produced by Single-Nozzle FSP retain their high crystallinity and purity, i.e., no secondary phases are formed, such as La2O3. La-doping is evidenced by a pink-hue color developed in La:SrTiO3 vs. the white SrTiO3. Upon increasing La-doping, the SrTiO3-particle size tends to increase, i.e., from 45 nm dXRD to 55 nm dXRD. Strikingly, the Specific Surface Area increases also, see Table 1. This counterintuitive observation is further analyzed in the BET-data analysis hereafter.
Double-Nozzle FSP used to form the SrTiO3/CuO heterojunctions, retains the high crystallinity of SrTiO3, see XRD data in Figure 2A, with a tendency towards larger particle sizes at increased Cu-content. In agreement with previous reports [28], this can be attributed to the increased contribution to the increased enthalpy that is added due to the second flame, thus increasing the overall synthesis temperature, and extending the temperature at the stage of agglomeration.
The N2-adsorption isotherms for selected nanomaterials are presented in Figure 3A, while the data for nanomaterials with different CuO percentages are shown in Figure S1 in the Supporting Information. All nanomaterials have the characteristic of a type-IV isotherm. The SSA values and pore-volume analysis, see Figure 4B, reveal some interesting trends: in the Single-Nozzle FSP, the La-dopped particles show an increase in their SSA values, see Table 1. However, taking into account the XRD data, this increased SSA is not concurring with the particle size. More insightful information is obtained by examining the pore size and pore volume trends, see Figure 4B, which reveals an increase in pore volume upon La-doping. Typically, in literature, SrTiO3 particles are reported to possess a total pore volume in the range of 15 cm3 g−1 [18,54], which is in the same range observed for our pristine SrTiO3 nanomaterial, Figure S2 in the Supporting Information. Upon La-doping, a sharp increase is observed for the SSA, but more importantly a 3-fold increase in the pore volume to 39 cm3 g−1. This trend can be attributed to a geometrical effect as depicted in the scheme in Figure 4. This is a result of the FSP process, where the La-doping decreases the packing/aggregation of the SrTiO3, even though some SrTiO3 might grow bigger. It is worth mentioning the relatively high surface area and pore volume in our nanomaterials, i.e., compared vs. previous synthesis methods for SrTiO3 that possessed low SSA, i.e., due to the application of high calcination temperatures [5,55].

3.2. Spectroscopic Characterization

3.2.1. Diffuse Reflectance UV-Vis Spectroscopy

SrTiO3, as a semiconductor, has been shown to possess an indirect bandgap of 3.25 eV and a direct bandgap of 3.75 eV [56]. It has been shown that La or Nb-dopings, or oxygen vacancies can produce an n-type alteration in the Density of States (DOS) [56], i.e., extra DOS are formed within the band gap right below the Conduction Band bottom. In our data in Figure 5B, the indirect band gap has been calculated using the Kubelka–Munk method [52]. Notice that its absorption starts at 390 nm [57], resulting in all the nanomaterials having Eg values close to 3.2 eV, listed in Table 1. Lanthanum/Copper incorporation in the SrTiO3 materials fundamentally changed the color of the nanoparticles, with a purple hue observed after the lanthanum doping, while a drastic brown tint was observed for the 2% CuO heterostructure [28], which is clearly visible from extra absorption in the 400–580 nm range, Figure 5A.
Overall, the present DRS-UV-Vis analysis verifies that the electronic structure of the SrTiO3 semiconductors produced by SN-FSP and DN-FSP, as well as their trends upon La-doping and Cu-heterojunction, are in accordance with literature data.

3.2.2. X-ray Photoelectron Spectroscopy

XPS data for Sr- or O-atoms are presented in Figure 6A,C, respectively. Figure 6B presents Cu-XPS data, there is an emphasis on copper in order to confirm the oxidation state of the Cu-NPs. The material STO/2Cu, i.e., with the higher Cu-loading is exemplified, since it possessed strong Cu-XPS signals, to ascertain the Cu-oxidation state with precision. La-could not be detected by XPS in any of our materials.
In Figure 6B, the binding energies located broadly at 932.8eV and 952.7eV correspond to the Cu 2p3/2 and Cu 2p1/2, respectively [58,59], while the presence of a strong satellite signal at 941eV and the broader peaks of Cu 2p3/2 and Cu 2p1/2 indicate that the Cu atoms are in the Cu(II) oxidation state [58,59]. Thus, XPS results confirm that the heterostructure is indeed SrTiO3/CuO. This is in agreement with the FSP settings, i.e., the P/D = 5/5 produces a typical oxidizing environment [19] that in our Cu-precursor promotes the formation of CuO. The binding energies at 132.2 eV and 134 eV were identified as the Sr 3d5/2 and Sr 3d3/2, corresponding to the Sr2+ state of the SrTiO3 [60,61], Figure 6A. For the oxygen species, Figure 6C, three peaks were observed, with the binding energies at 528.9 eV that can be attributed to lattice oxygen species of the SrTiO3 crystal structure [61,62]. The 531.1 eV is attributed to chemisorbed oxygen, and lastly, the peak at 532.4 eV is attributed to adsorbed oxygen on the particle surface or hydroxyl groups [61,62]. The oxygen species from the partially substituted materials, i.e., 0.25%La/0.5%CuO have distinctly different intensity ratios of the (lattice) vs. (chemisorbed) oxygen species. This can be possibly attributed to oxygen vacancies that are created from the bending of the lattice, impacting the adsorption and efficiency of the oxygen species at the catalyst surface [61,62]. For the Ti 2p, all materials have peaks located at 457.7 eV and 463.4 eV, which are the binding states of Ti 2p3/2 and Ti 2p1/2, respectively, that correspond to typical Ti4+ states, Figure S3 in the Supporting Information.
Overall, the present XPS data show that the FSP-made nano SrTiO3 consists of typical Ti4+ states with some O vacancies in all materials. In STO/Cu, the CuO particle was confirmed by XPS in accordance with the oxidizing FSP process used herein.

3.3. Photocatalytic Evaluation

Photocatalytic results for photocatalytic hydrogen production of the materials were evaluated using a mixture of H2O/CH3OH as a catalytic substrate. The only products were H2 or CH4, in all cases. The gas-production data are shown in Figure 7A,B for H2 or CH4, respectively. The corresponding rates are presented in Figure 7C,D, respectively. The highest H2-yield was 11980 umol g−1 h−1, achieved by the 0.9%La:SrTiO3, which is ~500% higher than for pristine SrTiO3 with H2-yield 2760 umol g−1 h−1. A clear beneficial trend is observed, i.e., higher La-doping promotes H2 production. This trend is in agreement with the literature [31,32]. Herein, however, we focus attention on the relative rates for CH4 also.
From Figure 7 we observe that in the absence of CuO, the production of CH4 from the photoreduction of methanol was minimal. The SrTiO3/CuO heterostructure portrays very different results, with the 2%CuO having almost the same H2 production as the pristine SrTiO3, however, there is a sharp increase in the CH4 (76 to 136 umol g−1 h−1). This selectivity towards CH4 is apparent for all CuO heterojunctions. Decreasing CuO content resulted in higher H2 and CH4 production, 2%CuO increasing the production to 704 umol g−1 h−1 CH4. Notice that with increasing Cu-loading the SrTiO3 particle surface is fully covered with CuO, which in turn might act as an inhibitor of light-absorbance by the SrTiO3. This ‘darkening’ effect plays a key role in diminishing the photocatalytic activity, i.e., both H2 and CH4. Most interestingly, the material 0.25%La/0.5%Cu showed an enhanced H2/CH4 selectivity, with the CH4 reaching 1469 umol g−1 h−1 and H2 to 5907 umol g−1 h−1.
Overall, the present data in Figure 7 show that: (i) in all instances, La-doping greatly increases the photocatalytic activity, (ii) CuO on SrTiO3 drives the products toward the reduction of CH3OH to CH4. Importantly, after the photocatalytic application, the particles fully retained their structure as evidenced by XRD data, see Figure S4 in the Supporting Information.

4. Discussion

The present data show that FSP offers a versatile technology for the production of nano SrTiO3, La:SrTiO3, SrTiO3/CuO, and La:SrTiO3/CuO. These materials show significant photocatalytic activity that can be tailored towards either pure H2 production or CH4/H2 production from an H2O/CH3OH mixture. Table 2 allows a comparison of H2 photogeneration by the present FSP particles vs. literature, SrTiO3-based nanocatalysts. From Table 2, and the literature generally, the use of methanol or ethanol as a sacrificial agent to enhance CH4 production has not been reported so far. Most research publications utilize Xenon radiation, with the highest yields shown in Table 2, although there are several with UV irradiation.
In the case of SrTiO3, full exploitation of its band gap > 3.2 eV dictates the use of UV light, in order to maximize the H2 yield [63,64,65,66,67], in comparison with visible-light H2 production [68,69]. In this context, Table 2, shows that the FSP-made 0.9%La:SrTiO3 nanocatalyst has the highest yield vs. literature data on pertinent La:SrTiO3 materials, although the irradiation method that was used herein (UV-Lamp) definitely boosts the photocatalytic efficiency. So far, stabilizing CuO or Cu2O NPs on metal oxides with lower conduction-band positions is a common strategy to enhance photocatalytic H2 production compared with pristine CuO and Cu2O [70,71]. In this context, it was reported that coupling Cu2O with TiO2 or ZnO leads to higher H2 photoproduction [70,72]. Although there are no previous data on CH4 production promotion by SrTiO3/CuO or La:SrTiO3/CuO heterojunctions, some insights can be gathered from other systems. For example, Xiong et al. demonstrated that on a TiO2-Pt-Cu2O nanohybrid, Cu2O promotes the CH4 production and at the same time suppresses the H2 evolution, whereas Pt favors H2O activation [73]. Moreover, Chen et al. had shown that the co-existence of Cu1+/Cu0 species on TiO2 enhances the photocatalytic efficiency, by increasing the lifetime of electrons leading to an enhancement of CO2 hydrogenation. In the same study, the reduction of CuO to Cu0 was found to be more efficient for the photocatalytic production of CH4. A Cu-Cu2O/TiO2 hybrid has shown an excellent selectivity for CH4 which is attributed to the suppression of CO formation from Cu0 species while Cu1+ species act as the active sites for CH4 production [74]. All these data pose the possibility that the promotion of CH4 production by SrTiO3/CuO or La:SrTiO3/CuO heterojunctions might involve the formation of Cu1+ or Cu0 species at the SrTiO3/CuO interface.

5. Conclusions

In the present work, we introduce Double-Nozzle Flame Spray Pyrolysis technology as a synthesis method for production of nano SrTiO3 perovskites, with in tandem control of La-Doping of SrTiO3 crystal plus a CuO/SrTiO3-heterojunction. The proposed FSP technology allows controlled production of La:SrTiO3, SrTiO3/CuO, or La:SrTiO3/CuO with fundamental advantages of the one-step production, and the potential for scalable production of complex nanomaterials.
The resulting nanomaterials showed distinct structural-electronic properties, with the La-doping inducing a characteristic increase in SSA via the formation of larger pore voids. Diligent control of the La-doping and SrTiO3/CuO heterostructure allowed a selective control of photocatalytic production of H2 or CH4 from an H2O/CH3OH mixture. La-doping in all cases increased the photocatalytic activity of SrTiO3 nanocatalysts, with the 0.9La:STO showing a benchmark H2-production rate of 12 mmol g−1 h−1. The incorporation of CuO drastically shifted the selectivity from H2 toward CH4. The highest production was achieved with in tandem incorporation of La and CuO, i.e., La:SrTiO3/CuO catalyst, with a CH4 production rate of 1.5 mmol g−1 h−1.
Thus, the present work exemplifies FSP as a potent technology for the production of complex nanocatalysts, at the same time bringing new insights into photocatalysis for H2/CH4 production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13030482/s1, Figure S1: Pore distribution and N2 adsorption-desorption isotherms of SrTiO3 with 2%, 1.2%, and 0.5% CuO materials; Figure S2: XRD and N2 adsorption-desorption isotherm of the pristine SrTiO3; Figure S3: XPS data for the nanomaterials for the Ti binding energies; Figure S4: XRD data for the before and after photocatalysis.

Author Contributions

Conceptualization, Y.D.; methodology, P.P. and A.Z.; formal analysis, P.P. and A.Z.; investigation, P.P., A.Z., C.P. and N.B.; data curation, P.P. and A.Z.; writing—original draft preparation, P.P.; writing—review and editing, P.P., A.Z. and Y.D.; supervision, Y.D.; funding acquisition, Y.D.; All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Hellenic Foundation for Research and Innovation (H.F.R.I) under the “First Call for H.F.R.I Research Projects to support Faculty members and Researchers and the procurement of high-cost research equipment grant” (HFRI-FM17-1888).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Dawood, F.; Anda, M.; Shafiullah, G.M. Hydrogen Production for Energy: An Overview. Int. J. Hydrog. Energy 2020, 45, 3847–3869. [Google Scholar] [CrossRef]
  2. Nikolaidis, P.; Poullikkas, A. A Comparative Overview of Hydrogen Production Processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. [Google Scholar] [CrossRef]
  3. Wang, Q.; Domen, K. Particulate Photocatalysts for Light-Driven Water Splitting: Mechanisms, Challenges, and Design Strategies. Chem. Rev. 2020, 120, 919–985. [Google Scholar] [CrossRef] [PubMed]
  4. Phoon, B.L.; Lai, C.W.; Juan, J.C.; Show, P.; Chen, W. A Review of Synthesis and Morphology of SrTiO3 for Energy and Other Applications. Int. J. Energy Res. 2019, 43, 5151–5174. [Google Scholar] [CrossRef]
  5. Suárez-Vázquez, S.I.; Gil, S.; García-Vargas, J.M.; Cruz-López, A.; Giroir-Fendler, A. Catalytic Oxidation of Toluene by SrTi1-XBXO3 (B = Cu and Mn) with Dendritic Morphology Synthesized by One Pot Hydrothermal Route. Appl. Catal. B Environ. 2018, 223, 201–208. [Google Scholar] [CrossRef]
  6. Moos, R.; Hardtl, K.H. Defect Chemistry of Donor-Doped and Undoped Strontium Titanate Ceramics between 1000° and 1400 °C. J. Am. Ceram. Soc. 2005, 80, 2549–2562. [Google Scholar] [CrossRef]
  7. Opoku, F.; Govender, K.K.; van Sittert, C.G.C.E.; Govender, P.P. Enhancing Charge Separation and Photocatalytic Activity of Cubic SrTiO3 with Perovskite-Type Materials MTaO3 (M=Na, K) for Environmental Remediation: A First-Principles Study. ChemistrySelect 2017, 2, 6304–6316. [Google Scholar] [CrossRef]
  8. Zhang, C.; Jiang, N.; Xu, S.; Li, Z.; Liu, X.; Cheng, T.; Han, A.; Lv, H.; Sun, W.; Hou, Y. Towards High Visible Light Photocatalytic Activity in Rare Earth and N Co-Doped SrTiO3: A First Principles Evaluation and Prediction. RSC Adv. 2017, 7, 16282–16289. [Google Scholar] [CrossRef] [Green Version]
  9. Fan, Y.; Liu, Y.; Cui, H.; Wang, W.; Shang, Q.; Shi, X.; Cui, G.; Tang, B. Photocatalytic Overall Water Splitting by SrTiO3 with Surface Oxygen Vacancies. Nanomaterials 2020, 10, 2572. [Google Scholar] [CrossRef]
  10. Chen, H.-C.; Huang, C.-W.; Wu, J.C.S.; Lin, S.-T. Theoretical Investigation of the Metal-Doped SrTiO3 Photocatalysts for Water Splitting. J. Phys. Chem. C 2012, 116, 7897–7903. [Google Scholar] [CrossRef]
  11. Ma, Y.; Wu, Z.; Wang, H.; Wang, G.; Zhang, Y.; Hu, P.; Li, Y.; Gao, D.; Pu, H.; Wang, B.; et al. Synthesis of Nanocrystalline Strontium Titanate by a Sol–Gel Assisted Solid Phase Method and Its Formation Mechanism and Photocatalytic Activity. CrystEngComm 2019, 21, 3982–3992. [Google Scholar] [CrossRef]
  12. Kakihana, M.; Okubo, T.; Arima, M.; Nakamura, Y.; Yashima, M.; Yoshimura, M. Polymerized Complex Route to the Synthesis of Pure SrTiO3 at Reduced Temperatures: Implication for Formation of Sr-Ti Heterometallic Citric Acid Complex. J. Sol-Gel Sci. Technol. 1998, 12, 95–109. [Google Scholar] [CrossRef]
  13. Kubacka, A.; Fernández-García, M.; Colón, G. Advanced Nanoarchitectures for Solar Photocatalytic Applications. Chem. Rev. 2012, 112, 1555–1614. [Google Scholar] [CrossRef]
  14. Madler, L.; Kammler, H.K.; Mueller, R.; Pratsinis, S.E. Controlled Synthesis of Nanostructured Particles by Ame Spray Pyrolysis. Aerosol. Sci. 2002, 33, 369–389. [Google Scholar] [CrossRef]
  15. Mädler, L.; Stark, W.J.; Pratsinis, S.E. Flame-Made Ceria Nanoparticles. J. Mater. Res. 2002, 17, 1356–1362. [Google Scholar] [CrossRef] [Green Version]
  16. Mueller, R.; Mädler, L.; Pratsinis, S.E. Nanoparticle Synthesis at High Production Rates by Flame Spray Pyrolysis. Chem. Eng. Sci. 2003, 58, 1969–1976. [Google Scholar] [CrossRef]
  17. Shih, S.-J.; Tzeng, W.-L. Manipulation of Morphology of Strontium Titanate Particles by Spray Pyrolysis. Powder Technol. 2014, 264, 291–297. [Google Scholar] [CrossRef]
  18. Kang, H.W.; Park, S.B. Doping of Fluorine into SrTiO3 by Spray Pyrolysis for H2 Evolution under Visible Light Irradiation. Chem. Eng. Sci. 2013, 100, 384–391. [Google Scholar] [CrossRef]
  19. Teoh, W.Y.; Amal, R.; Mädler, L. Flame Spray Pyrolysis: An Enabling Technology for Nanoparticles Design and Fabrication. Nanoscale 2010, 2, 1324. [Google Scholar] [CrossRef]
  20. Gröhn, A.J.; Pratsinis, S.E.; Sánchez-Ferrer, A.; Mezzenga, R.; Wegner, K. Scale-up of Nanoparticle Synthesis by Flame Spray Pyrolysis: The High-Temperature Particle Residence Time. Ind. Eng. Chem. Res. 2014, 53, 10734–10742. [Google Scholar] [CrossRef]
  21. Moularas, C.; Psathas, P.; Deligiannakis, Y. Electron Paramagnetic Resonance Study of Photo-Induced Hole/Electron Pairs in NaTaO3 Nanoparticles. Chem. Phys. Lett. 2021, 782, 139031. [Google Scholar] [CrossRef]
  22. Psathas, P.; Solakidou, M.; Mantzanis, A.; Deligiannakis, Y. Flame Spray Pyrolysis Engineering of Nanosized Mullite-Bi2Fe4O9 and Perovskite-BiFeO3 as Highly Efficient Photocatalysts for O2 Production from H2O Splitting. Energies 2021, 14, 5235. [Google Scholar] [CrossRef]
  23. Psathas, P.; Georgiou, Y.; Moularas, C.; Armatas, G.S.; Deligiannakis, Y. Controlled-Phase Synthesis of Bi2Fe4O9 & BiFeO3 by Flame Spray Pyrolysis and Their Evaluation as Non-Noble Metal Catalysts for Efficient Reduction of 4-Nitrophenol. Powder Technol. 2020, 368, 268–277. [Google Scholar] [CrossRef]
  24. Stathi, P.; Solakidou, M.; Deligiannakis, Y. Lattice Defects Engineering in W-, Zr-Doped BiVO4 by Flame Spray Pyrolysis: Enhancing Photocatalytic O2 Evolution. Nanomaterials 2021, 11, 501. [Google Scholar] [CrossRef] [PubMed]
  25. Kho, Y.K.; Teoh, W.Y.; Iwase, A.; Mädler, L.; Kudo, A.; Amal, R. Flame Preparation of Visible-Light-Responsive BiVO4 Oxygen Evolution Photocatalysts with Subsequent Activation via Aqueous Route. ACS Appl. Mater. Interfaces 2011, 3, 1997–2004. [Google Scholar] [CrossRef] [PubMed]
  26. Abe, Y.; Laine, R.M. Photocatalytic Plate-like La2Ti2O7 Nanoparticles Synthesized via Liquid-feed Flame Spray Pyrolysis (LF-FSP) of Metallo-organic Precursors. J. Am. Ceram. Soc. 2020, 103, 4832–4839. [Google Scholar] [CrossRef]
  27. Yuan, X.; Meng, L.; Zheng, C.; Zhao, H. Deep Insight into the Mechanism of Catalytic Combustion of CO and CH4 over SrTi1– xBxO3 (B = Co, Fe, Mn, Ni, and Cu) Perovskite via Flame Spray Pyrolysis. ACS Appl. Mater. Interfaces 2021, 13, 52571–52587. [Google Scholar] [CrossRef]
  28. Yuan, X.; Meng, L.; Xu, Z.; Zheng, C.; Zhao, H. CuO Quantum Dots Supported by SrTiO3 Perovskite Using the Flame Spray Pyrolysis Method: Enhanced Activity and Excellent Thermal Resistance for Catalytic Combustion of CO and CH4. Environ. Sci. Technol. 2021, 55, 14080–14086. [Google Scholar] [CrossRef]
  29. Ahmed, A.J.; Hossain, M.S.A.; Kazi Nazrul Islam, S.M.; Yun, F.; Yang, G.; Hossain, R.; Khan, A.; Na, J.; Eguchi, M.; Yamauchi, Y.; et al. Significant Improvement in Electrical Conductivity and Figure of Merit of Nanoarchitectured Porous SrTiO3 by La Doping Optimization. ACS Appl. Mater. Interfaces 2020, 12, 28057–28064. [Google Scholar] [CrossRef]
  30. Lu, Z.; Zhang, H.; Lei, W.; Sinclair, D.C.; Reaney, I.M. High-Figure-of-Merit Thermoelectric La-Doped A-Site-Deficient SrTiO3 Ceramics. Chem. Mater. 2016, 28, 925–935. [Google Scholar] [CrossRef]
  31. Jia, Y.; Shen, S.; Wang, D.; Wang, X.; Shi, J.; Zhang, F.; Han, H.; Li, C. Composite Sr2TiO4/SrTiO3(La,Cr) Heterojunction Based Photocatalyst for Hydrogen Production under Visible Light Irradiation. J. Mater. Chem. A 2013, 1, 7905. [Google Scholar] [CrossRef]
  32. Abdi, M.; Mahdikhah, V.; Sheibani, S. Visible Light Photocatalytic Performance of La-Fe Co-Doped SrTiO3 Perovskite Powder. Opt. Mater. 2020, 102, 109803. [Google Scholar] [CrossRef]
  33. Wang, Q.; Warnan, J.; Rodríguez-Jiménez, S.; Leung, J.J.; Kalathil, S.; Andrei, V.; Domen, K.; Reisner, E. Molecularly Engineered Photocatalyst Sheet for Scalable Solar Formate Production from Carbon Dioxide and Water. Nat. Energy 2020, 5, 703–710. [Google Scholar] [CrossRef]
  34. Moss, B.; Wang, Q.; Butler, K.T.; Grau-Crespo, R.; Selim, S.; Regoutz, A.; Hisatomi, T.; Godin, R.; Payne, D.J.; Kafizas, A.; et al. Linking in Situ Charge Accumulation to Electronic Structure in Doped SrTiO3 Reveals Design Principles for Hydrogen-Evolving Photocatalysts. Nat. Mater. 2021, 20, 511–517. [Google Scholar] [CrossRef]
  35. Wang, Q.; Hisatomi, T.; Suzuki, Y.; Pan, Z.; Seo, J.; Katayama, M.; Minegishi, T.; Nishiyama, H.; Takata, T.; Seki, K.; et al. Particulate Photocatalyst Sheets Based on Carbon Conductor Layer for Efficient Z-Scheme Pure-Water Splitting at Ambient Pressure. J. Am. Chem. Soc. 2017, 139, 1675–1683. [Google Scholar] [CrossRef]
  36. Iwashina, K.; Kudo, A. Rh-Doped SrTiO3 Photocatalyst Electrode Showing Cathodic Photocurrent for Water Splitting under Visible-Light Irradiation. J. Am. Chem. Soc. 2011, 133, 13272–13275. [Google Scholar] [CrossRef]
  37. Tse, J.; Aziz, A.; Flitcroft, J.M.; Skelton, J.M.; Gillie, L.J.; Parker, S.C.; Cooke, D.J.; Molinari, M. Unraveling the Impact of Graphene Addition to Thermoelectric SrTiO3 and La-Doped SrTiO3 Materials: A Density Functional Theory Study. ACS Appl. Mater. Interfaces 2021, 13, 41303–41314. [Google Scholar] [CrossRef]
  38. Wang, B.; Shen, S.; Guo, L. SrTiO3 single crystals enclosed with high-indexed {0 2 3} facets and {0 0 1} facets for photocatalytic hydrogen and oxygen evolution. Appl. Catal. B Environ. 2015, 166, 320–326. [Google Scholar] [CrossRef]
  39. Niishiro, R.; Tanaka, S.; Kudo, A. Hydrothermal-Synthesized SrTiO3 Photocatalyst Codoped with Rhodium and Antimony with Visible-Light Response for Sacrificial H2 and O2 Evolution and Application to Overall Water Splitting. Appl. Catal. B Environ. 2014, 150–151, 187–196. [Google Scholar] [CrossRef]
  40. Ali, S.; Razzaq, A.; Kim, H.; In, S.-I. Activity, Selectivity, and Stability of Earth-Abundant CuO/Cu2O/Cu0-Based Photocatalysts toward CO2 Reduction. Chem. Eng. J. 2022, 429, 131579. [Google Scholar] [CrossRef]
  41. Wang, Y.; Chen, E.; Tang, J. Insight on Reaction Pathways of Photocatalytic CO2 Conversion. ACS Catal. 2022, 12, 7300–7316. [Google Scholar] [CrossRef]
  42. Chen, W.-T.; Jovic, V.; Sun-Waterhouse, D.; Idriss, H.; Waterhouse, G.I.N. The Role of CuO in Promoting Photocatalytic Hydrogen Production over TiO2. Int. J. Hydrog. Energy 2013, 38, 15036–15048. [Google Scholar] [CrossRef]
  43. Yu, J.; Hai, Y.; Jaroniec, M. Photocatalytic Hydrogen Production over CuO-Modified Titania. J. Colloid Interface Sci. 2011, 357, 223–228. [Google Scholar] [CrossRef] [PubMed]
  44. Lalitha, K.; Sadanandam, G.; Kumari, V.D.; Subrahmanyam, M.; Sreedhar, B.; Hebalkar, N.Y. Highly Stabilized and Finely Dispersed Cu2O/TiO2: A Promising Visible Sensitive Photocatalyst for Continuous Production of Hydrogen from Glycerol:Water Mixtures. J. Phys. Chem. C 2010, 114, 22181–22189. [Google Scholar] [CrossRef]
  45. Chen, J.-L.; Liu, M.-M.; Xie, S.-Y.; Yue, L.-J.; Gong, F.-L.; Chai, K.-M.; Zhang, Y.-H. Cu2O-Loaded TiO2 Heterojunction Composites for Enhanced Photocatalytic H2 Production. J. Mol. Struct. 2022, 1247, 131294. [Google Scholar] [CrossRef]
  46. Strobel, R.; Mädler, L.; Piacentini, M.; Maciejewski, M.; Baiker, A.; Pratsinis, S.E. Two-Nozzle Flame Synthesis of Pt/Ba/Al2O3 for NOx Storage. Chem. Mater. 2006, 18, 2532–2537. [Google Scholar] [CrossRef]
  47. Lovell, E.C.; Großman, H.; Horlyck, J.; Scott, J.; Mädler, L.; Amal, R. Asymmetrical Double Flame Spray Pyrolysis-Designed SiO2/Ce0.7Zr0.3O2 for the Dry Reforming of Methane. ACS Appl. Mater. Interfaces 2019, 11, 25766–25777. [Google Scholar] [CrossRef]
  48. Solakidou, M.; Georgiou, Y.; Deligiannakis, Y. Double-Nozzle Flame Spray Pyrolysis as a Potent Technology to Engineer Noble Metal-TiO2 Nanophotocatalysts for Efficient H2 Production. Energies 2021, 14, 817. [Google Scholar] [CrossRef]
  49. Kumaravel, V.; Imam, M.; Badreldin, A.; Chava, R.; Do, J.; Kang, M.; Abdel-Wahab, A. Photocatalytic Hydrogen Production: Role of Sacrificial Reagents on the Activity of Oxide, Carbon, and Sulfide Catalysts. Catalysts 2019, 9, 276. [Google Scholar] [CrossRef] [Green Version]
  50. Patterson, A.L. The Scherrer Formula for X-Ray Particle Size Determination. Phys. Rev. 1939, 56, 978–982. [Google Scholar] [CrossRef]
  51. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of Gases in Multimolecular Layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  52. Tauc, J.; Grigorovici, R.; Vancu, A. Optical Properties and Electronic Structure of Amorphous Germanium. Phys. Status Solidi 1966, 15, 627–637. [Google Scholar] [CrossRef]
  53. Khan, M.A.; Nayan, N.; Shadiullah; Ahmad, M.K.; Soon, C.F. Surface Study of CuO Nanopetals by Advanced Nanocharacterization Techniques with Enhanced Optical and Catalytic Properties. Nanomaterials 2020, 10, 1298. [Google Scholar] [CrossRef]
  54. Baba-Ahmed, I.; Ghercă, D.; Iordan, A.-R.; Palamaru, M.N.; Mita, C.; Baghdad, R.; Ababei, G.; Lupu, N.; Benamar, M.A.; Abderrahmane, A.; et al. Sequential Synthesis Methodology Yielding Well-Defined Porous 75%SrTiO3/25%NiFe2O4 Nanocomposite. Nanomaterials 2021, 12, 138. [Google Scholar] [CrossRef] [PubMed]
  55. López-Suárez, F.E.; Parres-Esclapez, S.; Bueno-López, A.; Illán-Gómez, M.J.; Ura, B.; Trawczynski, J. Role of Surface and Lattice Copper Species in Copper-Containing (Mg/Sr)TiO3 Perovskite Catalysts for Soot Combustion. Appl. Catal. B Environ. 2009, 93, 82–89. [Google Scholar] [CrossRef]
  56. van Benthem, K.; Elsässer, C.; French, R.H. Bulk Electronic Structure of SrTiO3: Experiment and Theory. J. Appl. Phys. 2001, 90, 6156–6164. [Google Scholar] [CrossRef] [Green Version]
  57. Maček Kržmanc, M.; Daneu, N.; Čontala, A.; Santra, S.; Kamal, K.M.; Likozar, B.; Spreitzer, M. SrTiO3/Bi4Ti3O12 nanoheterostructural platelets synthesized by topotactic epitaxy as effective noble-metal-free photocatalysts for pH-neutral hydrogen evolution. ACS Appl. Mater. Interfaces 2020, 13, 370–381. [Google Scholar] [CrossRef]
  58. Mai, X.T.; Bui, D.N.; Pham, V.K.; Nguyen, T.H.L.; Nguyen, T.T.L.; Chau, H.D.; Tran, T.K.N. Effect of CuO Loading on the Photocatalytic Activity of SrTiO3 for Hydrogen Evolution. Inorganics 2022, 10, 130. [Google Scholar] [CrossRef]
  59. Xia, Y.; He, Z.; Hu, K.; Tang, B.; Su, J.; Liu, Y.; Li, X. Fabrication of N-SrTiO3/p-Cu2O Heterojunction Composites with Enhanced Photocatalytic Performance. J. Alloy. Compd. 2018, 753, 356–363. [Google Scholar] [CrossRef]
  60. Li, W.; Liu, S.; Wang, S.; Guo, Q.; Guo, J. The Roles of Reduced Ti Cations and Oxygen Vacancies in Water Adsorption and Dissociation on SrTiO3 (110). J. Phys. Chem. C 2014, 118, 2469–2474. [Google Scholar] [CrossRef]
  61. Ura, B.; Trawczyński, J.; Kotarba, A.; Bieniasz, W.; Illán-Gómez, M.J.; Bueno-López, A.; López-Suárez, F.E. Effect of Potassium Addition on Catalytic Activity of SrTiO3 Catalyst for Diesel Soot Combustion. Appl. Catal. B Environ. 2011, 101, 169–175. [Google Scholar] [CrossRef]
  62. Tan, H.; Zhao, Z.; Zhu, W.; Coker, E.N.; Li, B.; Zheng, M.; Yu, W.; Fan, H.; Sun, Z. Oxygen Vacancy Enhanced Photocatalytic Activity of Pervoskite SrTiO3. ACS Appl. Mater. Interfaces 2014, 6, 19184–19190. [Google Scholar] [CrossRef] [PubMed]
  63. Ramos-Sanchez, J.E.; Camposeco, R.; Lee, S.-W.; Rodríguez-González, V. Sustainable Synthesis of AgNPs/Strontium-Titanate-Perovskite-like Catalysts for the Photocatalytic Production of Hydrogen. Catal. Today 2020, 341, 112–119. [Google Scholar] [CrossRef]
  64. Liu, Y.; Xie, L.; Li, Y.; Yang, R.; Qu, J.; Li, Y.; Li, X. Synthesis and High Photocatalytic Hydrogen Production of SrTiO3 Nanoparticles from Water Splitting under UV Irradiation. J. Power Sources 2008, 183, 701–707. [Google Scholar] [CrossRef]
  65. Macaraig, L.; Chuangchote, S.; Sagawa, T. Electrospun SrTiO3 Nanofibers for Photocatalytic Hydrogen Generation. J. Mater. Res. 2014, 29, 123–130. [Google Scholar] [CrossRef] [Green Version]
  66. Qin, Y.; Wang, G.; Wang, Y. Study on the Photocatalytic Property of La-Doped CoO/SrTiO3 for Water Decomposition to Hydrogen. Catal. Commun. 2007, 8, 926–930. [Google Scholar] [CrossRef]
  67. Sakata, Y.; Miyoshi, Y.; Maeda, T.; Ishikiriyama, K.; Yamazaki, Y.; Imamura, H.; Ham, Y.; Hisatomi, T.; Kubota, J.; Yamakata, A.; et al. Photocatalytic Property of Metal Ion Added SrTiO3 to Overall H2O Splitting. Appl. Catal. A Gen. 2016, 521, 227–232. [Google Scholar] [CrossRef]
  68. Yu, K.; Zhang, C.; Chang, Y.; Feng, Y.; Yang, Z.; Yang, T.; Lou, L.-L.; Liu, S. Novel Three-Dimensionally Ordered Macroporous SrTiO3 Photocatalysts with Remarkably Enhanced Hydrogen Production Performance. Appl. Catal. B Environ. 2017, 200, 514–520. [Google Scholar] [CrossRef]
  69. Kumar, A.; Navakoteswara Rao, V.; Kumar, A.; Mushtaq, A.; Sharma, L.; Halder, A.; Pal, S.K.; Shankar, M.V.; Krishnan, V. Three-Dimensional Carbonaceous Aerogels Embedded with Rh-SrTiO3 for Enhanced Hydrogen Evolution Triggered by Efficient Charge Transfer and Light Absorption. ACS Appl. Energy Mater. 2020, 3, 12134–12147. [Google Scholar] [CrossRef]
  70. Lv, S.; Wang, Y.; Zhou, Y.; Liu, Q.; Song, C.; Wang, D. Oxygen Vacancy Stimulated Direct Z-Scheme of Mesoporous Cu2O/TiO2 for Enhanced Photocatalytic Hydrogen Production from Water and Seawater. J. Alloy. Compd. 2021, 868, 159144. [Google Scholar] [CrossRef]
  71. Lou, Y.; Zhang, Y.; Cheng, L.; Chen, J.; Zhao, Y. A Stable Plasmonic Cu@Cu2O/ZnO Heterojunction for Enhanced Photocatalytic Hydrogen Generation. ChemSusChem 2018, 11, 1505–1511. [Google Scholar] [CrossRef] [PubMed]
  72. Yoo, H.; Kahng, S.; Hyeun Kim, J. Z-Scheme Assisted ZnO/Cu2O-CuO Photocatalysts to Increase Photoactive Electrons in Hydrogen Evolution by Water Splitting. Sol. Energy Mater. Sol. Cells 2020, 204, 110211. [Google Scholar] [CrossRef]
  73. Xiong, Z.; Lei, Z.; Kuang, C.-C.; Chen, X.; Gong, B.; Zhao, Y.; Zhang, J.; Zheng, C.; Wu, J.C.S. Selective Photocatalytic Reduction of CO2 into CH4 over Pt-Cu2O TiO2 Nanocrystals: The Interaction between Pt and Cu2O Cocatalysts. Appl. Catal. B Environ. 2017, 202, 695–703. [Google Scholar] [CrossRef]
  74. Chen, B.-R.; Nguyen, V.-H.; Wu, J.C.S.; Martin, R.; Kočí, K. Production of Renewable Fuels by the Photohydrogenation of CO2: Effect of the Cu Species Loaded onto TiO2 Photocatalysts. Phys. Chem. Chem. Phys. 2016, 18, 4942–4951. [Google Scholar] [CrossRef]
Figure 2. (A) XRD patterns for the SrTiO3 nanomaterials. Inset: The unit-cell structure of the cubic SrTiO3 structure. (B) TEM image of the pristine SrTiO3 particle. (C) Size distribution graph obtained from several TEM images.
Figure 2. (A) XRD patterns for the SrTiO3 nanomaterials. Inset: The unit-cell structure of the cubic SrTiO3 structure. (B) TEM image of the pristine SrTiO3 particle. (C) Size distribution graph obtained from several TEM images.
Nanomaterials 13 00482 g002
Figure 3. (A,B) HRTEM images for the particle La:STO/Cu, with the distinct miller planes of CuO (110) deposited on the SrTiO3 surface. (CF) Scanning-TEM images with Ti, Sr, La, and Cu atom mapping for material La:STO/Cu.
Figure 3. (A,B) HRTEM images for the particle La:STO/Cu, with the distinct miller planes of CuO (110) deposited on the SrTiO3 surface. (CF) Scanning-TEM images with Ti, Sr, La, and Cu atom mapping for material La:STO/Cu.
Nanomaterials 13 00482 g003
Figure 4. (A) Nitrogen adsorption isotherms for the SrTiO3-based nanoparticles. (B) The corresponding pore-size distribution plot using the BJH method and the Cumulative Pore Volume. Bottom: Schematic visualization of the La-doping effect on the nanomaterial aggregation.
Figure 4. (A) Nitrogen adsorption isotherms for the SrTiO3-based nanoparticles. (B) The corresponding pore-size distribution plot using the BJH method and the Cumulative Pore Volume. Bottom: Schematic visualization of the La-doping effect on the nanomaterial aggregation.
Nanomaterials 13 00482 g004
Figure 5. (A) DRS-UV-Vis data for the SrTiO3 particles. (B) Tauc-Plots with the dotted arrows show the calculated bandgap. Side-photos: The powder nanoparticles showcase the color shift.
Figure 5. (A) DRS-UV-Vis data for the SrTiO3 particles. (B) Tauc-Plots with the dotted arrows show the calculated bandgap. Side-photos: The powder nanoparticles showcase the color shift.
Nanomaterials 13 00482 g005
Figure 6. (A) Sr 3d5/2 and Sr 3d3/2 XPS spectra for 0.9La:STO, STO/2Cu, and La:STO/Cu materials. (B) Cu 2p3/2 and Cu 2p1/2 XPS spectra of STO/2Cu for the binding energies. (C) Oxygen XPS spectra of 0.9La:STO, STO/2Cu, and La:STO/Cu.
Figure 6. (A) Sr 3d5/2 and Sr 3d3/2 XPS spectra for 0.9La:STO, STO/2Cu, and La:STO/Cu materials. (B) Cu 2p3/2 and Cu 2p1/2 XPS spectra of STO/2Cu for the binding energies. (C) Oxygen XPS spectra of 0.9La:STO, STO/2Cu, and La:STO/Cu.
Nanomaterials 13 00482 g006
Figure 7. Photocatalytic production by the nanomaterials from an H2O/CH3OH mixture. (A) H2-gas vs. time. (B) CH4-gas vs. time. (C,D) the corresponding gas production rates (umol per gr catalyst per hour).
Figure 7. Photocatalytic production by the nanomaterials from an H2O/CH3OH mixture. (A) H2-gas vs. time. (B) CH4-gas vs. time. (C,D) the corresponding gas production rates (umol per gr catalyst per hour).
Nanomaterials 13 00482 g007
Table 2. Hydrogen production rates and conditions reported for similar particles/methods.
Table 2. Hydrogen production rates and conditions reported for similar particles/methods.
Photo
Catalyst
Synthesis MethodCocatalyst
(wt%)
Light SourceReaction
Solution
H2 Yield (umol g−1 h−1)Ref.
0.9%La: SrTiO3 FSP1% PtHg (250 W)H2O + 20% CH3OH11,978This work
0.25%La:SrTiO/0.5% CuOFSP1% PtHg (250 W)H2O + 20% CH3OH5907This work
Ag-STOMicrowave-assisted hydrothermal 2% AgUV-low pressure Hg 50% H2O + 50% ethanol463[63]
Pt- SrTiO3Polymerized-Complex0.32% PtHg (500 W)Pure water + 40% methanol3200[64]
Nanofibers SrTiO3Electrospinning method-Hg (450 W)Pure water + 40% methanol160[65]
La: CoO/SrTiO3Solid statereaction-Hg (400 W)Pure water + Na2CO32800[66]
Na:SrTiO3polymerizable complex0.3% RhHg (450 W)Pure Water1600[67]
Macroporous SrTiO3emulsion polymerization method0.7% PtXe (300 W)Pure water + 25% methanol3599[68]
g-C3N4/Rh-SrTiO3/RGOMultiple methods0.58% RhXe (260 W)Pure water + 10% TEOA3467[69]
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

Psathas, P.; Zindrou, A.; Papachristodoulou, C.; Boukos, N.; Deligiannakis, Y. In Tandem Control of La-Doping and CuO-Heterojunction on SrTiO3 Perovskite by Double-Nozzle Flame Spray Pyrolysis: Selective H2 vs. CH4 Photocatalytic Production from H2O/CH3OH. Nanomaterials 2023, 13, 482. https://doi.org/10.3390/nano13030482

AMA Style

Psathas P, Zindrou A, Papachristodoulou C, Boukos N, Deligiannakis Y. In Tandem Control of La-Doping and CuO-Heterojunction on SrTiO3 Perovskite by Double-Nozzle Flame Spray Pyrolysis: Selective H2 vs. CH4 Photocatalytic Production from H2O/CH3OH. Nanomaterials. 2023; 13(3):482. https://doi.org/10.3390/nano13030482

Chicago/Turabian Style

Psathas, Pavlos, Areti Zindrou, Christina Papachristodoulou, Nikos Boukos, and Yiannis Deligiannakis. 2023. "In Tandem Control of La-Doping and CuO-Heterojunction on SrTiO3 Perovskite by Double-Nozzle Flame Spray Pyrolysis: Selective H2 vs. CH4 Photocatalytic Production from H2O/CH3OH" Nanomaterials 13, no. 3: 482. https://doi.org/10.3390/nano13030482

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