Next Article in Journal
The Mechanical Properties of Early Aged Shotcrete under Internal Sulfate Attack
Next Article in Special Issue
Properties of Metallic and Oxide Thin Films Based on Ti and Co Prepared by Magnetron Sputtering from Sintered Targets with Different Co-Content
Previous Article in Journal
Large Deformation and Energy Absorption Behaviour of Perforated Hollow Sphere Structures under Quasi-Static Compression
Previous Article in Special Issue
Thermoelectric Properties of Cu2Se Synthesized by Hydrothermal Method and Densified by SPS Technique
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

TiO2@Cu2O n-n Type Heterostructures for Photochemistry

by
Anita Trenczek-Zajac
1,*,
Joanna Banas-Gac
2 and
Marta Radecka
1
1
Department of Inorganic Chemistry, Faculty of Materials Science and Ceramics, AGH University of Science and Technology, al. A. Mickiewicza 30, 30-059 Krakow, Poland
2
Institute of Electronics, Faculty of Computer Science, Electronics and Telecommunications, AGH University of Science and Technology, al. A. Mickiewicza 30, 30-059 Krakow, Poland
*
Author to whom correspondence should be addressed.
Materials 2021, 14(13), 3725; https://doi.org/10.3390/ma14133725
Submission received: 2 June 2021 / Revised: 19 June 2021 / Accepted: 29 June 2021 / Published: 2 July 2021
(This article belongs to the Special Issue The 44th IMAPS Poland Conference)

Abstract

:
A TiO2@Cu2O semiconductor heterostructure with better photochemical response compared to TiO2 was obtained using an electrochemical deposition method of Cu2O on the surface of TiO2 nanotubes. The choice of 1D nanotubes was motivated by the possibility of achieving fast charge transfer, which is considered best suited for photochemical applications. The morphology and structural properties of the obtained heterojunction were determined using standard methods —SEM and Raman spectroscopy. Analysis of photoelectrochemical properties showed that TiO2@Cu2O heterostructures exhibit better properties resulting from an interaction with sunlight than TiO2. A close relationship between the morphology of the heterostructures and their photoproperties was also demonstrated. Investigations representing a combination of photoelectrochemical cells for hydrogen production and photocatalysis—photoelectrocatalysis—were also carried out and confirmed the observations on the photoproperties of heterostructures. Analysis of the Mott–Schottky plots as well as photoelectrochemical measurements (Iph-V, Iph-t) showed that TiO2 as well as, unusually, Cu2O exhibit n-type conductivity. On this basis, a new energy diagram of the TiO2@Cu2O system was proposed. It was found that TiO2@Cu2O n-n type heterostructure prevents the processes of photocorrosion of copper(I) oxide contained in a TiO2-based heterostructure.

Graphical Abstract

1. Introduction

The illumination of the semiconductor with light characterized by hν > Eg is related to the creation of electrons (e) in the conduction band and electron holes (h) in the valence band. After the creation of the carriers, the separation of the opposite charges in an electric field should occur. The charge carriers can be used indirectly to drive a chemical reaction (photocatalysis or photoelectrolysis) [1,2,3,4]. The difference between photoelectrolysis and photocatalysis is the subsequent mode of action of the photogenerated electron–hole pair, which can take part in the redox reaction in two basic configurations: the particulate system and the photoelectrochemical cell PEC [3]. In a photocatalytic process, the semiconductor powders are suspended in a solution and the photogenerated holes and electrons react with chemical species such as H2O, OH, or O2 to produce hydroxyl radicals (OH), superoxide radical anions (O2•−), and H2O2, which contribute to the decomposition of adsorbed molecules at the semiconductor surface. The photoelectrolysis is conducted in PEC. For example, an n-type semiconductor acts as a photoanode and a second electrode is metallic (when n-type semiconductor photoanode is used). In photoelectrolysis of water, holes oxidize water, which results in oxygen generation at the photocatalyst surface. The electrons are transported over an external circuit to the cathode, where hydrogen is generated via the reduction of water. The photoelectrocatalysis process, which combines both electrolysis and photocatalysis, is an excellent example of the ability to delay the recombination of electron-hole pairs, increasing the lifetime of the photocarriers. The photogenerated holes act as strong oxidizing species, whereas e play a role of potential reductor. The photogenerated holes act as strong oxidizing species, whereas e play a role of potential reductor [5]. The last two decades have seen a significant increase in the number of publications on the treatment of organic pollutants in wastewater [6,7,8,9]. Titanium dioxide (TiO2) is the preferred material for photoanodes in photoelectrocatalytic applications [3,6,8,9,10], although WO3, ZnO, and other materials are also proposed [3,6,8,9]. Titanium dioxide, TiO2, has been the first material applied as a photoanode in the PEC for photoelectrolysis of water into hydrogen and oxygen [11]. TiO2 is one of the most suitable candidates for photoanodes due to its high resistance to corrosion, stability, and negative flat band potential. However, the band-gap of TiO2 is in the order of 3 eV and, as a consequence, the absorption of sunlight is hindered. Several methods that could improve the photoresponse of TiO2 have been considered [2,3,10,12,13,14,15]. The strategies are focused on improving the performance of photocatalysts such by doping, co-doping, band-gap engineering, co-catalyst decoration, heterostructures junction formation, or modification of the microstructure and morphology. Among these propositions, the use of nanomaterials (0D, 1D, 2D, 3D) and heterostructures of metal oxide semiconductors based on a heterojunction with different relative positions of the edges of the conduction and valence bands deserves special attention. From the point of view of achieving fast charge transfer, 1D nanorods and nanotubes are considered the best suited for this purpose. Nanotubes present a natural means for the fast channeling of charge carriers [10,16,17]. In the literature, three classes of heterojunctions depending on the direction of charge transfer are reported [15,18]. For Type I, the conduction band minimum of one semiconductor (CB_1) is below CB_2 of the other one, and its valence band maximum (VB_1) is above VB_2. Type II occurs when both the conduction band minimum CB_1 and valence band maximum VB_1 of one semiconductor are below those of the second one. Type III is similar to type II, but with much bigger energetic separation between the band edges of the two semiconductors. In the case of the photoelectrocatalytic process, only the type of configuration where electrons can be transferred via an external circuit from the photoanode to the cathode, is suitable.
The efficiencies η of TiO2 in various forms (i.e., nanowires, nanotubes, thin layers) and heterostructures composed of TiO2 and different narrow and wide band-gap semiconductors are presented in Table 1. The coefficient η was calculated at the biased potential VB = 0 and 0.5 V based on the photocurrent vs. potential characteristics. The higher the potential difference is, the higher efficiency is reached. However, it should be emphasized that the “green” approach expects high efficiency of PEC with zero external potential difference. Among the TiO2 photoanodes, the highest efficiency equal to 1.9 was achieved at 0.5 V for a thin layer of TiO2. Relatively high values of η are also attributed to the TiO2@MoS2 system, but their interpretation is difficult due to the lack of information about VB at which they were obtained. In the case of TiO2 nanotubes modified with tin dioxide, the range of obtained efficiencies is wide (0.21–2.12%), which results from differences in the morphology and chemical properties of these materials. The modification of TiO2 nanowires with copper(I) oxide also results in an increase in η; however, the values are not too high.
Efficiency η (solar-to-chemical energy conversion efficiency) is expressed as follows:
η = I ph   V r     V B P
where Iph—photocurrent density in the circuit while the electrode is illuminating m A c m 2 , Vr—redox reaction potential (Vr = 1.23 V vs. NHE), VB—biased potential (V), and P—light power density m W c m 2 .
Copper(I) oxide (Cu2O) is a promising narrow band-gap semiconductor for photoelectrochemical applications due to its relative position of energy bands and band-gap energy of 2.0–2.5 eV [26,27,28,29]. However, the poor stability and the fast electron–hole recombination are a serious limitation for the application of Cu2O in photoelectrocatalysis. On the other hand, there are several approaches that can be used to stabilize and improve the efficiency of Cu2O as a photoelectrode. One of them is utilizing a suitable n-type semiconductor to combine with Cu2O. The formation of a junction between two semiconductors is favorable for inhibiting the fast recombination of photocharges. In addition, Cu2O joined with a wide band-gap semiconductor prevents the narrow band-gap semiconductor from photocorrosion.
In this work, we have studied titanium dioxide nanotubes (TiO2-NT) modified by Cu2O with the ultimate aim of determining band-gap alignment of TiO2-NT@Cu2O heterojunction. The aim of this research is to demonstrate the effect of TiO2-NT@Cu2O heterostructures, in particular, on the performance of the process of photoelectrolysis and photoelectrocatalysis in PEC cells for hydrogen generation and the decomposition of organics dyes.

2. Materials and Methods

2.1. Materials

Acetone (analytically pure), isopropanol (analytically pure), glycerol (analytically pure), Na2SO4 (analytically pure), CuSO4 (analytically pure), NaOH (pure), and methylene blue were purchased from Avantor Performance Materials (Gliwice, Poland). Lactic acid (88%, analytically pure) was acquired from Chempur (Piekary Slaskie, Poland). NH4F (≥98.0%, ACS reagent), Ti foil (0.127 mm, 99.7%), tert-butanol (≥99.0%, ACS reagent), ethylenediaminetetraacetic acid disodium salt (99.0–101.0%, ACS reagent) and p-benzoquinone (≥99.5%, HPLC) were purchased from Sigma-Aldrich (Saint Louis, MO, USA). Argon (pure) was purchased from Air Liquide (Paris, France).

2.2. Samples Preparation

Titanium dioxide nanotubes (TiO2-NT) were prepared via an anodization process according to the procedure described in our previous paper [24]. The Cu2O electrochemical deposition process was performed in a three-electrode cell: TiO2-NT—working electrode, Pt—counter electrode, Ag/AgCl—reference electrode. Before each deposition of Cu2O, a fresh electrolyte was prepared consisting of 50 cm3 0.4 M copper(II) sulfate(VI), 12.6 cm3 11.8 M lactic acid. The pH = 12 was established by adding the appropriate volume of 4.0 M NaOH. Electrodeposition was carried out in a solution heated up to a temperature of 60 °C. The potential difference was equal to −0.36 V during the process and the time lasted from 30 to 180 s. After deposition, TiO2-NT@Cu2O samples were rinsed with deionized water and dried in air at an ambient temperature. Detailed parameters of Cu2O deposition process are reported in Table 2.

2.3. Characterisation Techniques

Scanning Electron Microscope (SEM) NOVA NANO SEM 200 (FEI EUROPE COMPANY, Hillsboro, OR, USA) was used to observe the surface morphology. Raman spectra were recorded with the use of Witec Alpha 300M+ (Ulm, Germany) equipped with a blue laser (λ = 488 nm) and a confocal microscope with an Epiplan-Neofluar ZEISS (Oberkochen, Germany) lens (magnification of 100×). The photoelectrochemical studies were performed in a three-electrode custom-made photoelectrochemical cell (PEC) in the dark and under the illumination of white light. The photoanode was illuminated by a Xe lamp with a power of 450 W and a power density equal to 100 mW/cm2 was used. Three-electrode system consisting of TiO2-NT or TiO2@Cu2O photoanodes which served as a working electrode, a saturated calomel electrode (SCE) as the reference electrode, and Pt-electrode covered with Pt black as the counter electrode were used. Two types of characteristics were measured: current–voltage (Iph–V) and current–time (Iph–t). The following parameters were analyzed: current density measured in the dark, photocurrent density (Iph), flat band potential (Vfb). The stability of photoelectrodes was evaluated based on photocurrent kinetics. As an alternative method to photocatalysis, photoelectrocatalytic measurements were performed as a combination of photocatalysis and photoelectrolysis. Photoelectrocatalysis was carried out in a system designed for PEC measurements—commercial PECC-2 (ZAHNER-Elektrik GmbH & CoKG, Kronach, Germany). Three electrodes—photoelectrode, Pt, and Ag/AgCl—were immersed in the electrolyte consisting of 0.1 M Na2SO4 solution and 1.25 · 10−5 M methylene blue (MB). A potential difference equal to 1 V was applied to the electrodes during the decomposition process. The external voltage causes the separation of photogenerated charge carriers and thus limits their recombination. The protoelectrocatalytic procedure was as follows. First, determination of the adsorption–desorption equilibrium—the electrolyte with electrodes immersed in it was stirred in the dark for 30 min. Second, the Xe lamp was switched on and a constant potential difference of 1 V was applied. At specified intervals, 4 mL of solution was collected and injected into the measurement cuvettes. The absorbance of the solution was measured in quartz cuvettes using a V-670 UV-VIS-NIR spectrophotometer (Jasco, Tokyo, Japan). After that, the solution was placed back into the PECC-2. For comparison purposes, photoelectrocatalytic decomposition was also performed in the dark and with no external voltage. To determine which reactions occur during the process of photoelectrocatalytic decomposition of MB, three types of scavengers that show affinity to different reactive species were used [30,31,32,33]: tert-butanol—hydroxyl radicals, p-benzoquinone—superoxide radical anions, and ethylenediaminetetraacetic acid disodium—electron holes. The type of conductivity was determined with the use of electrochemical impedance spectroscopy based on admittance spectra and the Mott–Schottky plot.

3. Results

SEM surface and cross-section images for TiO2-NT and TiO2-NT@Cu2O system prepared in the 5–180 s electrodeposition process are presented in Figure 1 and Figure S1. The sample obtained during 5 s deposition is characterized by Cu2O crystals of the smallest size, not exceeding 100 nm. An increase in deposition time up to 30 s slightly increases the crystal sizes up to approximately 300 nm and their quantity. On the other hand, electrodeposition lasting for 180 s leads not only to the complete coverage of the surface of the nanotubes, but also to a significant increase in the size of Cu2O crystals ranging from 0.2 to 3.0 μm. The thickness of the Cu2O layer is 1.67 μm.
Figure 2 shows Raman spectra of TiO2 nanotubes and selected TiO2-NT@Cu2O heterostructures. The spectrum of titanium dioxide nanotubes displays five Raman modes: 150, 203, 395, 516, and 635 cm−1. All of them can be assigned to the TiO2 anatase polymorph [34]. After the electrodeposition of Cu2O, as expected, the modes arising from TiO2 are widened due to partial overlap with new modes derived from copper(I) oxide. What is more, as the prolongation of the electrodeposition time from 5 to 180 s results in a complete coverage of the surface of nanotubes by Cu2O (see Figure 1), Cu2O Raman modes in the spectrum become better pronounced.
Photoelectrochemical properties of TiO2-NT@Cu2O heterostructures were determined based on the results of current–time and current–voltage measurements. First of all, the tested photoelectrodes, both TiO2-NT and heterostructures, demonstrate stability under photoelectrochemical measurement conditions. Figure 3a shows the changes in the kinetics of the photocurrent induced by sudden switching on and off the light illuminating the photoanode. Comparison of the results obtained for TiO2 nanotubes and TiO2-NT@Cu2O heterostructures reveals that the shortest deposition time of copper(I) oxide on the titanium dioxide surface leads to the increase in the photocurrent density of ca. 20%. On the other hand, the longest time of Cu2O electrodeposition results in a nearly two-fold decrease in Iph compared to TiO2-NT. Figure 3b presents the Iph–V characteristics of the PEC cell. Current–voltage curves take a typical shape for a n-type semiconductor photoanode. Heterostructures prepared in the process of deposition lasting no longer than 30 s allowed us to obtain photocurrent density values higher than that for TiO2 nanotubes. The consequence of the longest electrodeposition (180 s) is the noticeable reduction in Iph values. The ratio of photocurrent density obtained for heterostructural photoanodes (Istruct) was also compared to that for TiO2 nanotubes (ITiO2-NT). Istruct/ITiO2-NT higher than one means that the photocurrent density for heterostructural photoanodes is higher than for TiO2-NT. The flat band potential is negative for all electrodes and is equal to −0.54 ± 0.03 V. Efficiencies calculated for the selected TiO2@Cu2O heterostructures have shown that there is a significant improvement in relation to TiO2 nanotubes and equal to 3.27 and 2.32 @ 0.5 V, respectively. What is more, the value it achieves is higher not only compared to TiO2 thin film layer, but also compared to that for the TiO2@SnO2 (Table 1).
Photoelectrochemical properties of photoanodes, i.e., flat band potential, photocurrent density at U = 0 and 1 V, and the Istruct/ITiO2-NT ratio at U = 0 and 1 V are summarized in Table 3.
There is a morphological requirement for heterostructures that are applicable in photocatalysis and photoelectrochemistry. For such a system to work efficiently, it is necessary to have quadruple points, i.e., places where access to light and electrolytes is provided to all components of the heterostructure remaining in direct contact with each other. If the layer–particles system is regarded, fulfillment of such a requirement is possible when one of the components is dispersed on the surface of the other. At the same time, it is important that there are many such points in the heterostructure to ensure good photoactivity. The results presented above (Table 3) are in conformity with the observations of morphological differences. The requirements for light and electrolyte access for both components of the heterostructure are fulfilled by the sample obtained during a 5 s deposition. Small copper(I) oxide crystals are distributed evenly on the surface of the nanotubes. On the other hand, the complete coverage of the surface of TiO2 nanotubes with a layer of Cu2O (Figure 1) means that after illumination of the electrode, chemical reactions occur only on the surface of Cu2O. Thus, meeting the morphological criterion in the case of heterostructures with a broad-band semiconductor allows for obtaining a photoanode with better properties than TiO2 nanotubes.
TiO2-NT@Cu2O heterostructures were also tested in the process of photoelectrocatalysis. The choice of organic dyes is very wide. From among them, methylene blue (MB) belongs to the group of cationic dyes. It is introduced into nature as a water pollutant from the textile industry and is highly toxic and carcinogenic. It is also a model pollutant in the study of photocatalytic activity of semiconductors [35]. That is why the ability of the photoanodes to degrade pollutants was evaluated against MB under sunlight simulated by an Xe bulb and applied a potential difference equal to 1V. Typical spectral dependences of MB absorbance obtained before and during photoelectrocatalytic decomposition are presented in Figure 4a. The time required to establish the adsorption-desorption balance of methylene blue was 30 min. Since the presence of leuco methylene blue was excluded (please refer to Supplementary Materials, Figure S2), the decrease in absorbance can be directly correlated with the decreasing concentration of the partially decomposed dye. Figure 4b shows the percentage of decomposed MB after 60 min of the process for TiO2 nanotubes and heterostructures. It is worth noting that all heterostructures show greater activity towards methylene blue decomposition than TiO2-NT, even though it removes 49% of MB. The results obtained for TiO2-NT@Cu2O heterostructures prepared in the electrodeposition process lasting for 5, 15, 30, and 180 s are 99%, 86%, 61%, and 62%, respectively. These results correlate well with the results of photoelectrochemical measurements (see Figure 3b).
In terms of the practical application of photocatalysts, stability under measurement conditions is an important issue. Representative TiO2@Cu2O heterostructure 5/NT was examined in recycle experiments. Photoelectrocatalytic degradation of MB was repeated for four cycles (Figure 5). After every 120 min cycle, almost 90% of the dye was degraded. This indicates that the TiO2@Cu2O heterostructure is completely stable under those conditions.
The comparison of the percent of decomposed methylene blue in the process of photocatalytic or photoelectrocatalytic decomposition under different conditions is presented in Figure 6. TiO2@Cu2O heterostructure was tested both in the dark and under illumination with white light. When the anode was not biased and not illuminated, the percentage of decomposed MB was negligible (0.1%), however, after illumination this value increased to 6.01%. The application of a potential difference of 1 V leads to a sharp increase in the photoactivity of TiO2@Cu2O and after a time as short as 15 min, 29.92% of MB is degraded. This also means that it is necessary to simultaneously illuminate and polarize the working electrode for the process of photoelectrocatalytic decomposition of methylene blue occurred. For comparison, TiO2 nanotubes under the same conditions (1 V, white light) allowed us to obtain only 5.25%. An almost six-fold difference in the percent of decomposed methylene blue clearly indicates an improvement in the photoelectrocatalytic properties of TiO2 nanotubes associated with the presence of copper(I) oxide in the TiO2@Cu2O heterostructure. There are several factors contributing to this effect. First, the range of light absorption of TiO2 (UV) is extended by visible light due to the presence of Cu2O in the heterostructure. Second, recombination of charge carriers is prevented through the separation of electrons and holes between components of the heterostructure. Similar results regarding the TiO2@Cu2O system were described by Ma et al. [36]. In their study, they used trichlorophenol as a model for organic pollutant. They postulated that the improved photoactivity of TiO2@Cu2O in comparison to bare TiO2 results not only from the widened absorption range of light, but also from the fact that TiO2 and Cu2O formed a type II heterostructure. This confirms our assumptions, as type II heterojunction provides an efficient separation of charge carriers between heterostructure elements. Additionally, in the case of photoelectrocatalysis, this is supported by the external voltage.
The process of photoelectrocatalytic decomposition of organic dyes occurs near the photoanode, which indicates the oxidation process is responsible for the degradation [36,37]. The reactive oxygen species that are formed after excitation of the semiconductor with light are as follows:
h + + H 2 O     H + + OH
e + O 2   O 2
O 2 + H +   HO 2
2   HO 2   H 2 O 2 + O 2
H 2 O 2 + e   OH + OH
However, according to da Silva et al. [38], the process of photoelectrocatalytic decomposition of methylene blue occurs mainly as a result of reactions involving electron holes and hydroxyl radicals. To study which of the reactive species are involved in the decomposition reaction of methylene blue in the system with TiO2@Cu2O (5/NT) photoanode, photoelectrocatalytic processes were carried out with the use of different scavengers that show affinity to hydroxyl radicals (2-propanol), oxygen radicals (p-Benzoquinone), and electron holes (ethylenediaminetetraacetic acid disodium). The results obtained at 1 V and after illumination with white light for 15 min are presented in Figure 6. The elimination of oxygen radicals causes degradation of 29.00% of the dye. This is only a percentage point less than that obtained without scavengers under the same conditions. Removal of OH and h , on the other hand, allows to eliminate 19.26 and 21.12%, respectively, which is approximately 10 percentage points less than without scavengers. On this basis, it can be concluded that oxygen radicals do not participate in the process of the degradation of methylene blue. Instead, the removal of hydroxyl radicals and electron holes significantly contributes to the reduction in the degradation process. This effect suggests that OH and h are the basis of the methylene blue degradation mechanism in the discussed process.

4. Discussion

The type of heterojunction is closely related to the type of semiconductor, i.e., p-type or n-type, band-gap energy and band edges position. The n- and p-type conductivity of the photoelectrode based on titanium dioxide nanotubes TiO2-NT and copper(I) oxide Cu2O were determined on the basis of the analysis of the photocurrent under illumination with white light as a function of the applied bias voltage (VB) and Mott–Schottky (M–S) plots. Figure 7 illustrates the methods of analysis of the experimental data for a hypothetical photoanode (n-type) and photocathode (p-type), and the results obtained for the photoelectrodes studied in this work. The M–S plot will possess a negative slope for p-type materials and a positive slope for n-type materials.
In the case of a PEC with a photoanode, the photocurrent characteristic shows a high value of the anodic current, in contrast to the photocathode, which is dominated by the cathodic current. Based on the impedance spectroscopy measurements performed for Cu2O, TiO2, and TiO2@Cu2O, Mott–Schottky plots were drawn and are presented in Figure 7 as the C−2–VB dependence. The slope of the rectilinear range of C−2–VB dependence for the tested photoelectrodes takes positive values. The photocurrent–voltage characteristics also confirm n-type conductivity of the samples. At low temperature, undoped titanium dioxide is well known to be an n-type semiconductor. Although copper(I) oxide is as p-type semiconductor, it was also proved to show n-type conductivity. It is possible to control the type of Cu2O conductivity by careful design of the conditions of solvothermal [39] or electrochemical deposition process [40,41].
The current–voltage characteristics of the photoelectrochemical cell with TiO2@Cu2O photoanode show an increase in photocurrent and solar-to-chemical energy conversion efficiency in relation to the TiO2 anode. Similarly, the photoelectrocatalytic degradation of MB reveals an increase in the activity of heterostructures compared to TiO2. The nature of Cu2O conductivity in the vast majority identified as p-type indicates a possibility of creating a TiO2@Cu2O p-n system of either type II junction (Figure 8a) or a Z-scheme (Figure 8b). Both configurations of the p-n type TiO2@Cu2O heterostructure can be considered when the system acts as a photocatalyst. Aguirre et al. [42] indicated that the TiO2@Cu2O system plays not only the role of a photocatalyst in the CO2 reduction reaction, but also prevents Cu2O from photocorrosion if the heterostructure acts as a Z-scheme. In this case, the electron holes from the Cu2O valence band in the aftermath of recombination with electrons from the TiO2 conduction band cannot participate in the process of Cu2O photocorrosion. The photoanode based on TiO2 modified with copper(I) oxide in a photoelectrochemical cell requires a directed flow of electrons from the anode to the cathode, which means that processes occurring according to the Z-scheme can be difficult. As mentioned before, n-type conductivity of Cu2O in the form of layers obtained by the electrodeposition is well known in the literature [43,44]. In this work, n-type conductivity both for Cu2O and TiO2@Cu2O was proved based on both PEC current–voltage characteristics and Mott–Schottky plots. Additionally, taking into account the fact that the flat band potential can be practically identified as the edge of the conduction band, the value of the flat band potential obtained from the M–S plot and Iph–V measurements makes it possible to determine the bottom of the conduction band. Flat band potential Vfb for Cu2O is equal to −0.29 V, whereas for TiO2 nanotubes and TiO2@Cu2O it is comparable and equal to −0.50 and −0.56 V, respectively.
On the basis of the structural, electrical, and photoelectrochemical results, TiO2@Cu2O heterostructure was proposed to create a heterojunction resulting from the combination of TiO2 and Cu2O semiconductors in type II n-n configuration (Figure 8c).
In the case of Cu2O, the potentials for anodic and cathodic decomposition are located within the band-gap [45], and hence it is necessary to prevent its corrosion. Paracchino et al. named the conditions that should be met in the case of a n-type semiconductor/p-type Cu2O heterojunction to not only achieve high efficiency of photoelectrochemical processes but also provide protection against photocorrosion [46]. Among them, conditions of the relative position of the band edges of n-type semiconductor and redox potentials of water; the conduction band edge should be placed above the water reduction level Ered (H+/H2). On the other hand, within its band-gap, there should be no decomposition potentials. The results obtained in this work indicate that in the case of a n-n semiconductor system based on TiO2@Cu2O, titanium dioxide prevents copper(I) oxide from corrosion by accelerating the transport of electron holes into the electrolyte. This is directly confirmed by the results of repeated photoelectrocatalytic decomposition of methylene blue shown in Figure 5. The percentage of MB decomposed in the experiments conducted did not decrease after four uses. Such a behavior was previously demonstrated by us for TiO2@MoS2 in photoelectrochemical water decomposition [24].

5. Conclusions

Electrodeposition carried out using a mixture of CuSO4, lactic acid, and NaOH, with pH = 12, using the difference potential of −0.36 V and short reaction time, allowed to obtain a discontinuous layer of Cu2O. This effect is beneficial from the point of view of the photoelectrochemical properties of the resulting Cu2O@TiO2-NT heterojunction, because it allows absorption in both the UV (TiO2) and vis (Cu2O) range and electrolyte access to both components of the heterostructure. The obtained Cu2O@TiO2-NT system was successfully used as a photoelectrode during the alternative to photocatalysis, photoelectrocatalytic decomposition of methylene blue, as a result of which an improvement in properties relative to unmodified titanium dioxide was demonstrated. A new configuration of band-gap alignment for heterojunction of Cu2O@TiO2-NT was proposed. The analysis of research results indicated that the n-n semiconductor type II heterojunction of Cu2O/TiO2 had formed. It was found that it prevents the processes of photocorrosion of a semiconductor with a narrower band gap contained in a TiO2-based heterostructure.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma14133725/s1; Figure S1. SEM images of TiO2-NT: (a) surface and (b) cross-section; Figure S2. Absorption spectra of the MB + Na2SO4 mixture before and after photoelectrocatalysis of MB obtained for 5/NT heterostructure. Inset: position of the band derived from MB (blue) and LMB (orange).

Author Contributions

Conceptualization, A.T.-Z. and M.R.; Data curation, A.T.-Z. and J.B.-G.; Formal analysis, A.T.-Z., J.B.-G. and M.R.; Funding acquisition, M.R.; Investigation, A.T.-Z. and J.B.-G.; Methodology, A.T.-Z., J.B.-G. and M.R.; Resources, J.B.-G. and M.R.; Supervision, A.T.-Z. and M.R.; Validation, A.T.-Z., J.B.-G. and M.R.; Visualization, A.T.-Z., J.B.-G. and M.R.; Writing—original draft, A.T.-Z., J.B.-G. and M.R.; Writing—review and editing, A.T.-Z. and M.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Center NCN, grant no. 2016/21/B/ST8/00457.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

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. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  2. Marschall, R. Semiconductor composites: Strategies for enhancing charge carrier separation to improve photocatalytic activity. Adv. Funct. Mater. 2014, 24, 2421–2440. [Google Scholar] [CrossRef]
  3. Li, J.; Wu, N. Semiconductor-based photocatalysts and photoelectrochemical cells for solar fuel generation: A review. Catal. Sci. Technol. 2015, 5, 1360–1384. [Google Scholar] [CrossRef]
  4. Thalluri, S.M.; Bai, L.; Lv, C.; Huang, Z.; Hu, X.; Liu, L. Strategies for Semiconductor/Electrocatalyst Coupling toward Solar-Driven Water Splitting. Adv. Sci. 2020, 7. [Google Scholar] [CrossRef] [Green Version]
  5. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  6. Garcia-Segura, S.; Brillas, E. Applied photoelectrocatalysis on the degradation of organic pollutants in wastewaters. J. Photochem. Photobiol. C Photochem. Rev. 2017, 31, 1–35. [Google Scholar] [CrossRef]
  7. Oturan, M.A.; Aaron, J.J. Advanced oxidation processes in water/wastewater treatment: Principles and applications. A review. Crit. Rev. Environ. Sci. Technol. 2014, 44, 2577–2641. [Google Scholar] [CrossRef]
  8. Daghrir, R.; Drogui, P.; Robert, D. Photoelectrocatalytic technologies for environmental applications. J. Photochem. Photobiol. A Chem. 2012, 238, 41–52. [Google Scholar] [CrossRef]
  9. Bessegato, G.G.; Guaraldo, T.T.; de Brito, J.F.; Brugnera, M.F.; Zanoni, M.V.B. Achievements and Trends in Photoelectrocatalysis: From Environmental to Energy Applications. Electrocatalysis 2015, 6, 415–441. [Google Scholar] [CrossRef] [Green Version]
  10. Radecka, M.; Kusior, A.; Trenczek-Zajac, A.; Zakrzewska, K. Oxide Nanomaterials for Photoelectrochemical Hydrogen Energy Sources. In Materials for Sustainable Energy; Academic Press: Cambridge, MA, USA, 2018; Volume 72, pp. 145–183. ISBN 9780128150771. [Google Scholar]
  11. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  12. Henderson, M.A. A surface science perspective on TiO2 photocatalysis. Surf. Sci. Rep. 2011, 66, 185–297. [Google Scholar] [CrossRef]
  13. Ahmad, H.; Kamarudin, S.K.; Minggu, L.J.; Kassim, M. Hydrogen from photo-catalytic water splitting process: A review. Renew. Sustain. Energy Rev. 2015, 43, 599–610. [Google Scholar] [CrossRef]
  14. Tee, S.Y.; Win, K.Y.; Teo, W.S.; Koh, L.D.; Liu, S.; Teng, C.P.; Han, M.Y. Recent Progress in Energy-Driven Water Splitting. Adv. Sci. 2017, 4, 1600337. [Google Scholar] [CrossRef] [Green Version]
  15. Zhang, L.; Jaroniec, M. Toward designing semiconductor-semiconductor heterojunctions for photocatalytic applications. Appl. Surf. Sci. 2018, 430, 2–17. [Google Scholar] [CrossRef]
  16. Roy, P.; Berger, S.; Schmuki, P. TiO2 nanotubes: Synthesis and applications. Angew. Chemie Int. Ed. 2011, 50, 2904–2939. [Google Scholar] [CrossRef]
  17. Wehrenfennig, C.; Palumbiny, C.M.; Snaith, H.J.; Johnston, M.B.; Schmidt-Mende, L.; Herz, L.M. Fast Charge-Carrier Trapping in TiO2 Nanotubes. J. Phys. Chem. C 2015, 119, 9159–9168. [Google Scholar] [CrossRef]
  18. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1–20. [Google Scholar] [CrossRef]
  19. Sun, B.; Shi, T.; Peng, Z.; Sheng, W.; Jiang, T.; Liao, G. Controlled fabrication of Sn/TiO2 nanorods for photoelectrochemical water splitting. Nanoscale Res. Lett. 2013, 8, 1–8. [Google Scholar] [CrossRef] [Green Version]
  20. Sun, Y.; Yan, K.P. Effect of anodization voltage on performance of TiO2 nanotube arrays for hydrogen generation in a two-compartment photoelectrochemical cell. Int. J. Hydrogen Energy 2014, 39, 11368–11375. [Google Scholar] [CrossRef]
  21. Radecka, M.; Rekas, M.; Trenczek-Zajac, A.; Zakrzewska, K. Importance of the band gap energy and flat band potential for application of modified TiO2 photoanodes in water photolysis. J. Power Sources 2008, 181, 46–55. [Google Scholar] [CrossRef]
  22. Menon, H.; Gopakumar, G.; Nair, V.S.; Nair, S.V.; Shanmugam, M. 2D-Layered MoS2-Incorporated TiO2-Nanofiber-Based Dye-Sensitized Solar Cells. ChemistrySelect 2018, 3, 5801–5807. [Google Scholar] [CrossRef]
  23. Cheng, C.; Ren, W.; Zhang, H. 3D TiO2/SnO2 hierarchically branched nanowires on transparent FTO substrate as photoanode for efficient water splitting. Nano Energy 2014, 5, 132–138. [Google Scholar] [CrossRef]
  24. Radecka, M.; Wnuk, A.; Trenczek-Zajac⁠, A.; Schneider, K.; Zakrzewska⁠, K. TiO2/SnO2 nanotubes for hydrogen generation by photoelectrochemical water splitting. Int. J. Hydrogen Energy 2015, 40, 841–851. [Google Scholar] [CrossRef]
  25. Yuan, W.; Yuan, J.; Xie, J.; Li, C.M. Polymer-Mediated Self-Assembly of TiO2@Cu2O Core-Shell Nanowire Array for Highly Efficient Photoelectrochemical Water Oxidation. ACS Appl. Mater. Interfaces 2016, 8, 6082–6092. [Google Scholar] [CrossRef]
  26. Yang, Y.; Xu, D.; Wu, Q.; Diao, P. Cu2O/CuO bilayered composite as a high-efficiency photocathode for photoelectrochemical hydrogen evolution reaction. Sci. Rep. 2016, 6, 35158. [Google Scholar] [CrossRef] [Green Version]
  27. Zhang, Q.; Zhang, K.; Xu, D.; Yang, G.; Huang, H.; Nie, F.; Liu, C.; Yang, S. CuO nanostructures: Synthesis, characterization, growth mechanisms, fundamental properties, and applications. Prog. Mater. Sci. 2014, 60, 208–337. [Google Scholar] [CrossRef]
  28. Koiki, B.A.; Arotiba, O.A. Cu2O as an emerging semiconductor in photocatalytic and photoelectrocatalytic treatment of water contaminated with organic substances: A review. RSC Adv. 2020, 10, 36514–36525. [Google Scholar] [CrossRef]
  29. Sun, S.; Zhang, X.; Yang, Q.; Liang, S.; Zhang, X.; Yang, Z. Cuprous oxide (Cu2O) crystals with tailored architectures: A comprehensive review on synthesis, fundamental properties, functional modifications and applications. Prog. Mater. Sci. 2018, 96, 111–173. [Google Scholar] [CrossRef] [Green Version]
  30. Kusior, A.; Michalec, K.; Jelen, P.; Radecka, M. Shaped Fe2O3 nanoparticles – Synthesis and enhanced photocatalytic degradation towards RhB. Appl. Surf. Sci. 2019, 476, 342–352. [Google Scholar] [CrossRef]
  31. Lin, Z.; Li, L.; Yu, L.; Li, W.; Yang, G. Dual-functional photocatalysis for hydrogen evolution from industrial wastewaters. Phys. Chem. Chem. Phys. 2017, 19, 8356–8362. [Google Scholar] [CrossRef]
  32. Trenczek-Zajac, A. Thermally oxidized CdS as a photoactive material. New J. Chem. 2019, 43, 8892–8902. [Google Scholar] [CrossRef]
  33. Shen, H.; Zhao, X.; Duan, L.; Liu, R.; Wu, H.; Hou, T.; Jiang, X.; Gao, H. Influence of interface combination of RGO-photosensitized SnO2@RGO core-shell structures on their photocatalytic performance. Appl. Surf. Sci. 2017, 391, 627–634. [Google Scholar] [CrossRef]
  34. Ohsaka, T.; Izumi, F.; Fujiki, Y. Raman spectrum of anatase, TiO2. J. Raman Spectrosc. 1978, 7, 321–324. [Google Scholar] [CrossRef]
  35. Xu, Y.H.; Liang, D.H.; Liu, M.L.; Liu, D. zhong Preparation and characterization of Cu2O-TiO2: Efficient photocatalytic degradation of methylene blue. Mater. Res. Bull. 2008, 43, 3474–3482. [Google Scholar] [CrossRef]
  36. Ma, Q.; Zhang, H.; Cui, Y.; Deng, X.; Guo, R.; Cheng, X.; Xie, M.; Cheng, Q. Fabrication of Cu2O/TiO2 nano-tube arrays photoelectrode and its enhanced photoelectrocatalytic performance for degradation of 2,4,6-trichlorophenol. J. Ind. Eng. Chem. 2018, 57, 181–187. [Google Scholar] [CrossRef]
  37. Liu, D.; Tian, R.; Wang, J.; Nie, E.; Piao, X.; Li, X.; Sun, Z. Photoelectrocatalytic degradation of methylene blue using F doped TiO2 photoelectrode under visible light irradiation. Chemosphere 2017, 185, 574–581. [Google Scholar] [CrossRef] [PubMed]
  38. da Silva, M.R.; Lucilha, A.C.; Afonso, R.; Dall’Antonia, L.H.; de Andrade Scalvi, L.V. Photoelectrochemical properties of FTO/m-BiVO4 electrode in different electrolytes solutions under visible light irradiation. Ionics 2014, 20, 105–113. [Google Scholar] [CrossRef]
  39. Xiong, L.; Huang, S.; Yang, X.; Qiu, M.; Chen, Z.; Yu, Y. P-Type and n-type Cu2O semiconductor thin films: Controllable preparation by simple solvothermal method and photoelectrochemical properties. Electrochim. Acta 2011, 56, 2735–2739. [Google Scholar] [CrossRef]
  40. Fernando, C.A.N.; De Silva, P.H.C.; Wethasinha, S.K.; Dharmadasa, I.M.; Delsol, T.; Simmonds, M.C. Investigation of n-type Cu2O layers prepared by a low cost chemical method for use in photo-voltaic thin film solar cells. Renew. Energy 2002, 26, 521–529. [Google Scholar] [CrossRef]
  41. McShane, C.M.; Choi, K.S. Photocurrent enhancement of n-type Cu2O electrodes achieved by controlling dendritic branching growth. J. Am. Chem. Soc. 2009, 131, 2561–2569. [Google Scholar] [CrossRef]
  42. Aguirre, E.M.; Zhou, R.; Eugene, A.J.; Guzman, M.I.; Grela, M.A. Cu2O/TiO2 heterostructures for CO2 reduction through a direct Z-scheme: Protecting Cu2O from photocorrosion. Appl. Catal. B Environ. 2017, 217, 485–493. [Google Scholar] [CrossRef]
  43. Siripala, W. Electrodeposition of n-type Cuprous Oxide Thin Films. ECS Trans. 2008, 11, 1–10. [Google Scholar] [CrossRef]
  44. Garuthara, R.; Siripala, W. Photoluminescence characterization of polycrystalline n-type Cu2O films. J. Lumin. 2006, 121, 173–178. [Google Scholar] [CrossRef]
  45. Gerischer, H. Solar Photoelectrolysis with Semiconductor Electrodes. In Solar Energy Conversion; Seraphin, B.O., Ed.; Springer: Berlin/Heidelberg, Germany, 1979; Volume 31, p. 338. ISBN 978-3-662-30849-3. [Google Scholar]
  46. Paracchino, A.; Laporte, V.; Sivula, K.; Grätzel, M.; Thimsen, E. Highly active oxide photocathode for photoelectrochemical water reduction. Nat. Mater. 2011, 10, 456–461. [Google Scholar] [CrossRef]
Figure 1. SEM images of TiO2-NT@Cu2O heterostructures obtained by electrodeposition method in a solution of pH = 12. Deposition parameters: U = −0.36 V, t = 5–180 s.
Figure 1. SEM images of TiO2-NT@Cu2O heterostructures obtained by electrodeposition method in a solution of pH = 12. Deposition parameters: U = −0.36 V, t = 5–180 s.
Materials 14 03725 g001
Figure 2. Raman spectra of TiO2-NT and TiO2-NT@Cu2O heterostructures obtained after 5 and 180 s of electrodeposition.
Figure 2. Raman spectra of TiO2-NT and TiO2-NT@Cu2O heterostructures obtained after 5 and 180 s of electrodeposition.
Materials 14 03725 g002
Figure 3. Photoelectrochemical characteristics of TiO2-NT and TiO2-NT@Cu2O anodes: (a) current-time and (b) current-voltage curves.
Figure 3. Photoelectrochemical characteristics of TiO2-NT and TiO2-NT@Cu2O anodes: (a) current-time and (b) current-voltage curves.
Materials 14 03725 g003
Figure 4. (a) Absorption spectra of the MB + Na2SO4 mixture before and after photoelectrocatalysis of MB obtained for 5/NT heterostructure. (b) Comparison of the amount of photoelectrocatalytically decomposed MB after 120 min for different TiO2@Cu2O heterostructures.
Figure 4. (a) Absorption spectra of the MB + Na2SO4 mixture before and after photoelectrocatalysis of MB obtained for 5/NT heterostructure. (b) Comparison of the amount of photoelectrocatalytically decomposed MB after 120 min for different TiO2@Cu2O heterostructures.
Materials 14 03725 g004
Figure 5. Recycled photoelectrocatalytic test of 5/NT sample.
Figure 5. Recycled photoelectrocatalytic test of 5/NT sample.
Materials 14 03725 g005
Figure 6. Comparison of the amount of photocatalytically and photoelectrocatalytically decomposed MB after 15 min of process under different conditions: light—illumination with white light, dark—no illumination, 0 V—zero voltage, 1 V—external voltage, pB—p-benzoquinone, EDTA:Na2—ethylenediaminetetraacetic acid disodium.
Figure 6. Comparison of the amount of photocatalytically and photoelectrocatalytically decomposed MB after 15 min of process under different conditions: light—illumination with white light, dark—no illumination, 0 V—zero voltage, 1 V—external voltage, pB—p-benzoquinone, EDTA:Na2—ethylenediaminetetraacetic acid disodium.
Materials 14 03725 g006
Figure 7. Examples of theoretical and experimental results of selected experimental methods based on photoelectrochemical and electrochemical measurements for determination of the type of conductivity—n- or p-type.
Figure 7. Examples of theoretical and experimental results of selected experimental methods based on photoelectrochemical and electrochemical measurements for determination of the type of conductivity—n- or p-type.
Materials 14 03725 g007
Figure 8. Schematic presentation of possible TiO2@Cu2O type II heterostructures: p-n heterojunction [42] (a), Z-scheme (b), and n-n heterojunction (c). ECB—conduction band of semiconductor, EVB—valence band of semiconductor, 1, 2—first and second semiconductor, EF—Fermi level.
Figure 8. Schematic presentation of possible TiO2@Cu2O type II heterostructures: p-n heterojunction [42] (a), Z-scheme (b), and n-n heterojunction (c). ECB—conduction band of semiconductor, EVB—valence band of semiconductor, 1, 2—first and second semiconductor, EF—Fermi level.
Materials 14 03725 g008
Table 1. Solar-to-chemical energy conversion efficiencies η of TiO2 and TiO2-based heterostructures. Potential difference was measured relative to the SCE electrode unless stated otherwise.
Table 1. Solar-to-chemical energy conversion efficiencies η of TiO2 and TiO2-based heterostructures. Potential difference was measured relative to the SCE electrode unless stated otherwise.
SystemPhotoelectrodeEfficiency η (%) Ref.
at 0 (V)at 0.5 (V)
TiO2TiO2 nanorods0.15 1[19]
TiO2 nanotubes0.25–1.6[20]
TiO2 thin film1.651.90[21]
TiO2@MoS2TiO2 nanofibers4.7 2[22]
TiO2@MoS26.0 2
TiO2@SnO2TiO2 nanowires[23]
TiO2@SnO20.21 1
TiO2 nanotubes0.340.39[24]
TiO2@SnO20.39–1.610.47–2.12
TiO2@Cu2OTiO2 nanowires0.14[25]
TiO2@Cu2O0.39
1 Potential difference for Ag/AgCl electrode. 2 No information about the potential difference.
Table 2. Parameters of TiO2-NT@Cu2O heterostructures formation process.
Table 2. Parameters of TiO2-NT@Cu2O heterostructures formation process.
Electrode NamepHPotential Difference (V)Time (s)Substrate
5/NT12−0.365TiO2-NT
15/NT15
30/NT30
180/NT180
Table 3. Photoelectrochemical parameters and conductivity type of TiO2-NT and TiO2-NT@Cu2O heterostructures.
Table 3. Photoelectrochemical parameters and conductivity type of TiO2-NT and TiO2-NT@Cu2O heterostructures.
Electrode NameIph (μA/cm2)Ihetero/ITiO2-NTVfb (V)
at 0 (V)at 1 (V)at 0 (V)at 1 (V)Mott-SchottkyI-V 1
TiO2-NT61.9324.9−0.50−0.55
5/NT94.9470.71.531.45−0.52
15/NT94.5398.91.531.23−0.56−0.56
30/NT86.8392.41.401.21−0.46
180/NT50.0218.90.810.67−0.50
1 For determination method, see [21].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Trenczek-Zajac, A.; Banas-Gac, J.; Radecka, M. TiO2@Cu2O n-n Type Heterostructures for Photochemistry. Materials 2021, 14, 3725. https://doi.org/10.3390/ma14133725

AMA Style

Trenczek-Zajac A, Banas-Gac J, Radecka M. TiO2@Cu2O n-n Type Heterostructures for Photochemistry. Materials. 2021; 14(13):3725. https://doi.org/10.3390/ma14133725

Chicago/Turabian Style

Trenczek-Zajac, Anita, Joanna Banas-Gac, and Marta Radecka. 2021. "TiO2@Cu2O n-n Type Heterostructures for Photochemistry" Materials 14, no. 13: 3725. https://doi.org/10.3390/ma14133725

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