Next Article in Journal
Synthesis and Modification of Hydroxyapatite Nanofiber for Poly(Lactic Acid) Composites with Enhanced Mechanical Strength and Bioactivity
Next Article in Special Issue
Impact of Iron on the Fe–Co–Ni Ternary Nanocomposites Structural and Magnetic Features Obtained via Chemical Precipitation Followed by Reduction Process for Various Magnetically Coupled Devices Applications
Previous Article in Journal
A Rational Fabrication Method for Low Switching-Temperature VO2
Previous Article in Special Issue
Styrene-Based Elastomer Composites with Functionalized Graphene Oxide and Silica Nanofiber Fillers: Mechanical and Thermal Conductivity Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Trace Amounts of Co3O4 Nano-Particles Modified TiO2 Nanorod Arrays for Boosted Photoelectrocatalytic Removal of Organic Pollutants in Water

1
College of Chemistry and Chemical Engineering, Lanzhou University, Lanzhou 730000, China
2
Key Laboratory of Evidence Science Techniques Research and Application, Gansu University of Political Science and Law, Lanzhou 730070, China
3
College of Chemistry and Chemical Engineering, Northwest Normal University, Lanzhou 730000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2021, 11(1), 214; https://doi.org/10.3390/nano11010214
Submission received: 30 November 2020 / Revised: 2 January 2021 / Accepted: 4 January 2021 / Published: 15 January 2021

Abstract

:
Trace amounts of Co3O4 modified TiO2 nanorod arrays were successfully fabricated through the photochemical deposition method without adding any nocuous reagents. The Co3O4/TiO2 nanorod arrays fabricated in acid solution had the highest photo-electrochemical activity. We elaborated on the mechanism of Co3O4-TiO2 fabricated in different pH value solutions. The Co3O4-TiO2 had a more remarkable photo-electrochemical performance than the pure TiO2 nanorod arrays owing to the heterojunction between Co3O4 and TiO2. The degradation of methylene blue and hydroquinone was selected as the model reactions to evaluate the photo-electrochemical performance of Co3O4-TiO2 nanorod arrays. The Co3O4/TiO2 nanorod arrays had great potential in waste water treatment.

Graphical Abstract

1. Introduction

From the comprehensive point of view, semiconductor metal oxides have been widely used as stable photo-catalysts for the cosmopolitan energy crisis [1,2,3,4,5]. TiO2 is the most extensively used semiconductor photo-catalyst owing to its exceptional properties, such as high photo-catalytic activity, chemical stability, environmental-friendliness, and low cost [3,4,6,7,8]. However, because of the large bandgap of TiO2 (3.2 eV), the practical applications are hampered by its low electrical conductivity, strong reflection, and weak light-harvesting ability [9,10], as well as the rapid combination of photo-generated electron and hole pairs. Various strategies have been utilized to improve the photo-catalytic efficiency of TiO2 materials, such as tuning their crystallite size and structure, sensitizing them by organic dye and quantum dots [8,11,12], and modifying them with metals (e.g., Pt, Ru, Ag, Au, Rh, Pd, Ni, and Co) [3,10,13,14,15,16,17] or transition metal oxides (e.g., Co3O4, CoO, Cu2O, and Fe2O3) [18,19,20,21] with a narrow band gap semiconductor. The formation of hetero-junctions between metal oxides and semiconductors is a useful strategy to suppress the recombination of photo-generated electrons and holes in TiO2, and extend photon absorption into the visible regime, which can enhance the photochemical efficiency of TiO2 nanomaterials.
Semiconductor-based hetero-junctions are able to facilitate fast charge separation and enhance the photo-catalytic efficiency of TiO2 nanomaterials. It is an effective strategy to construct TiO2-based hetero-junction structures with transition metal oxides. Heterojunction structures have the potential to facilitate electron–hole separation. TiO2 is a n-type semiconductor and, combined with a p-type semiconductor in a suitable band gap position to form a p-n heterojunction, it is an effective tactic to expand light absorption, enhance the separation effects of electrons and holes, prolong the lifetime of the electron and hole, and heighten the photocatalytic activity. Transition metal oxides, such as Ag2O, Cu2O, CuO, and Co3O4, have been used to form a p-n junction to promote an interfacial electron transfer process and increase the separation effect [18,22,23,24]. Cobalt oxides have received attention because of their excellent photo-catalytic activity in carbon dioxide reduction, oxygen reduction, and environment restoration [25,26]. Based on its properties of outstanding photocatalytic activity and low cost, cobalt oxide becomes a feasible material to fabricate the heterojunction structure with other semiconductor photocatalysts. In this work, the Co3O4-TiO2 nanorod arrays were synthesized by photochemical deposition (a green method). The performance of Co3O4-TiO2 nanorod arrays could be controlled by adjusting the pH value and the concentration of the Co precursor.

2. Experimental

2.1. Preparation of TiO2 Nanorod Arrays

The TiO2 nanorod arrays on Fluorine doped tin oxide (FTO) were fabricated through the modified hydrothermal method [7,8]. Typically, 30 mL HCl (6 mol/L) was mixed with 0.4 mL tertrabutyl titanium by strong stirring for 10 min. Then, the above solution was transferred into a Teflon pot, in which an FTO glass electrode with the coated layer facing down was placed against the wall of the Teflon pot. The hydrothermal synthesis was carried out at 150 °C for 6 h in an oven. Then, the TiO2 nanorod arrays was washed with high pure water, and dried by high pure N2.

2.2. Fabrication of Ultra Small Co3O4 Coated TiO2 Nanorod Arrays

The Co3O4-coated TiO2 nanorod arrays were prepared by photo-chemical deposition in a 1:1 ethanol/water solution containing different concentrations of Co(NO3)2 under Xe lamp with a powder intensity of 100 mW/cm2. The driving force of the Co deposition on the TiO2 nanorod arrays was high energetic photons, which could photo-excite electrons from the valance band of TiO2 to the conductor band and leave holes on the valance band. The holes were depleted by the hole receptor of ethanol. The electrons can reduce the cobalt ions onto the surface of TiO2 nanorod arrays. Nano cobalt metal was stable and easily oxidized to Co3O4.

2.3. Photo-Electrochemical Studies of the Co3O4 Modified TiO2 Nanorod Arrays

To study the photo-electrochemical response of the Co3O4-modified TiO2 nanorods and pure TiO2 nanorod arrays (1.0 cm × 1.0 cm), open circuit potential (OCP) and the ampere-metric method (IT) were conducted in 0.1 mol/L Na2SO4 solution at room temperature after being deoxidized by high pure N2 for 15 min.

2.4. The Photo-Electrochemical Degradation Research

Methylene blue and hydroquinone were used to test the photo-electrochemical activity of Co3O4 modified TiO2 nanorod arrays and pure TiO2 nanorod arrays. The Co3O4 modified TiO2 and pure TiO2 nanorod arrays with a geometric area of 1.0 cm × 1.0 cm were used as the working electrode, as well as platinum as the counter electrode and saturated calomel electrode (SCE) as the reference electrode. A 500 W Xe lamp was used to simulate sunlight with a powder intensity of 100 mW/cm2. A solution containing 10 mg/L methylene blue or hydroquinone, 0.1 mol/L Na2SO4, and 10 mmol/L H2O2 was used as the investigated subject. During the photo-electrochemical degradation process, the electrode was added with 1.0 V bias potential (vs. SCE) and was light-illuminated to degrade methylene blue or hydroquinone. To avoid the effects of solution temperature, the reaction system was placed in a constant temperature system with a circulating water device (Beijing LabTech Instruments Co., Ltd., Beijing, China). The photo-electrochemical degradation process of methylene blue and hydroquinone was measured by UV/vis spectrum.

3. Results and Discussions

3.1. XRD Analysis

The Co3O4-TiO2 and TiO2 nanorod arrays were investigated by X-ray diffraction measurement (XRD) using copper target to identify the phase of all samples. The Co3O4-TiO2 and TiO2 samples in Figure 1 show the rutile phase of TiO2 (JCPDS No1-1292). No new XRD peaks of Co3O4 were observed for the Co3O4-TiO2 nanorod arrays, which might because of the small amount of cobalt oxide in the Co3O4-TiO2 hybrid catalyst.

3.2. Scanning Electron Microscopy (SEM) and Transmission Electron Microscopy (TEM) Measurement

The morphologies and microstructures of the TiO2 nanorod arrays and Co3O4-TiO2 nanorod arrays were characterized by scanning electron microscopy (SEM). Figure 2 shows the top view SEM images of the TiO2 and Co3O4-TiO2 nanorod arrays. Highly ordered and large scale TiO2 nanorod arrays are vertically aligned on both pure TiO2 and Co3O4-TiO2 nanorod arrays in Figure 2.
The average diameter of TiO2 nanorod is about 180 nm with a rectangular cross section. There are almost no different features between Co3O4-TiO2 nanorod arrays and pure TiO2 nanorod arrays, which might because of the small amount of Co element in the Co3O4-TiO2 nanorod arrays. The small amount of Co element could be confirmed by Energy Dispersive X-ray (EDX) in Supplementary Materials Figure S1. Compared with the super strong signal of Ti and O element, the signal of Co element was very low, which means the cobalt content was very low.
Figure 3a–d show the TEM images of TiO2 nanorods and Co3O4-TiO2 nanorod arrays, respectively. It can be clearly seen that the TiO2 nanorod arrays have the dimension of 4 nm with a clearly lattice structure. Figure 3b is the high resolution of TEM image of TiO2. Compared with the clear lattice structure of pure TiO2 nanorods, the surface of Co3O4-TiO2 nanorod arrays (Figure 3c) was covered with something similar to fog, which makes the lattice structure of Co3O4-TiO2 nanorods not obvious. High resolution TEM of Co3O4-TiO2 nanorod in Figure 3d showed that there were some amorphous Co3O4 on the surface of TiO2.

3.3. X-ray Photoelectron Spectroscopy (XPS) Analysis

To confirm the composition of Co3O4-TiO2 nanorod arrays sample, the XPS method was used to study the chemical composition and valence state of Co3O4-TiO2. The survey spectra illustrated in Supplementary Materials Figure S2 demonstrates the existence of Co, Ti, and O elements. Figure 4a shows the XPS spectrum of O 1s. Figure 4b shows the XPS spectrum of Ti 2p obital of the Co3O4-TiO2 nanorod array. The Ti 2p3/2 and 2p1/2 located at 458.4 eV and 464.1 eV can be assigned to Ti4+, which coincided with TiO2. The band energy of 780.50 eV and 797.43 eV in Figure 4c corresponded to Co 2p3/2 and Co 2p1/2, respectively. The peaks are the typical signature of Co3O4 and are consistent with the previous literature [27]. Compared with the strong intensity of the Ti and O element, the XPS spectra strength of cobalt was very weak, which meant the amount of cobalt element in Co3O4-TiO2 nano-materials was small.
The experimental data were in accordance with the EDX results in Supplementary Materials Table S1. The amount of Co was small, because of the high activity of cobalt metal. Cobalt element is more vivacious than hydrogen element, which means there is a competitive reaction between cobalt ions and hydrogen ions during photochemical deposition. The competition between cobalt ions and hydrogen ions decreases the amount of cobalt deposition on the TiO2 surface. Furthermore, the Co metal easily dissolved into Co ions under acid circumstance. Therefore, the competitive reaction and the instability of cobalt lead to the ultra-small amount of Co on the TiO2 nanorod arrays.

3.4. Photocurrent Test

The transient photocurrent was further used to confirm the generation, transfer, and separation processes of the photo-induced electrons and holes on both Co3O4 modified TiO2 nanorod arrays and the pure TiO2 nanorod arrays. To illustrate the effect of Co3O4 on photocatalytic activity of Co3O4-TiO2 nanorod array, the photocurrent response of the TiO2 nanorod arrays and Co3O4-TiO2 nanorod arrays was measured by chopping light. The curves of both Co3O4-TiO2 and TiO2 samples in Figure 5a had outstanding responses to chopping light cycles. The current value of Co3O4-TiO2 is five times higher than pure TiO2 nanorod arrays. This means that the Co3O4-TiO2 nanorod arrays had higher photoelectrocatalytic activity than the pure TiO2 nanorod arrays.
Compared with pure TiO2 nanorod arrays, the Co3O4-TiO2 nanorod arrays exhibited an obviously higher photocurrent, which indicated that Co3O4-TiO2 nanorod arrays had the higher photo-electrochemical activity. Under light illumination, the electron in the valence band was excited to the conductor band and left a hole in the valence band. The electrons were accumulated on the conductor band and holes were assembled on the valence band by persistent light illumination [28]. The relative OCP value of the samples was measured to compare the performance in a different semiconductor. In comparison, the relative OCP value of Co3O4-TiO2 was higher than the pure TiO2 nanorod arrays, which demonstrated that separation of e-h+ pairs in the Co3O4-TiO2 heterojunction is significantly improved by the addition of Co3O4 (Supplementary Materials Figure S3).

3.5. The Formation Mechanisms of Co3O4 Nanoparticles

Further observation found that Co3O4-TiO2 nanorods fabricated in different pH values had diverse responses to visible light, as shown in Figure 5b. The Co3O4-TiO2 nanorods fabricated under pH 4.12 had the highest photo-electrochemical response compared with others fabricated at pH 7.82 and 7.00. The formation mechanisms of Co3O4 nanoparticles in different pH solutions were different, which led to a different response to visible light. In neutral solution, the Co3O4 were formed under photochemical deposition.
3 Co 8 e + 4 H 2 O = Co 3 O 4 + 8 H +
Co nanoparticles were easily deposited onto the semiconductor in alkaline solution. Such as Co-ZnO was fabricated, which had high catalytic activity in oxygen production [29]. And the high catalytic activity of Ni-CdS nanorods was fabricated through photochemical deposition in NaOH solution containing Ni ions and methanol [30]. Holes have high energy, which could oxidize methanol adsorbing on the semiconductor. Considering the high concentration of ethanol, ethanol was oxidized to formaldhyde and holes were decomposed. It is very difficult to deposit transition metal or metal oxide onto semiconductors in the acid solution because cobalt metal is more active than the hydron element. There were two different competitive reactions during the photodeposition reaction. They are listed below:
Co 2 + + 2 e = Co
2 H + + 2 e = H 2
Furthermore, the high activity of transition metals meant they easily dissolved in acid solution. The Co metal was not stable because of oxidization reaction and photocorrosion. Considering the above reasons, the amount of Co is very small.
Co 2 e = Co 2 +
3 Co 8 e + 4 OH = Co 3 O 4 + 4 H +
In alkaline solution, the Co ions mainly existed through Co(OH)42−, which was reduced to Co metal by electrons photo-excited on the TiO2 surface. The Co metal was not stable and could easily be oxidized to Co3O4 by oxygen or OH radical produced in photochemical deposition.
Co ( OH ) 4 2 + 2 e = Co + 4 OH
The amount of Co3O4 can be controlled through regulating the concentration of Co ions from 0.1 mmol/L to 1 mmol/L. We found that Co3O4-TiO2 nanorod array fabricated in 0.5 mmol/L cobalt ions had the highest photocatalytic activity. It can be clearly seen that the photocurrent increased with the concentration of Co ions increasing from 0.1 mmol/L to 0.5 mmol/L (in Figure 6). However, the photo-electrochemical currents decreased when the Co ions’ concentrations were further increased, which may be because of the formation of a thick Co3O4 layer and increase in the carrier recombination rate.

3.6. Photo-Electrochemical Activity of Co3O4 Modified TiO2 Nanorod Arrays

Methylene blue was used as the probe to evaluate the photo-electrochemical activity of pure TiO2 nanorod arrays and Co3O4 modified TiO2 nanorod arrays. As shown in Figure 7, the TiO2 nanorod array exhibits a moderate catalytic performance for the photo-electrochemical degradation of methylene blue, which could have contributed to the ordered arrays effect. The Co3O4 modified TiO2 nanorod arrays exhibit excellent photo-electrochemical degradation activity to methylene blue. TiO2 nanorod arrays modified with Co3O4 had good photochemical catalytic activity, which was attributed to the disjunction of Co3O4-TiO2 and the charge transfer between Co3O4 and TiO2 nanometer rod components. Furthermore, the isolated Co3O4 islands on the surface of TiO2 nanorod arrays acted as reactive sites to enhance the photo-electrochemical activity. The degradation of methylene blue by Co3O4 modified TiO2 was greater than that of pure TiO2 nanorod arrays. Methylene blue was totally degraded on Co3O4 modified TiO2 nanorod arrays. Compared with Co3O4-TiO2 nanorod arrays, only 60% methylene blue was degraded on pure TiO2 nanorod arrays.
The TiO2 nanorod arrays only absorbed ultraviolet photons to generate e-h+ pairs to degrade the methylene blue molecule, while Co3O4-TiO2 nanorod arrays provide a p-n junction between Co3O4 nanoparticles and TiO2 nanorod arrays. The enhancement of photo-electrochemical activity of Co3O4/TiO2 heterojunction samples could have contributed to the formation of the type-II p-n hetero-junction between Co3O4 and TiO2. TiO2 is an n-type wide band gap semiconductor with the conduction band at 0.14 V, and Co3O4 is a p-type narrow band gap with band energy of 2.07 eV [31]. When Co3O4 nanoparticles are deposited onto the surface of TiO2 nanorod arrays, a p-n heterojunction can be formed at the surface of TiO2 nanorod arrays, and the electrons can be transferred from the Co3O4 to TiO2 nanorod arrays until their Fermi levels are equal [32]. The equilibrium can be broken by methylene blue, which acted as a holes receptor. With the photo-electrochemical reaction going on, methylene blue was degraded. The positive potential was loaded onto the Co3O4-TiO2 electrode to further increase the efficiency of the separation of hole and electrons. The Co3O4 has Co2+ and Co3+ valence state. Co2+ and Co3+ can be easily oxidized to Co4+, and Co4+/Co3+ had high activity to oxidize methylene blue to water and carbon dioxide. Thus, Co3O4-TiO2 had higher photo-electrochemical activity than pure TiO2 nanorod arrays.
Hydroquinone was used to further study the photo-electrochemical activity of Co3O4 modified TiO2 and pure TiO2 nanorod arrays. It can be seen clearly from Figure 8a,b that, compared with pure TiO2, the Co3O4 modified TiO2 nanorod arrays had higher photo-electrochemical activity. Only 30% hydroquinone was degraded on pure TiO2 nanorod arrays, and hydroquinone displayed total degradation on Co3O4 modified TiO2 nanorod arrays.
In order to clearly measure the photo-electrochemical activity of Co3O4 modified TiO2 nanorod arrays and pure TiO2 nanorod arrays, the corresponding kinetic constant was computed through fitting the experimental degradation of hydroquinone by the following equation.
ln ( C t C 0 ) = k t
where Ct is the concentration of hydroquinone at a certain reaction time, C0 is the original concentration, k is the apparent first rate, and t is photo-electrochemical time. The model was suitable for the photo-electrochemical hydroquinone degradation process. The k-values for the Co3O4-TiO2 and pure TiO2 nanorod arrays are 0.91745 and 0.1206, respectively. The k-value of Co3O4-TiO2 is almost eight times that of the pure TiO2 nanorod arrays, which further confirmed that Co3O4 addition greatly enhanced the photo-electrochemical activity.

3.7. Photo-Electro-Catalytic Degradation Mechanism

The photo-electrochemical activity of the photo-catalyst was mainly determined by light absorption, charges holes separation, and charge transfer from the inner to the surface of catalyzer. The band gap of the semiconductor was the key factor, which had a great influence on the photo-activity of the catalyst. The bandgap of Co3O4-TiO2 nanorod arrays and pure TiO2 nanorod arrays was determined from UV/vis diffuse reflectance spectra using the Tauc function, as shown in Figure 9a. The bandgap of TiO2 nanorod arrays is 3.2 eV and is very close to that from the former literature [33]. The bandgap of Co3O4-TiO2 is 2.85 eV, which could have contributed to the intrinsic narrow band gap of Co3O4 hybrid with TiO2. The addition of Co3O4 reduced the band gap of the nanomaterial and then enhanced its catalytic activity.
The influence of Co3O4 on the energy level of the photo-catalyst was studied through the Mott–Schottky electrochemical method in 0.1 mol/L Na2SO4 with 1000 Hz for Co3O4-TiO2 and pure TiO2 nanorod arrays. Both TiO2 and Co3O4-TiO2 had positive slopes, meaning that both pure TiO2 and Co3O4-TiO2 are n-type semiconductors and electrons as the majority carriers. It can be seen that the addition of Co3O4 did not change the semiconductor type of TiO2, but greatly changed the Fermi level and the flat band potential of TiO2. This phenomenon is consistent with other reported work that the addition of cobalt would change the band energy of the nanomaterial [34]. The flat-band potential (Vb) and the carrier density of nano-materials can be calculated according to the following equation:
N D = 2 e ε 0 ε ( d E d ( 1 C 2 ) )
where e 0 is the dielectronic constant of the material, ε is the permittivity of the vacuum, e is the element charge, ND is the donor density, and C is the capacitance. From the slope in the plot of 1/C2 versus V in the Figure 9b, the smaller slope for the Co3O4-TiO2 reflects a higher electron donor density. The higher ND means lower resistance, faster charge transfer, and higher electrochemical activity.
The Co3O4-TiO2 nanorod arrays had greatly enhanced photo-electrochemical performance, which could be attributed to the synergetic effects of the formation of a p-n junction between Co3O4 and TiO2. The catalytic performance of semiconductor nanomaterials depends on the bandgap of the semiconductor nanomaterial, the separation of electrons and holes, and the lifetime of electrons and holes generated by photo exciting. The hole and electron can move the surface to react with the adsorbed reactant. However, the electron and hole could recombine easily in a short time, which greatly abates the activity of the catalyst. Therefore, the catalyst’s activity can be greatly influenced by the life-time of the photo-induced electron-holes pairs. The p-n junction could greatly enhance the life-time of the electron and holes. The longer life-time of holes and electrons greatly enhances the activity of catalysts. Co3O4 is a p-type semiconductor with a band gap of 2.19 eV [31], while TiO2 is an n-type semiconductor with a band gap of 3.2 eV [1]. Thus, Co3O4 participating in the TiO2 nanorod arrays could change the structure of TiO2 in three aspects: (1) broaden the absorption range from ultraviolet light to visible light; (2) form a p-n junction to enhance the life-time of the electron; or (3) the disjunct Co3O4 nanoparticles act as an activation point to improve the photo-electrochemical activity. The conduction band (CB) position of TiO2 is more anodic than Co3O4, so the excited electrons on the CB of TiO2 could not transfer to Co3O4, while the holes could transfer from TiO2 to Co3O4. The recombination of electron and hole could be reduced just as shown in Scheme 1. At the hetero-junction in thermal equilibrium, the p-type and n-type regions have completely opposite charges, and the n-type regions become positive, while the p-type region becomes negative. When the n-p hetero-junction semiconductor is excited by visible light with high energy to band gap, the photo-generated electrons can move to the CB of the n-type TiO2 and holes can move to the VB of the n-type semiconductor for the formation of the inner electric field in the Co3O4/TiO2 sample, which effectively impedes the recombination of electron–hole pairs. The biased voltage could further restrain the photo-excited electrons and holes recombination through the following mechanism. The positive bias voltage depletes the electrons and, as a result, the holes can be excluded to the surface. Then, the absorbed molecules on the surface of Co3O4 can react with the holes to form a series of radicals, such as OH radical and other radicals. These radicals have great energy to oxidize the organic waste to water and carbon dioxide.
Co 3 O 4 + visible   light Co 3 O 4 ( e ) + Co 3 O 4 ( h + )
TiO 2 + bias   voltage TiO 2 ( h + )
Co 3 O 4 + bias   voltage Co 3 O 4 + holes
H 2 O 2 + hole 2 OH
MB + 2 OH CO 2 + H 2 O
According to the above results and discussion, consequently, under external bias voltage, electrons transfer along the external wires to auxiliary electrode and leave holes on the surface of Co3O4, which could oxidize the organic waste to water and carbon dioxide, as the schematic shows.

4. Conclusions

Ultra small amounts of Co3O4-modified TiO2 nanorod arrays were successfully fabricated thorough green photochemical deposition methods without adding any nocuous reagents. The Co3O4/TiO2 nanorod arrays fabricated in acid solution had the highest photo-electrochemical activity. We elaborated on the mechanism of Co3O4-TiO2 fabricated in different pH value solutions. The amount of Co3O4 could be controlled by adjusting the concentration of Co ions. The small amount of Co3O4 made many disjunct active points, which acted as active sites to mineralize organic wastes during photoelectrochemical degradation. The Co3O4/TiO2 nanorod arrays had higher photo-electrochemical activity to degrade organic waste than pure TiO2 nanorod arrays, which had great potential in waste water treatment.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/11/1/214/s1, Figure S1: EDX of Co3O4 modified TiO2 nanorod arrays, Figure S2: Full XPS data of Co3O4 modified TiO2 nanorod arrays, Figure S3: OCP response of Co3O4 modified TiO2 nanorod arrays fabricated in different pH value. Table S1: the content of Co3O4 modified TiO2 nanorod arrays.

Author Contributions

Conceptualization, Z.Z. and C.W.; methodology, Y.D.; software, X.Z.; validation, Z.Z., Y.D. and C.W.; formal analysis, Y.D.; in-vestigation, W.C.; resources, Y.D.; data curation, C.L.; writing—original draft preparation, Y.D.; writing—review and editing, Z.Z.; visualization, C.L.; supervision, Z.Z.; project administration, Z.B.; funding acquisition, Z.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (51862029), Lanzhou Talent Innovation and Entrepreneurship Project (2019-RC-99), Industry Support and Guidance Project for Colleges and Universities in Gansu Province in 2020 (2020C-32), and Major Scientific Projects of Gan Su Institute of Political Science and Law (GZF2018XZD03).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Data is contained within the article or Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Naldoni, A.; Allieta, M.; Santangelo, S.; Marelli, M.; Fabbri, F.; Cappelli, S.; Bianchi, C.L.; Psaro, R.; Dal, V.S. Effect of nature and location of defects on bandgap narrowing in black TiO2 nanoparticles. J. Am. Chem. Soc. 2012, 134, 7600–7603. [Google Scholar] [CrossRef]
  2. Khan, S.U.; Shahry, M.; Ingler, W.B., Jr. Efficient photochemical water splitting by a chemically modified n-TiO2. Science 2002, 297, 2243–2245. [Google Scholar] [CrossRef] [PubMed]
  3. Neatu, S.; Macia, A.J.A.; Concepcion, P.; Garcia, H. Gold-copper nanoalloys supported on TiO2 as photocatalysts for CO2 reduction by water. J. Am. Chem. Soc. 2014, 136, 15969–15976. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, L.; Luo, S.; Li, Y.; Xiao, Y.; Kang, Q.; Cai, Q. High efficient photocatalytic degradation of p-nitrophenol on a unique Cu2O/TiO2 p-n heterojunction network catalyst. Environ. Sci. Technol. 2010, 44, 7641–7646. [Google Scholar] [CrossRef] [PubMed]
  5. Lu, W.; Zhang, Y.; Zhang, J.; Xu, P. Reduction of Gas CO2 to CO with High Selectivity by Ag Nanocube-Based Membrane Cathodes in a Photoelectrochemical System. Ind. Eng. Chem. Res. 2020, 59, 5536–5545. [Google Scholar] [CrossRef]
  6. Dhandole, L.K.; Mahadik, M.A.; Kim, S.G.; Chung, H.S.; Seo, Y.S.; Cho, M.; Ryu, J.H.; Jang, J.S. Boosting Photocatalytic Performance of Inactive Rutile TiO2 Nanorods under Solar Light Irradiation: Synergistic Effect of Acid Treatment and Metal Oxide Co-catalysts. ACS Appl. Mater. Interfaces 2017, 9, 23602–23613. [Google Scholar] [CrossRef]
  7. Huang, H.; Pan, L.; Lim, C.K.; Gong, H.; Guo, J.; Tse, M.S.; Tan, O.K. Hydrothermal Growth of TiO2 Nanorod Arrays and In Situ Conversion to Nanotube Arrays for Highly Efficient Quantum Dot-Sensitized Solar Cells. Small 2013, 9, 3153–3160. [Google Scholar] [CrossRef]
  8. Li, J.; Cushing, S.K.; Zheng, P.; Senty, T.; Meng, F.; Bristow, A.D.; Manivannan, A.; Wu, N. Solar hydrogen generation by a CdS-Au-TiO2 sandwich nanorod array enhanced with Au nanoparticle as electron relay and plasmonic photosensitizer. J. Am. Chem. Soc. 2014, 136, 8438–8449. [Google Scholar] [CrossRef]
  9. Liu, J.; Li, Y.; Ke, J.; Wang, S.; Wang, L.; Xiao, H. Black NiO-TiO2 nanorods for solar photocatalysis: Recognition of electronic structure and reaction mechanism. Appl. Catal. B Environ. 2018, 224, 705–714. [Google Scholar] [CrossRef]
  10. Mishra, S.; Yogi, P.; Sagdeo, P.R.; Kumar, R. TiO2–Co3O4 Core–Shell Nanorods: Bifunctional Role in Better Energy Storage and Electrochromism. ACS Appl. Energy Mater. 2018, 1, 790–798. [Google Scholar] [CrossRef]
  11. Yang, M.; Ding, B.; Lee, S.; Lee, J.K. Carrier Transport in Dye-Sensitized Solar Cells Using Single Crystalline TiO2 Nanorods Grown by a Microwave-Assisted Hydrothermal Reaction. J. Phys. Chem. C 2011, 115, 14534–14541. [Google Scholar] [CrossRef]
  12. Li, T.L.; Lee, Y.L.; Teng, H. CuInS2 quantum dots coated with CdS as high-performance sensitizers for TiO2 electrodes in photoelectrochemical cells. J. Mater. Chem. 2011, 21, 5089–5098. [Google Scholar] [CrossRef]
  13. Zhang, J.; Jin, X.; Morales, G.P.I.; Yu, X.; Liu, H.; Zhang, H.; Razzari, L.; Claverie, J.P. Engineering the Absorption and Field Enhancement Properties of Au-TiO2 Nanohybrids via Whispering Gallery Mode Resonances for Photocatalytic Water Splitting. ACS Nano 2016, 10, 4496–4503. [Google Scholar] [CrossRef] [PubMed]
  14. Biyoghe, L.; Ndong, B.; Ibondou, M.P.; Gu, X.; Lu, S.; Qiu, Z.; Sui, Q.; Mbadinga, S.M. Enhanced Photocatalytic Activity of TiO2 Nanosheets by Doping with Cu for Chlorinated Solvent Pollutants Degradation. Ind. Eng. Chem. Res. 2014, 53, 1368–1376. [Google Scholar] [CrossRef]
  15. Sun, T.; Fan, J.; Liu, E.; Liu, L.; Wang, Y.; Dai, H.; Yang, Y.; Hou, W.; Hu, X.; Jiang, Z. Fe and Ni co-doped TiO2 nanoparticles prepared by alcohol-thermal method: Application in hydrogen evolution by water splitting under visible light irradiation. Powder Technol. 2012, 228, 210–218. [Google Scholar] [CrossRef]
  16. Seh, Z.W.; Liu, S.; Low, M.; Zhang, S.Y.; Liu, Z.; Mlayah, A.; Han, M.Y. Janus Au-TiO2 photocatalysts with strong localization of plasmonic near-fields for efficient visible-light hydrogen generation. Adv. Mater. 2012, 24, 2310–2314. [Google Scholar] [CrossRef]
  17. Melvin, A.A.; Illath, K.; Das, T.; Raja, T.; Bhattacharyya, S.; Gopinath, C.S. M-Au/TiO2 (M = Ag, Pd and Pt) nanophotocatalyst for overall solar water splitting: Role of interfaces. Nanoscale 2015, 7, 13477–13488. [Google Scholar] [CrossRef]
  18. Bessekhouad, Y.; Robert, D.; Weber, J.V. Photocatalytic activity of Cu2O/TiO2, Bi2O3/TiO2 and ZnMn2O4/TiO2 heterojunctions. Catal. Today 2005, 101, 315–321. [Google Scholar] [CrossRef]
  19. Santamaria, M.; Conigliaro, G.; Franco, F.; Quarto, F. Photoelectrochemical Evidence of Cu2O/TiO2 Nanotubes Hetero-Junctions formation and their Physicochemical Characterization. Electrochim. Acta 2014, 144, 315–323. [Google Scholar] [CrossRef]
  20. Kupfer, B.; Majhi, K.; Keller, D.A.; Bouhadana, Y.; Rühle, S.; Barad, H.N.; Anderson, A.Y.; Zaban, A. Thin Film Co3O4/TiO2 Heterojunction Solar Cells. Adv. Energy Mater. 2015, 5, 1401007. [Google Scholar] [CrossRef]
  21. Liu, L.; Ji, Z.; Zou, W.; Gu, X.; Deng, Y.; Gao, F.; Tang, C.; Dong, L. In Situ Loading Transition Metal Oxide Clusters on TiO2 Nanosheets As Co-catalysts for Exceptional High Photoactivity. ACS Catal. 2013, 3, 2052–2061. [Google Scholar] [CrossRef]
  22. Sarkar, D.; Ghosh, C.K.; Mukherjee, S.; Chattopadhyay, K.K. Three dimensional Ag2O/TiO2 type-II (p-n) nanoheterojunctions for superior photocatalytic activity. ACS Appl. Mater. Interfaces 2013, 5, 331–337. [Google Scholar] [CrossRef] [PubMed]
  23. Shao, Z.; Zhang, Y.; Yang, X.; Zhong, M. Au-Mediated Charge Transfer Process of Ternary Cu2O/Au/TiO2-NAs Nanoheterostructures for Improved Photoelectrochemical Performance. ACS Omega 2020, 5, 7503–7518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wang, M.; Sun, L.; Lin, Z.; Cai, J.; Xie, K.; Lin, C. P–n heterojunction photoelectrodes composed of Cu2O-loaded TiO2 nanotube arrays with enhanced photoelectrochemical and photoelectrocatalytic activities. Energy Environ. Sci. 2013, 6, 1211–1220. [Google Scholar] [CrossRef]
  25. Zhang, G.; Huang, C.; Wang, X. Dispersing molecular cobalt in graphitic carbon nitride frameworks for photocatalytic water oxidation. Small 2015, 11, 1215–1221. [Google Scholar] [CrossRef]
  26. Downes, C.A.; Marinescu, S.C. Efficient Electrochemical and Photoelectrochemical H2 Production from Water by a Cobalt Dithiolene One-Dimensional Metal-Organic Surface. J. Am. Chem. Soc. 2015, 137, 13740–13743. [Google Scholar] [CrossRef]
  27. Gao, Z.; Zhang, L.; Ma, C.; Zhou, Q.; Tang, Y.; Tu, Z.; Yang, W.; Cui, L.; Li, Y. TiO2 decorated Co3O4 acicular nanotube arrays and its application as a non-enzymatic glucose sensor. Biosens. Bioelectron. 2016, 80, 511–518. [Google Scholar] [CrossRef]
  28. Cho, I.S.; Chen, Z.; Forman, A.J.; Kim, D.R.; Rao, P.M.; Jaramillo, T.F.; Zheng, X. Branched TiO(2) nanorods for photoelectrochemical hydrogen production. Nano Lett. 2011, 11, 4978–4984. [Google Scholar] [CrossRef]
  29. Steinmiller, E.M.; Choi, K.S. Photochemical deposition of cobalt-based oxygen evolving catalyst on a semiconductor photoanode for solar oxygen production. Proc. Natl. Acad. Sci. USA 2009, 106, 20633–20636. [Google Scholar] [CrossRef] [Green Version]
  30. Simon, T.; Bouchonville, N.; Berr, M.J.; Vaneski, A.; Adrovic, A.; Volbers, D.; Wyrwich, R.; Doblinger, M.; Susha, A.S.; Rogach, A.L.; et al. Redox shuttle mechanism enhances photocatalytic H2 generation on Ni-decorated CdS nanorods. Nat. Mater. 2014, 13, 1013–1018. [Google Scholar] [CrossRef]
  31. Chang, X.; Wang, T.; Zhang, P.; Zhang, J.; Li, A.; Gong, J. Enhanced Surface Reaction Kinetics and Charge Separation of p-n Heterojunction Co3O4/BiVO4 Photoanodes. J. Am. Chem. Soc. 2015, 137, 8356–8359. [Google Scholar] [CrossRef] [PubMed]
  32. Kibria, M.G.; Zhao, S.; Chowdhury, F.A.; Wang, Q.; Nguyen, H.P.; Trudeau, M.L.; Guo, H.; Mi, Z. Tuning the surface Fermi level on p-type gallium nitride nanowires for efficient overall water splitting. Nat. Commun. 2014, 5, 1–6. [Google Scholar] [CrossRef] [PubMed]
  33. Lee, J.S.; You, K.H.; Park, C.B. Highly photoactive, low bandgap TiO2 nanoparticles wrapped by graphene. Adv. Mater. 2012, 24, 1084–1088. [Google Scholar] [CrossRef] [PubMed]
  34. Sadanandam, G.; Lalitha, K.; Kumari, V.D.; Shankar, M.V.; Subrahmanyam, M. Cobalt doped TiO2: A stable and efficient photocatalyst for continuous hydrogen production from glycerol: Water mixtures under solar light irradiation. Int. J. Hydrogen Energy 2013, 38, 9655–9664. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of Co3O4-TiO2 nanorod arrays and pure TiO2 nanorod arrays.
Figure 1. XRD pattern of Co3O4-TiO2 nanorod arrays and pure TiO2 nanorod arrays.
Nanomaterials 11 00214 g001
Figure 2. Scanning electron microscopy (SEM) images of (a) TiO2 nanorod arrays and (b) Co3O4-TiO2 nanorod arrays.
Figure 2. Scanning electron microscopy (SEM) images of (a) TiO2 nanorod arrays and (b) Co3O4-TiO2 nanorod arrays.
Nanomaterials 11 00214 g002
Figure 3. TEM and high resolution TEM of TiO2 nanorod arrays (a,b) and Co3O4-TiO2 nanorod arrays (c,d).
Figure 3. TEM and high resolution TEM of TiO2 nanorod arrays (a,b) and Co3O4-TiO2 nanorod arrays (c,d).
Nanomaterials 11 00214 g003
Figure 4. The XPS spectra of Co3O4-TiO2 nanorod arrays: (a) O 1s, (b) Ti3d, and (c) Co2p.
Figure 4. The XPS spectra of Co3O4-TiO2 nanorod arrays: (a) O 1s, (b) Ti3d, and (c) Co2p.
Nanomaterials 11 00214 g004
Figure 5. The IT curves measured at applied potential of 0 V vs. saturated calomel electrode (SCE) under chopped illumination in 0.1 mol/L Na2SO4 solution: (a) TiO2 and Co3O4-TiO2 fabricated at pH 4.12; (b) Co3O4-TiO2 fabricated at different pH value.
Figure 5. The IT curves measured at applied potential of 0 V vs. saturated calomel electrode (SCE) under chopped illumination in 0.1 mol/L Na2SO4 solution: (a) TiO2 and Co3O4-TiO2 fabricated at pH 4.12; (b) Co3O4-TiO2 fabricated at different pH value.
Nanomaterials 11 00214 g005
Figure 6. Open circuit potential (OCP) curve of Co3O-TiO2 fabricated in different concentrations of Co ions.
Figure 6. Open circuit potential (OCP) curve of Co3O-TiO2 fabricated in different concentrations of Co ions.
Nanomaterials 11 00214 g006
Figure 7. Photo-electrochemical degradation of methylene blue on Co3O4-TiO2 and TiO2 nanorod arrays: (a) Co3O4-TiO2; (b) TiO2 nanorod arrays; and (c) full curve of methylene blue degradation on Co3O4-TiO2 and pure TiO2.
Figure 7. Photo-electrochemical degradation of methylene blue on Co3O4-TiO2 and TiO2 nanorod arrays: (a) Co3O4-TiO2; (b) TiO2 nanorod arrays; and (c) full curve of methylene blue degradation on Co3O4-TiO2 and pure TiO2.
Nanomaterials 11 00214 g007
Figure 8. (a) Hydroquinone degradation on Co3O4-TiO2 nanorod arrays; (b) hydroquinone degradation on pure TiO2 nanorod arrays; and (c) comparison of degradation rate of hydroquinone on Co3O4-TiO2 nanorod arrays and pure TiO2 nanorod arrays.
Figure 8. (a) Hydroquinone degradation on Co3O4-TiO2 nanorod arrays; (b) hydroquinone degradation on pure TiO2 nanorod arrays; and (c) comparison of degradation rate of hydroquinone on Co3O4-TiO2 nanorod arrays and pure TiO2 nanorod arrays.
Nanomaterials 11 00214 g008
Figure 9. (a) Plots of (ahv)2 vs. photo energy of Co3O4-TiO2 and TiO2 nanorod arrays. (b) Mott–Schottky plots of Co3O4-TiO2 and pure TiO2 nanorod arrays in 0.5 mol/L Na2SO4 at a frequency of 1 KHz.
Figure 9. (a) Plots of (ahv)2 vs. photo energy of Co3O4-TiO2 and TiO2 nanorod arrays. (b) Mott–Schottky plots of Co3O4-TiO2 and pure TiO2 nanorod arrays in 0.5 mol/L Na2SO4 at a frequency of 1 KHz.
Nanomaterials 11 00214 g009
Scheme 1. Schematic mechanism of the photo-electrochemical degradation of organic pollution on Co3O4-TiO2 at a constant positive bias potential.
Scheme 1. Schematic mechanism of the photo-electrochemical degradation of organic pollution on Co3O4-TiO2 at a constant positive bias potential.
Nanomaterials 11 00214 sch001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Du, Y.; Zheng, Z.; Chang, W.; Liu, C.; Bai, Z.; Zhao, X.; Wang, C. Trace Amounts of Co3O4 Nano-Particles Modified TiO2 Nanorod Arrays for Boosted Photoelectrocatalytic Removal of Organic Pollutants in Water. Nanomaterials 2021, 11, 214. https://doi.org/10.3390/nano11010214

AMA Style

Du Y, Zheng Z, Chang W, Liu C, Bai Z, Zhao X, Wang C. Trace Amounts of Co3O4 Nano-Particles Modified TiO2 Nanorod Arrays for Boosted Photoelectrocatalytic Removal of Organic Pollutants in Water. Nanomaterials. 2021; 11(1):214. https://doi.org/10.3390/nano11010214

Chicago/Turabian Style

Du, Yongling, Zhixiang Zheng, Wenzhuo Chang, Chunyan Liu, Zhiyong Bai, Xinyin Zhao, and Chunming Wang. 2021. "Trace Amounts of Co3O4 Nano-Particles Modified TiO2 Nanorod Arrays for Boosted Photoelectrocatalytic Removal of Organic Pollutants in Water" Nanomaterials 11, no. 1: 214. https://doi.org/10.3390/nano11010214

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