Next Article in Journal
Thermal Energy Transfer between Helium Gas and Graphene Surface According to Molecular Dynamics Simulations and the Monte Carlo Method
Previous Article in Journal
How Hydrodynamic Phonon Transport Determines the Convergence of Thermal Conductivity in Two-Dimensional Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Co3O4 Nanopetals Grown on the Porous CuO Network for the Photocatalytic Degradation

1
School of Materials Science and Engineering, Hebei University of Technology, Tianjin 300401, China
2
Key Laboratory for New Type of Functional Materials in Hebei Province, Hebei University of Technology, Tianjin 300401, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(16), 2850; https://doi.org/10.3390/nano12162850
Submission received: 4 July 2022 / Revised: 11 August 2022 / Accepted: 16 August 2022 / Published: 18 August 2022

Abstract

:
Designing a novel photocatalytic composite for the efficient degradation of organic dyes remains a serious challenge. Herein, the multi-layered Co3O4@NP-CuO photocatalyst with unique features, i.e., the self-supporting, hierarchical porous network as well as the construction of heterojunction between Co3O4 and CuO, are synthesized by dealloying-electrodeposition and subsequent thermal treatment techniques. It is found that the interwoven ultrathin Co3O4 nanopetals evenly grow on the nanoporous CuO network (Co3O4@NP-CuO). The three-dimensional (3D) hierarchical porous structure for the catalyst provides more surface area to act as active sites and facilitates the absorption of visible light in the photodegradation reaction. Compared with the commercial CuO and Co3O4 powders, the newly designed Co3O4@NP-CuO composite exhibits superior photodegradation performance for RhB. The enhanced performance is mainly due to the construction of heterojunction of Co3O4/CuO, greatly promoting the efficient carrier separation for photocatalysis. Furthermore, the possible photocatalytic mechanism is analyzed in detail. This work provides a promising strategy for the fabrication of a new controllable heterojunction to improve photocatalytic activity.

1. Introduction

Water is an essential natural resource for life, inseparable from human health and the sustainability of economic development. Due to the rapid development of industrialization, most of the freshwater resources have been consumed, resulting in a large amount of wastewater contaminated with organic compounds, heavy metal ions, and dyes, which has a serious impact on human production and life. Dye wastewater (from hair dyes, leather, textiles, and solar luminescence technology), one of the most common pollutants, is highly toxic and carcinogenic [1,2,3,4,5,6,7]. Nowadays, a series of strategies are adopted to address environmental issues, such as biodegradation technology [8], physical adsorption [9], and advanced oxidation processes (AOPs) [10,11,12]. Especially, the AOPs, with the generation of strongly oxidizing hydroxyl radicals (·OH) and superoxide radicals (·O2) in the catalytic reactions, have attracted considerable attention due to the ability of rapid organic decomposition and efficient wastewater purification. Photocatalytic semiconductor degradation, one of the AOPs, has many advantages, including no secondary pollution, low expenses, and complete oxidation of organic dyes, etc. [13,14].
As far, a p-type semiconductor Co3O4 (the band gap of ~1.7–2.3 eV) [15,16,17], with the different morphologies of nanoparticles [18], nanosheets [19], and quantum dots [20], has been utilized in electrochemical sensors, electrodes, and magnetic materials, respectively. Considering the theoretical good visible light response range of Co3O4 [21], it is promising in the field of photocatalytic degradation. Hence, it is of great importance to study the photocatalytic degradation properties of Co3O4.
Unfortunately, numerous semiconductors (including Co3O4) [22,23,24,25] encounter a major problem of narrow optical response range, which results in high photogenerated-carriers recombination rate. Among the variety of modifications [26,27], the construction of heterojunction/Schottky junction by epitaxial growth can efficiently facilitate spatial charge separation and further improve the photocatalytic performance, as exemplified for Co3O4/TiO2 [28], CuxO/ZnO@Au [29] and Co3O4/g-C3N4 [30]. Another issue is how to reduce the agglomeration of powders-based composite photocatalysts obtained by conventional preparations [31,32]. To better solve the above-mentioned structural and functional drawbacks, it is necessary to design a new type of photocatalyst simultaneously possessing a self-supporting structure and heterojunction or Schottky junction structure.
Recently, nanoporous metals (NPMs) and metallic composites have remarkable features in the field of catalysis on account of fertile active sites exposed in the unique three-dimensional (3D) nanoporous structure [33,34,35,36,37,38,39], attracting much attention in the photocatalysis field. For instance, Wang et al. [40] reported a high degradable activity of 3D nanostructure rutile TiO2 photocatalyst by dealloying Cu60Ti30Y10 amorphous in HNO3 aqueous solutions. In our previous work, Cu2O/CuO nanowires/nanosheets [41,42] and Au nanoparticles modified CuO nanowires grew on nanoporous Cu (NPC) [43] have been successfully synthesized via a simple dealloying-anodizing technique and the composites exhibited much-improved degradation performance.
The work in this paper synthesized a novel Co3O4@CuO composite catalyst by utilizing the dealloying Cu40Zr60 amorphous alloy followed by the electrodeposition methods [44] along with heat treatment. The Co3O4@CuO composite exhibits an intriguing multi-layered structure and concurrently possesses the self-supporting, hierarchical porous network as well as the Co3O4/CuO heterojunction, which not only avoids the agglomeration of powder-like catalysts but also accelerates the transport of photogenerated carriers and reduces the recombination of electrons (e) and holes (h+). The photodegradable performance in rhodamine b (RhB) was further measured. Moreover, the underlying degradation mechanism in the photocatalytic process was analyzed in detail by using active radicals capture experiments.

2. Materials and Methods

2.1. Synthesis of Nanoporous Cu (NPC)

Alloy ingots of Cu40Zr60 (at. %) were prepared by arc-melting a mixture of pure Cu (99.99 wt%) and Zr (99.99 wt%) under a Zr-gettered argon atmosphere. The Cu40Zr60 amorphous ribbons with a width of 2 mm and a thickness of ~25 μm were fabricated by remelting and quenching onto a rotating copper wheel at a linear speed of 34 m/s [45]. Then, the nanoporous Cu (NPC) was fabricated by free-chemical dealloying the as-spun Cu40Zr60 ribbons immersed in 0.05 M HF solution for 2 h at 298 K [43]. The as-dealloyed ribbons (NPC) were cleaned with deionized water and dried at 333 K in a vacuum.

2.2. Synthesis of Co3O4@NP-CuO Composite

The as-dealloyed ribbons (NPC, with a length of 35 mm) were directly used as the deposition substrate network. Then, the Co(OH)2 nanopetals were electrodeposited on the NPC substrate network (Co(OH)2@NPC) in the 60 mM Co(NO3)2·6H2O via the electrodeposition method [44], where NPC served as the working electrode, a platinum network as a counter electrode, and Ag/AgCl as a reference electrode. The electrochemical deposition was performed at −0.8 V for 30 min. The as-calcined ribbons (Co3O4@NP-CuO composite) were further obtained by calcining the as-deposited ribbons (Co(OH)2@NPC) at 573 K for 2 h in the air.

2.3. Microstructure Characterization

The phase and crystal structure of the prepared samples were characterized by an X-ray diffractometer (XRD, D8, Cu-Kα, Bruker, Karlsruhe, Germany). The chemical valences of elements were detected using an X-ray photoelectron spectrometer (XPS, Thermo Fisher Scientific, Waltham, MA, USA). Transmission electron microscopy (TEM, JEOL JEM 2100F, Tokyo, Japan) and scanning electron microscopy (SEM, Nova nanoSEM 450, FEI, Hillsboro, OR, USA) were applied to characterize the microstructure of the samples. The photoluminescence spectra (PL) were detected on a fluorescence spectrometer (FL3-22). The ASAP 2020M+C type-specific surface area and porosity analyzer were used to measure specific surface area and the pore diameter distribution of the samples via BET and BJH methods, respectively.

2.4. Degradation Experiments

The simultaneous degradation experiments were carried out under the irradiation of a xenon lamp light source (100 mW/cm2, λ ≥ 420 nm) to study the degradation performance of photocatalysts. During the experiments, the catalyst with an effective catalytic area of 1 cm2 (ca 2 mg) was added into the mixed 6 mL of 10 mg/L RhB solution and 2 mL of 30 mass% H2O2. The concentration of RhB was measured on a UV-vis spectrophotometer (UV-1920) every 2 min. Tert-butanol (TBA), p-benzoquinone (PBQ), silver nitrate (AgNO3), and disodium ethylenediaminetetraacetate (EDTA-2Na) with a consistent amount (2 mL of 1 mM) were added to the above solutions to capture hydroxyl radicals (·OH), superoxide radicals (·O2), electrons (e) and holes (h+), respectively to identify the essential active species. The active species were further determined by the electron paramagnetic resonance (EPR) using Bruker A300 electron paramagnetic resonance spectrometer.

3. Results and Discussion

3.1. Design of Self-Supporting Co3O4@NP-CuO Composite

Figure 1a shows the SEM image of the as-dealloyed ribbon (NPC) obtained by dealloying Cu40Zr60 ribbons in 0.05 M HF solutions for 2 h. It is clearly seen that the as-dealloyed ribbon exhibits a 3D bicontinuous pore/ligament structure [43]. The as-dealloyed ribbon is chosen as the deposition substrate network. First, to explore the effect of deposition potential on the surface morphology grown on the as-dealloyed ribbon substrate, Figure S1 shows that SEM images of the as-deposited ribbons after electrodeposition in 60 mM cobalt nitrate solutions, at applied potentials of −0.6, −0.8, −1.0, and −1.2 V for 30 min, respectively. Unlike the uneven distribution of nanopetals (Figure S1a) and the coarsen nanopetals (Figure S1c,d), it is observed that the most uniform and fine nanopetals grow on the NPC substrate (Figure S1b) when the deposited voltage is at −0.8 V. Thus, the electrodeposition voltage and time of −0.8 V and 30 min, respectively, are selected as the following electrodeposition conditions.
After depositing (Figure S1b), the as-deposited samples are further calcined at 573 K for 2 h, shown in Figure 1b. In Figure 1b, it is clearly observed that the interwoven nanopetals uniformly grow on the as-calcined sample surface [44], which is similar to those of as-deposited samples before heat treatment.
Figure 1c displays the cross-sectional SEM image of the as-calcined ribbon. It can be clearly observed that the as-calcined ribbon exhibits a unique three-layer structure in which the interfaces are in close contact with each other [44]. Their corresponding EDS spectra are shown in Figure S2. For the nanopetals deposition layer (region A), the EDS spectrum (Figure S2a) mainly shows the peaks of Co and O with an atomic ratio of ~3:4, implying the formation of the Co3O4 nanopetals layer. The nanoporous layer (region B) is composed of Cu and O elements in Figure S2b, corresponding to the nanoporous CuO layer (NP-CuO). It should be noticed that, compared with the morphology of NPC (Figure 1a), the ligaments of nanoporous layer (the inset image of Figure 1c) become coarsened, which is caused by the calcination treatment and the oxidation of Cu substrate network, but the CuO inherits the 3D bicontinuous nanoporous structure of NPC. Furthermore, the EDS spectrum of the inter-region C (Figure S2c) consists mainly of Cu and Zr elements with an atomic ratio of ~4:6, indicating that the region belongs to the Cu40Zr60 amorphous layer. Based on the above results, it can be concluded that the Co3O4@NP-CuO composite sample, synthesized via the dealloying and electrodeposition method followed by the calcination, exhibits a multilayer structure. In addition, uniform nanoporous structures of the NPC substrate network (Figure S3a) and uniform growth of nanopetals on the NPC (Figure S3b) can be observed at a larger scale range, which guarantees the homogeneity of Co3O4/NP-CuO composite samples. Moreover, the entire composite material keeps a free-standing structure, which is attributed to the amorphous alloy interlayer support [46,47].
The nitrogen adsorption-desorption isotherms and pores size distribution curves of Co3O4@NP-CuO are shown in Figure 1d. It belongs to the type IV isotherm curve [43], indicating that the Co3O4@NP-CuO sample possesses a mesoporous structure [44,48]. The specific surface area of Co3O4@NP-CuO is about 49.58 m2/g. The pores size of the material (the inset of Figure 1d) is mainly in the range of 5–15 nm, which derives from the pores on the NP-CuO matrix and the pores formed by the interweaving of nanopetals in the Co3O4 layer.
The XRD patterns of all samples are shown in Figure 2. The as-spun Cu40Zr60 ribbons have a broad diffraction peak (Figure 2a), showing the formation of an amorphous structure [43,49]. After dealloying, the fcc Cu (JCPDS #04-0836) crystalline peaks (Figure 2a) are indexed in the XRD pattern of as-dealloyed ribbons, and no Zr element is detected, implying that Zr element selectively dissolves into the HF solution and the formation of nanoporous Cu (NPC) shown in Figure 1a [43]. On the other hand, the NPC substrate is selected for electrodeposition at −0.8 V for 30 min and calcination treatment at 573 K for 2 h in air. Figure 2b shows the XRD patterns of as-deposited and as-calcined ribbons. In addition to the metallic Cu peaks, the as-deposited ribbons (without calcination) show new characteristic diffraction peaks at 19.1°, 32.5°, 38.0°, and 51.5°, which are indexed as the (001), (100), (011), and (012) planes of Co(OH)2 (JCPDS #74-1057), respectively [44]. After calcination, the crystalline phases of the as-calcined ribbons are identified to be CuO (JCPDS #45-0937) and Co3O4 (JCPDS #42-1467), revealing the formation of nanoporous CuO (NP-CuO) on the NPC and Co(OH)2 nanopetals change to be Co3O4 nanopetals [50]. The results can be verified in the SEM image in Figure 1c and the EDS spectra of Figure S2.
As for the as-calcined ribbon exhibiting the multi-layer structure (Figure 1c), it is important to make clear the surface composition and chemical states of either the nanopetal layer or nanoporous layer of the as-calcined ribbon. Figure 3 shows XPS analysis applied to the nanopetal layer or nanoporous layer. From Figure 3a, the strong photoelectron peaks of Co appear in the nanopetals layer, whereas the Cu peaks significantly reduce as compared to those in the nanoporous layer. Figure 3b shows the deconvolution of the Co 2p peaks obtained from the nanopetals layer. The binding energy peaks of the Co 2p spectrum at 779.9 and 795.0 eV are related to Co 2p3/2 and Co 2p1/2 of Co2+ with shakeup satellite peaks at 789.4 and 804.7 eV, respectively [50]. Additionally, the two peaks centered at 781.2 and 796.4 eV are assigned to the characteristic Co 2p3/2 and Co 2p1/2 of Co3+, revealing the coexistence of Co2+ and Co3+ species in the Co3O4@NP-CuO [51]. Moreover, the peak intensity of Cu 2p is not obviously seen in Figure S4, which might attribute to the thicker deposited nanopetals layer. On the other hand, for the nanoporous layer just beneath the nanopetals layer, the XPS survey spectrum of the nanoporous layer (Figure 3a) identifies the Cu and O elements. In Figure 3c, the peaks of Cu 2p spectra located around 934.0 and 953.7 eV correspond to Cu 2p3/2 and Cu 2p1/2 of Cu2+, respectively [52]. Moreover, the Zr element is not detected in the nanoporous layer, inferring that Zr has been selectively etched after dealloying the Cu40Zr60 amorphous ribbon [43]. For both the layers displayed in Figure 3d, the two peaks of O 1s at 529.9 and 531.2 eV correspond to the metal oxides (OM) and absorbed oxygen (OH), respectively [53,54,55], demonstrating that the nanopetals layer is mainly Co3O4 and the nanoporous layer is CuO.
More morphologies and structure characterizations of the as-calcined multi-layered ribbon are further detected by TEM (Figure 4 and Figure 5). TEM images of the nanopetals layer for the Co3O4@NP-CuO composite are shown in Figure 4. Figure 4a shows 3D interwoven ultrathin nanopetals in the Co3O4 nanopetals layer. The SAED pattern in Figure 4b shows the (111), (220), (311), and (400) crystal orientation index, confirming the presence of the Co3O4 phase [56]. Meanwhile, from the HRTEM of nanopetals shown in Figure 4c, the lattice fringes with a spacing of 0.24 nm, 0.28 nm, and 0.47 nm are assigned to (311), (220), and (111) planes of Co3O4 (JCPDS #42-1467), respectively [51,56]. The elemental distribution of the Co3O4 nanopetals is further measured by EDS mapping, as depicted in Figure 4d–g. From the dark field image of the nanopetals shown in Figure 4d, it is clearly observed that the interwoven and superimposed structure of Co3O4 nanopetals. Obviously, Co and O elements uniformly accumulate in the Co3O4 nanopetals, revealing that the nanopetals are mainly composed of Co and O elements.
The nanopetals layer of the as-calcined ribbon is confirmed to be mainly Co3O4, while the CuO nanoporous layer needs to be more characterized. Figure 5a,b show the SEM and TEM images of the CuO nanoporous layer, respectively. It is observed that ligaments after calcination become coarsened, and the pore features are inherited. From the HRTEM of the nanoporous layer (Figure 5c), it is found that the lattice fringes spacing of 0.252 and 0.232 nm are the (11-1) and (111) of CuO, respectively [43,57]. The SAED pattern (Figure 5d) further proves the existence of the CuO phase. Moreover, the elemental mapping of the nanoporous layer is shown in Figure 5e,f. The Cu and O elements evenly distribute within the 3D ligaments, confirming the NPC is oxidized to be NP-CuO. Hence, the multi-layer porous Co3O4@NP-CuO composite with a 3D bicontinuous pore/ligament skeleton structure, self-supporting as well as the Co3O4/CuO heterojunction is successfully synthesized via the dealloying and electrodeposition methods followed by the calcination.

3.2. Photocatalytic Degradation Activity of Co3O4@NP-CuO Composite

The photocatalytic activity of the Co3O4@NP-CuO composite was evaluated by degrading RhB at room temperature. Figure 6a displays photocatalytic degradation results of Co3O4@NP-CuO composite in a mixed solution of RhB and H2O2 under illumination. The strong characteristic adsorption peak of RhB at 553 nm continuously decreases with time and becomes very weak after 11 min. From the inset of Figure 6a, the color change process of the RhB dye is shown, eventually becoming transparent. Figure 6b demonstrates the different degradation efficiencies of RhB by the commercial CuO, Co3O4 powders, and the composite. Compared to the commercial CuO and Co3O4, the newly designed Co3O4@NP-CuO composite exhibits an outstanding photocatalytic activity towards RhB decomposition due to the synergistic effect of the porous NP-CuO and the ultrathin Co3O4 nanopetals.
In order to explore the influence of catalyst, light-irradiation, and oxidizing agent H2O2 on the degradation performance, Figure 6c shows the performance graphs of RhB degradation in different conditions. The RhB is partially reduced in the absence of H2O2, indicating that H2O2 acts an auxiliary role in the photocatalytic process. It is worth noting that RhB is only degraded by 15–20% without a catalyst or light-irradiation. Notably, the degradation performance is significantly enhanced when all the Co3O4@NP-CuO catalyst, H2O2 and light-irradiation are applied to the RhB reaction system. The results indicate that the catalyst, H2O2, and light-irradiation are indispensable for the rapid degradation of RhB. The synergistic effect of the catalyst, irradiation, and H2O2 will be proposed in the degradation mechanism. On the other hand, the photocatalytic degradation efficiencies, and recycling performance of RhB over the Co3O4@NP-CuO were further measured. The Co3O4@NP-CuO composite maintains a high RhB degradation rate of over 85% within seven cycles, demonstrating high recyclability.
The kinetic model of the photodegradation process can be well evaluated by the photodegradation rate and the activation energy of the photodegradation reaction possesses. Figure 7 clearly shows the photocatalytic degradation rate and activation energy of the Co3O4@NP-CuO composite. The performance curves of Co3O4@NP-CuO composite for RhB degradation at different temperatures are shown in Figure 7a. As the temperature of the degradation process increases from 298 K to 318 K, the degradation rate of RhB increases continuously. The degradation rate of RhB follows the pseudo-first-order kinetic model [41,42,43], as shown in Equation (1):
ln ( C t C 0 ) = kt
where C0 and Ct correspond to the transient concentration of at time 0 and at time t, respectively, k is the reaction rate constant of the photocatalytic degradation process. Combining Figure 7a with Equation (1), −ln(Ct/C0) versus t for RhB photodegradation at different temperatures is drawn and fitted in Figure 7b. At different temperatures, −ln(Ct/C0) shows a good linear relationship with t. The calculated photocatalytic degradation rates k at 298, 308, and 318 K are 0.25, 0.42, and 0.60 min−1, respectively. Furthermore, the activation energy of the photocatalytic degradation process can be evaluated according to the Arrhenius Equation (2):
lnk = E a RT + lnA
In the formula, k is the photocatalytic rate constant at different temperatures, Ea is the apparent activation energy, R is the ideal molar gas constant, T is the thermodynamic temperature, and A is the frequency factor. The curve in Figure 7c is established with −lnk as the ordinate and 1000/RT as the abscissa. There is a good linear relationship between −lnk and 1000/RT. According to Equation (2), the activation energy of Co3O4@NP-CuO is calculated to be ~34.7 kJ/mol for the RhB degradation reaction from 298 to 318 K. The lower activation energy reflects that the photocatalytic degradation process is easier to carry out, which also proves the photocatalytic performance of the Co3O4@NP-CuO composite.

3.3. Degradation Mechanism

It is well-known that the recombination of photogenerated e and h+ for the semiconductor catalysts greatly influences the degradation efficiency. The recombination of photogenerated e and h+ can release energy in the form of luminescence and exotherm [41,42]. Therefore, the photoluminescence (PL) spectrum of CuO, Co3O4, and Co3O4@NP-CuO composite can be utilized to evaluate photogenerated charges recombination behavior. Figure 8a shows the PL spectra of Co3O4@NP-CuO composite, CuO and Co3O4. They exhibit different fluorescence intensities around λ = 610 nm. Compared with the fluorescence intensities of CuO and Co3O4, the fluorescence intensity of the Co3O4@NP-CuO composite is significantly lower, indicating that the Co3O4/CuO heterostructure is beneficial in reducing the recombination rate of photogenerated carriers [15,57].
In the photocatalytic degradation process, active radicals always play a significant role. To unveil the photocatalytic degradation mechanism of Co3O4@NP-CuO composite, radical trapping experiments were carried out (Figure 8b). During the photocatalytic reaction, TBA, PBQ, AgNO3, and ethyl EDTA-2Na were added to the RhB solution as trapping agents for OH, O2−, e, and h+, respectively [43]. When the TBA is introduced, the degradation rate of RhB is apparently limited, and the degradation efficiency of the Co3O4@NP-CuO composite decreases from 92% to 40%, demonstrating that ·OH is the dominant species in the degradation of RhB. After the addition of AgNO3, the degradation rate of RhB is limited to a certain extent, and the degradation rate decreases from 92% to 63%, implying that e is also the main active species in the photocatalytic degradation process. On the contrary, the degradation rate changes slightly when PBQ and EDTA-2Na are added, inferring that ·O2− and h+ have a minor effect on the rapid degradation reaction of RhB. Based on the above analysis, ·OH and e are the main active species in the process of RhB degradation for the present catalyst.
The electron paramagnetic resonance (EPR) technology was implemented to further disclose the generation process of free radicals (Figure 8c). It is found that a small amount of ·OH is generated in the dark, mainly due to the Fenton-like reaction of CuO and Co3O4 such as Equations (3)–(6) [41,43,58,59,60]:
Cu 2 + + H 2 O 2     Cu + + HOO · + H +
Cu + + H 2 O 2     Cu 2 + + · OH + OH
Co 2 + + H 2 O 2     Co 3 + + · OH + OH
Co 3 + + H 2 O 2     Co 2 + + HOO · + H +
The production of ·OH is greatly improved after the addition of H2O2, indicating that H2O2 is essential in the production of ·OH. Besides the Fenton-like reaction, it should be noted that the role of e- cannot be ignored because H2O2 acts as an acceptor of e- and react to generate ·OH (Equation (7)) [43]:
e + H 2 O 2 OH + · OH
Combined with the above results and analysis, the photocatalytic degradation mechanism of Co3O4@NP-CuO is proposed. The degradation of RhB includes the Fenton-like reaction (Equations (3)–(6)) and photocatalytic reaction (Equation (7)). As shown in Figure 6c, the photocatalytic reaction dominates the degradation process. The efficient degradation is attributed to the synergistic effect of light-irradiation and H2O2. Light and H2O2 are essential to produce ·OH [43]. Based on Equation (7), the large-scale generation of ·OH is mainly due to the reaction between photogenerated electrons produced by light-irradiation and H2O2 in Co3O4@NP-CuO. On the other hand, the 3D bicontinuous pore/ligament skeleton structure and Co3O4/CuO heterojunction together contribute to promoting photocatalytic performance [41,42,43]. The abundant Co3O4 nanopetals can absorb more light and generate more photogenerated e and h+. The 3D mesoporous structure of NP-CuO accelerates the transfer of photogenerated carriers, thereby effectively reducing the undesirable recombination of e and h+ [48]. In addition, the exposed surface of Co3O4@NP-CuO can provide more active sites for the improvement of photocatalytic degradation efficiency. Overall, the novel Co3O4@NP-CuO composite with a 3D bicontinuous pore/ligament skeleton structure exhibits excellent performance, opening a new avenue for designing photocatalytic materials.

4. Conclusions

In this work, the multi-layered homogenous Co3O4@NP-CuO composite simultaneously exhibiting the self-supporting, porous network as well as the Co3O4/CuO heterojunction is successfully achieved via the dealloying-electrodeposition methods followed by the calcination. The interwoven “petal-like” Co3O4 ultrathin nanosheets uniformly grow on the nanoporous CuO substrate network. The present Co3O4@NP-CuO composite exhibits much higher photodegradable performance for RhB as compared to the commercial CuO and Co3O4 powders. The rich porosity of multi-layered homogenous composite catalyst facilities the absorption of visible light and transportation of RhB dyes to generate more electrons (e) and ·OH. Furthermore, the construction of heterojunction between Co3O4 and CuO effectively promotes the separation of photogenerated carriers, which is attributed to the higher performance under the synergistic effect of H2O2.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12162850/s1, Figure S1: SEM images of deposition at a voltage of −0.6 V (a), −0.8 V (b), −1.0 V (c), and −1.2 V (d); Figure S2: EDS analysis for the cross-sectional image of as-calcined ribbon: nanopetals layer (a), nanoporous layer (b), amorphous layer (c); Figure S3: SEM images of a wide range of Cu40Zr60 amorphous alloy ribbon dealloyed in 0.05 M HF for 2 h (a) and the composite sample via deposition at −0.8 V for 30 min followed by the calcination (b); Figure S4: XPS spectrum of Cu 2p for nanopetals layer.

Author Contributions

Conceptualization, C.Q.; methodology, Y.S., C.W. and S.Q.; software, Y.S., C.W. and S.Q.; validation, C.Q.; formal analysis, C.Q. and F.P.; investigation, Y.S., C.W., S.Q. and F.P.; resources, C.Q.; data curation, Y.S. and C.W.; writing—original draft preparation, Y.S.; writing—review and editing, C.Q. and F.P.; visualization, Z.W. and Y.L.; supervision, C.Q. and F.P.; project administration, C.Q.; funding acquisition, Y.L., Z.W. and C.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financially supported by the National Natural Science Foundation of China (52071125), Natural Science Foundation of Hebei Province, China (E2020202176; E2020202071), Science and Technology Project of Hebei Education Department, China (ZD2020177).

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.

References

  1. Li, S.J.; Xue, B.; Wu, G.Y.; Liu, Y.P.; Zhang, H.Q.; Ma, D.Y.; Zuo, J.C. A Novel Flower-Like Ag/AgCl/BiOCOOH Ternary Heterojunction Photocatalyst: Facile Construction and Its Superior Photocatalytic Performance for the Removal of Toxic Pollutants. Nanomaterials 2019, 9, 1562. [Google Scholar] [CrossRef] [PubMed]
  2. Uthirakumar, P.; Devendiran, M.; Kuznetsov, A.Y.; Kim, G.C.; Lee, I.H. Efficient, recyclable, and affordable daylight induced Cu/Cu2O/CuI photocatalyst via an inexpensive iodine sublimation process. Appl. Surf. Sci. 2021, 537, 147007. [Google Scholar] [CrossRef]
  3. Yuan, X.Q.; Pei, F.; Luo, X.L.; Hu, H.T.; Qian, H.M.; Wen, P.; Miao, K.K.; Guo, S.F.; Wang, W.; Feng, G.D. Fabrication of ZnO/Au@Cu2O heterojunction towards deeply oxidative photodegradation of organic dyes. Sep. Purif. Technol. 2021, 262, 118301. [Google Scholar] [CrossRef]
  4. Qiang, T.T.; Chen, L.; Xia, Y.J.; Qin, X.T. Dual modified MoS2/SnS2 photocatalyst with Z-scheme heterojunction and vacancies defects to achieve a superior performance in Cr (VI) reduction and dyes degradation. J. Clean. Prod. 2021, 291, 125213. [Google Scholar] [CrossRef]
  5. Nohynek, G.; Hueber-Becker, F.; Meulingy, W.; Dufour, E.; Bolt, H.; Debie, A. Occupational exposure of hairdressers to [C-14]-para-phenylenediamine-containing oxidative hair dyes. Toxicol. Lett. 2007, 172, S30–S31. [Google Scholar] [CrossRef]
  6. Cassano, A.; Molinari, R.; Romano, M.; Drioli, E. Treatment of aqueous effluents of the leather industry by membrane processes A review. J. Membr. Sci. 2001, 181, 111–126. [Google Scholar] [CrossRef]
  7. Albano, G.; Colli, T.; Biver, T.; Aronica, L.A.; Pucci, A. Photophysical properties of new p-phenylene- and benzodithio-phene-based fluorophores for luminescent solar concentrators (LSCs). Dyes Pigm. 2020, 178, 108368. [Google Scholar] [CrossRef]
  8. Li, Y.; Li, L.; Chen, Z.X.; Zhang, J.; Gong, L.; Wang, Y.X.; Zhao, H.Q.; Mu, Y. Carbonate-activated hydrogen peroxide oxidation process for azo dye decolorization: Process, kinetics, and mechanisms. Chemosphere 2018, 192, 372–378. [Google Scholar] [CrossRef]
  9. Sakkayawong, N.; Thiravetyan, P.; Nakbanpote, W. Adsorption mechanism of synthetic reactive dye wastewater by chitosan. J. Colloid Interface Sci. 2005, 286, 36–42. [Google Scholar] [CrossRef]
  10. Liu, L.M.; Chen, Z.; Zhang, J.W.; Shan, D.; Wu, Y.; Bai, L.M.; Wang, B.Q. Treatment of industrial dye wastewater and pharmaceutical residue wastewater by advanced oxidation processes and its combination with nanocatalysts: A review. J. Water Process Eng. 2021, 42, 102122. [Google Scholar] [CrossRef]
  11. Li, L.; Han, Q.; Wang, L.; Liu, B.; Wang, K.K.; Wang, Z.Y. Dual roles of MoS2 nanosheets in advanced oxidation Processes: Activating permonosulfate and quenching radicals. Chem. Eng. J. 2022, 440, 135866. [Google Scholar] [CrossRef]
  12. Fan, J.X.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Lu, J.M. Adsorption and biodegradation of dye in wastewater with Fe3O4@MIL-100 (Fe) core-shell bio-nanocomposites. Chemosphere 2018, 191, 315–323. [Google Scholar] [CrossRef] [PubMed]
  13. Akerdi, A.G.; Bahrami, S.H. Application of heterogeneous nano-semiconductors for photocatalytic advanced oxidation of organic compounds: A review. J. Environ. Chem. Eng. 2019, 7, 103283. [Google Scholar] [CrossRef]
  14. Xin, L.; Hu, J.W.; Xiang, Y.Q.; Li, C.F.; Fu, L.Y.; Li, Q.H.; Wei, X.H. Carbon-Based Nanocomposites as Fenton-Like Catalysts in Wastewater Treatment Applications: A Review. Materials 2021, 14, 2643. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, H.Y.; Tian, W.J.; Zhou, L.; Sun, H.Q.; Tade, M.; Wang, S.B. Monodisperse Co3O4 quantum dots on porous carbon nitride nanosheets for enhanced visible-light-driven water oxidation. Appl. Catal. B Environ. 2018, 223, 2–9. [Google Scholar] [CrossRef]
  16. Chen, Z.F.; Pan, J.Q.; Mei, J.; Yu, Q.; Wang, P.H.; Wang, P.P.; Wang, J.J.; Song, C.S.; Zheng, Y.Y.; Li, C.R. Ternary Co3O4/CdS/SrTiO3 core-shell pn junctions toward enhanced photocatalytic hydrogen production activity. J. Environ. Chem. Eng. 2021, 9, 104895. [Google Scholar] [CrossRef]
  17. Wang, L.; Wan, J.W.; Zhao, Y.S.; Yang, N.L.; Wang, D. Hollow Multi-Shelled Structures of Co3O4 Dodecahedron with Unique Crystal Orientation for Enhanced Photocatalytic CO2 Reduction. J. Am. Chem. Soc. 2019, 141, 2238–2241. [Google Scholar] [CrossRef]
  18. Bilge, S.; Karadurmus, L.; Atici, E.B.; Sinag, A.; Ozkan, S.A. A novel electrochemical sensor based on magnetic Co3O4 nanoparticles/carbon recycled from waste sponges for sensitive determination of anticancer ruxolitinib. Sens. Actuator B Chem. 2022, 367, 132127. [Google Scholar] [CrossRef]
  19. You, D.J.; Lou, J.Y.; Li, X.Q.; Zhou, Y.L.; Sun, X.Q.; Wang, X.L. Investigation of advanced catalytic effect of Co3O4 nanosheets modified carbon felts as vanadium flow battery electrodes. J. Power Sources 2021, 494, 229775. [Google Scholar] [CrossRef]
  20. Bronzato, J.D.; Tofanello, A.; Oliveira, M.T.; Bettini, J.; Brito, A.M.M.; Costa, S.A.; Costa, S.M.; Lanfredi, A.J.C.; Nascimento, O.R.; Nantes-Cardoso, I.L. Virucidal, photocatalytic and chiromagnetic cobalt oxide quantum dots. Appl. Surf. Sci. 2022, 576, 151847. [Google Scholar] [CrossRef]
  21. Liu, D.L.; Pervaiz, E.; Adimi, S.; Thomas, T.; Qu, F.D.; Huang, C.Z.; Wang, R.; Jiang, H.; Yang, M.H. Theoretical study on W-Co3O4 surface: Acetone adsorption and sensing mechanism. Appl. Surf. Sci. 2021, 566, 150642. [Google Scholar] [CrossRef]
  22. Wang, Y.L.; Yu, D.; Wang, W.; Gao, P.; Zhong, S.; Zhang, L.S.; Zhao, Q.Q.; Liu, B.J. Synthesizing Co3O4-BiVO4/g-C3N4 heterojunction composites for superior photocatalytic redox activity. Sep. Purif. Technol. 2020, 239, 116562. [Google Scholar] [CrossRef]
  23. Cho, E.C.; Chang-Jian, C.W.; Huang, J.H.; Huang, T.Y.; Wu, N.J.; Li, M.T.; Chen, Y.L.; Hsu, S.C.; Weng, H.C.; Lee, K.C. Preparation of Ni(OH)2/CuO heterostructures for improved photocatalytic degradation of organic pollutants and microorganism. Chemosphere 2022, 300, 134484. [Google Scholar] [CrossRef] [PubMed]
  24. Si, Y.L.; Cao, S.; Wu, Z.J.; Ji, Y.L.; Mi, Y.; Wu, X.C.; Liu, X.F.; Piao, L.Y. The effect of directed photogenerated carrier separation on photocatalytic hydrogen production. Nano Energy 2017, 41, 488–493. [Google Scholar] [CrossRef]
  25. Xiao, F.X. Construction of Highly Ordered ZnO-TiO2 Nanotube Arrays (ZnO/TNTs) Heterostructure for Photocatalytic Application. ACS Appl. Mater. Interfaces 2012, 4, 7054–7062. [Google Scholar] [CrossRef] [PubMed]
  26. Bisaria, K.; Sinha, S.; Singh, R.; Iqbal, H.M.N. Recent advances in structural modifications of photo-catalysts for organic pollutants degradation-A comprehensive review. Chemosphere 2021, 284, 131263. [Google Scholar] [CrossRef]
  27. Qi, K.Z.; Cheng, B.; Yu, J.G.; Ho, W.K. Review on the improvement of the photocatalytic and antibacterial activities of ZnO. J. Alloy. Compd. 2017, 727, 792–820. [Google Scholar] [CrossRef]
  28. Meng, A.Y.; Zhang, J.; Xu, D.F.; Cheng, B.; Yu, J.G. Enhanced photocatalytic H2 production activity of anatase TiO2 nanosheet by selectively depositing dual-cocatalysts on {101} and {001} facets. Appl. Catal. B Environ. 2016, 198, 286–294. [Google Scholar] [CrossRef]
  29. Zhou, G.; Xu, X.Y.; Ding, T.; Feng, B.; Bao, Z.J.; Hu, J.G. Well-Steered Charge-Carrier Transfer in 3D Branched CuxO/ZnO@Au Heterostructures for Efficient Photocatalytic Hydrogen Evolution. ACS Appl. Mater. Interfaces 2015, 7, 26819–26827. [Google Scholar] [CrossRef]
  30. Jin, C.Y.; Wang, M.; Li, Z.L.; Kang, J.; Zhao, Y.; Han, J.; Wu, Z.M. Two dimensional Co3O4/g-C3N4 Z-scheme heterojunction: Mechanism insight into enhanced peroxymonosulfate-mediated visible light photocatalytic performance. Chem. Eng. J. 2020, 398, 125569. [Google Scholar] [CrossRef]
  31. Chahar, D.; Taneja, S.; Bisht, S.; Kesarwani, S.; Thakur, P.; Thakur, A.; Sharma, P.B. Photocatalytic activity of cobalt substituted zinc ferrite for the degradation of methylene blue dye under visible light irradiation. J. Alloy. Compd. 2021, 851, 156878. [Google Scholar] [CrossRef]
  32. Zhang, H.Y.; Wang, Z.W.; Li, R.N.; Guo, J.L.; Li, Y.; Zhu, J.M.; Xie, X.Y. TiO2 supported on reed straw biochar as an adsorptive and photocatalytic composite for the efficient degradation of sulfamethoxazole in aqueous matrices. Chemosphere 2017, 185, 351–360. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, Y.L.; Dan, Z.H.; Liang, Y.F.; Wang, Y.; Qin, F.X.; Chang, H. Asynchronous Evolution of Nanoporous Silver on Dual-Phase Ag-Sn Alloys by Potentiostatic Dealloying in Hydrochloric Acid Solution. Nanomaterials 2019, 9, 743. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, X.X.; Xu, W.C.; Cao, S.; Zhu, S.L.; Liang, Y.Q.; Cui, Z.D.; Yang, X.J.; Li, Z.Y.; Wu, S.L.; Inoue, A.; et al. An amorphous nanoporous PdCuNi-S hybrid electrocatalyst for highly efficient hydrogen production. Appl. Catal. B Environ. 2019, 246, 156–165. [Google Scholar] [CrossRef]
  35. Wada, T.; Yubuta, K.; Inoue, A.; Kato, H. Dealloying by metallic melt. Mater. Lett. 2011, 65, 1076–1078. [Google Scholar] [CrossRef]
  36. Pang, F.J.; Wang, Z.F.; Zhang, K.; He, J.; Zhang, W.Q.; Guo, C.X.; Ding, Y. Bimodal nanoporous Pd3Cu1 alloy with restrained hydrogen evolution for stable and high yield electrochemical nitrogen reduction. Nano Energy 2019, 58, 834–841. [Google Scholar] [CrossRef]
  37. Wang, R.Y.; Xu, C.X.; Bi, X.X.; Ding, Y. Nanoporous surface alloys as highly active and durable oxygen reduction reaction electrocatalysts. Energy Environ. Sci. 2012, 5, 5281–5286. [Google Scholar] [CrossRef]
  38. Li, R.; Liu, X.J.; Wu, R.Y.; Wang, J.; Li, Z.B.; Chan, K.C.; Wang, H.; Wu, Y.; Lu, Z.P. Flexible Honeycombed Nanoporous/Glassy Hybrid for Efficient Electrocatalytic Hydrogen Generation. Adv. Mater. 2019, 31, 1904989. [Google Scholar] [CrossRef]
  39. Xu, W.C.; Zhu, S.L.; Liang, Y.Q.; Li, Z.Y.; Cui, Z.D.; Yang, X.J.; Inoue, A. Nanoporous CuS with excellent photocatalytic property. Sci. Rep. 2015, 5, 18125. [Google Scholar] [CrossRef]
  40. Wang, N.; Pan, Y.; Wu, S.K.; Zhang, E.M.; Dai, W.J. Rapid synthesis of rutile TiO2 nano-flowers by dealloying Cu60Ti30Y10 metallic glasses. Appl. Surf. Sci. 2018, 428, 328–337. [Google Scholar] [CrossRef]
  41. Li, M.; Li, Y.Y.; Zhang, Q.; Qin, C.L.; Zhao, W.M.; Wang, Z.F.; Inoue, A. Ultrafine Cu2O/CuO nanosheet arrays integrated with NPC/BMG composite rod for photocatalytic degradation. Appl. Surf. Sci. 2019, 483, 285–293. [Google Scholar] [CrossRef]
  42. Li, M.; Wang, Z.F.; Zhang, Q.; Qin, C.L.; Inoue, A.; Guo, W.B. Formation and evolution of ultrathin Cu2O nanowires on NPC ribbon by anodizing for photocatalytic degradation. Appl. Surf. Sci. 2020, 506, 144819. [Google Scholar] [CrossRef]
  43. Qin, S.Y.; Liu, Y.; Liu, S.M.; Wang, X.Y.; Li, Y.Y.; Qin, C.L.; Wang, Z.F.; Li, M. Self-standing porous Au/CuO nanowires with remarkably enhanced visible light absorption and photocatalytic performance. Appl. Surf. Sci. 2022, 594, 153443. [Google Scholar] [CrossRef]
  44. Sun, X.H.; Zheng, D.H.; Pan, F.D.; Qin, C.L.; Li, Y.Y.; Wang, Z.F.; Liu, Y. 3D nanoporous Ni@NiO/metallic glass sandwich electrodes without corrosion cracks for flexible supercapacitor application. Appl. Surf. Sci. 2021, 545, 149043. [Google Scholar] [CrossRef]
  45. Zhao, F.; Zheng, D.H.; Liu, Y.; Pan, F.D.; Deng, Q.B.; Qin, C.L.; Li, Y.Y.; Wang, Z.F. Flexible Co(OH)2/NiOxHy@Ni hybrid electrodes for high energy density supercapacitors. Chem. Eng. J. 2021, 415, 128871. [Google Scholar] [CrossRef]
  46. Zhang, Y.; Zheng, D.H.; Liu, S.M.; Qin, S.Y.; Sun, X.H.; Wang, Z.F.; Qin, C.L.; Li, Y.Y.; Zhou, J. Flexible porous Ni(OH)2 nanopetals sandwiches for wearable non-enzyme glucose sensors. Appl. Surf. Sci. 2021, 552, 149529. [Google Scholar] [CrossRef]
  47. Zheng, D.H.; Li, M.; Li, Y.Y.; Qin, C.L.; Wang, Y.C.; Wang, Z.F. A Ni(OH)2 nanopetals network for high-performance supercapacitors synthesized by immersing Ni nanofoam in water. Beilstein J. Nanotechnol. 2019, 10, 281–293. [Google Scholar] [CrossRef]
  48. Pradhan, A.C.; Uyar, T. Morphological Control of Mesoporosity and Nanoparticles within Co3O4-CuO Electrospun Nanofibers: Quantum Confinement and Visible Light Photocatalysis Performance. ACS Appl. Mater. Interfaces 2017, 9, 35757–35774. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, Q.; Man, L.Y.; Qin, C.L.; Wang, Z.F.; Zhao, W.M.; Li, Y.Y. Flexible Free-Standing CuxO/Ag2O (x = 1, 2) Nanowires Integrated with Nanoporous Cu-Ag Network Composite for Glucose Sensing. Nanomaterials 2020, 10, 357. [Google Scholar] [CrossRef]
  50. Rabani, I.; Zafar, R.; Subalakshmi, K.; Kim, H.S.; Bathula, C.; Seo, Y.S. A facile mechanochemical preparation of Co3O4@g-C3N4 for application in supercapacitors and degradation of pollutants in water. J. Hazard. Mater. 2021, 407, 124360. [Google Scholar] [CrossRef]
  51. Xie, X.B.; Ni, C.; Lin, Z.H.; Wu, D.; Sun, X.Q.; Zhang, Y.P.; Wang, B.; Du, W. Phase and morphology evolution of high dielectric CoO/Co3O4 particles with Co3O4 nanoneedles on surface for excellent microwave absorption application. Chem. Eng. J. 2020, 396, 125205. [Google Scholar] [CrossRef]
  52. Qin, C.L.; Zhang, M.M.; Li, B.E.; Li, Y.Y.; Wang, Z.F. Ag particles modified CuxO (x = 1, 2) nanowires on nanoporous Cu-Ag bimetal network for antibacterial applications. Mater. Lett. 2020, 258, 126823. [Google Scholar] [CrossRef]
  53. Qin, C.L.; Zhang, W.; Asami, K.; Kimura, H.; Wang, X.M.; Inoue, A. A novel Cu-based BMG composite with high corrosion resistance and excellent mechanical properties. Acta Mater. 2006, 54, 3713–3719. [Google Scholar] [CrossRef]
  54. Qin, C.L.; Oak, J.J.; Ohtsu, N.; Asami, K.; Inoue, A. XPS study on the surface films of a newly designed Ni-free Ti-based bulk metallic glass. Acta Mater. 2007, 55, 2057–2063. [Google Scholar] [CrossRef]
  55. Qin, C.L.; Zheng, D.H.; Hu, Q.F.; Zhang, X.M.; Wang, Z.F.; Li, Y.Y.; Zhu, J.S.; Ou, J.Z.; Yang, C.H.; Wang, Y.C. Flexible integrated metallic glass-based sandwich electrodes for high-performance wearable all-solid-state supercapacitors. Appl. Mater. Today 2020, 19, 100539. [Google Scholar] [CrossRef]
  56. Yao, M.M.; Hu, Z.H.; Xu, Z.J.; Liu, Y.F. Template synthesis of 1D hierarchical hollow Co3O4 nanotubes as high performance supercapacitor materials. J. Alloy. Compd. 2015, 644, 721–728. [Google Scholar] [CrossRef]
  57. Zhang, Y.; Huang, J.W.; Ding, Y. Porous Co3O4/CuO hollow polyhedral nanocages derived from metal-organic frameworks with heterojunctions as efficient photocatalytic water oxidation catalysts. Appl. Catal. B Environ. 2016, 198, 447–456. [Google Scholar] [CrossRef]
  58. Wang, S.S.; Liu, L. Fabrication of novel nanoporous copper powder catalyst by dealloying of ZrCuNiAl amorphous powders for the application of wastewater treatments. J. Hazard. Mater. 2017, 340, 445–453. [Google Scholar] [CrossRef]
  59. Bokare, A.D.; Choi, W. Review of iron-free Fenton-like systems for activating H2O2 in advanced oxidation processes. J. Hazard. Mater. 2014, 275, 121–135. [Google Scholar] [CrossRef]
  60. Cheng, M.; Liu, Y.; Huang, D.L.; Lai, C.; Zeng, G.M.; Huang, J.H.; Liu, Z.F.; Zhang, C.; Zhou, C.Y.; Qin, L.; et al. Prussian blue analogue derived magnetic Cu-Fe oxide as a recyclable photo-Fenton catalyst for the efficient removal of sulfamethazine at near neutral pH values. Chem. Eng. J. 2019, 362, 865–876. [Google Scholar] [CrossRef]
Figure 1. (a) SEM images of Cu40Zr60 amorphous alloy ribbon de-alloyed in 0.05 M HF for 2 h; The composite sample via deposition at −0.8 V for 30 min followed by the calcination: (b) Plan-view SEM images; (c) Cross-sectional SEM image; (d) Nitrogen adsorption-desorption curve and pore size distribution of Co3O4@NP-CuO.
Figure 1. (a) SEM images of Cu40Zr60 amorphous alloy ribbon de-alloyed in 0.05 M HF for 2 h; The composite sample via deposition at −0.8 V for 30 min followed by the calcination: (b) Plan-view SEM images; (c) Cross-sectional SEM image; (d) Nitrogen adsorption-desorption curve and pore size distribution of Co3O4@NP-CuO.
Nanomaterials 12 02850 g001
Figure 2. (a) XRD patterns of the as-spun Cu40Zr60 ribbon and as-dealloyed sample immersed in 0.05 M HF for 2 h; (b) XRD patterns of as-deposited sample immersed in 60 mM Co(NO3)2 at −0.8 V for 30 min and as-calcined sample at 573 K for 2 h.
Figure 2. (a) XRD patterns of the as-spun Cu40Zr60 ribbon and as-dealloyed sample immersed in 0.05 M HF for 2 h; (b) XRD patterns of as-deposited sample immersed in 60 mM Co(NO3)2 at −0.8 V for 30 min and as-calcined sample at 573 K for 2 h.
Nanomaterials 12 02850 g002
Figure 3. (a) XPS survey of the Co3O4 nanopetals layer and CuO nanoporous layer for the Co3O4@NP-CuO composite; (b) The deconvolution of Co 2p spectrum of the nanopetals layer; (c) Cu 2p spectrum of nanoporous layer; (d) O 1s spectra of Co3O4 nanopetals layer and CuO nanoporous layer for the Co3O4@NP-CuO.
Figure 3. (a) XPS survey of the Co3O4 nanopetals layer and CuO nanoporous layer for the Co3O4@NP-CuO composite; (b) The deconvolution of Co 2p spectrum of the nanopetals layer; (c) Cu 2p spectrum of nanoporous layer; (d) O 1s spectra of Co3O4 nanopetals layer and CuO nanoporous layer for the Co3O4@NP-CuO.
Nanomaterials 12 02850 g003
Figure 4. TEM images of nanopetals layer of the Co3O4@NP-CuO composite: (a) high magnification image; (b) SAED pattern; (c) HRTEM image; EDS mapping images of nanopetals of (d) dark field image; (e) Co and O mixed; (f) Co; (g) O.
Figure 4. TEM images of nanopetals layer of the Co3O4@NP-CuO composite: (a) high magnification image; (b) SAED pattern; (c) HRTEM image; EDS mapping images of nanopetals of (d) dark field image; (e) Co and O mixed; (f) Co; (g) O.
Nanomaterials 12 02850 g004
Figure 5. SEM and TEM images of nanoporous layer of the Co3O4@NP-CuO: (a) high magnification SEM image; (b) TEM image; (c) HRTEM image; (d) SAED pattern; EDS mapping images of (e) O; (f) Cu.
Figure 5. SEM and TEM images of nanoporous layer of the Co3O4@NP-CuO: (a) high magnification SEM image; (b) TEM image; (c) HRTEM image; (d) SAED pattern; EDS mapping images of (e) O; (f) Cu.
Nanomaterials 12 02850 g005
Figure 6. (a) UV-visible absorption spectra of RhB by the Co3O4@NP-CuO composite in the presence of H2O2 for different times; (b) degradation performance of the Co3O4@NP-CuO composite, commercial CuO, and Co3O4 powders; (c) effects of different conditions on degradation of RhB; (d) cyclic stability of Co3O4@NP-CuO composite.
Figure 6. (a) UV-visible absorption spectra of RhB by the Co3O4@NP-CuO composite in the presence of H2O2 for different times; (b) degradation performance of the Co3O4@NP-CuO composite, commercial CuO, and Co3O4 powders; (c) effects of different conditions on degradation of RhB; (d) cyclic stability of Co3O4@NP-CuO composite.
Nanomaterials 12 02850 g006
Figure 7. (a) Degradation properties of Co3O4@NP-CuO composite at different temperatures, (b) corresponding L-H pseudo-first-order dynamic model and, (c) Arrhenius plot.
Figure 7. (a) Degradation properties of Co3O4@NP-CuO composite at different temperatures, (b) corresponding L-H pseudo-first-order dynamic model and, (c) Arrhenius plot.
Nanomaterials 12 02850 g007
Figure 8. (a) PL spectra of Co3O4@NP-CuO, CuO, and Co3O4; (b) Effect of different free radicals on photocatalytic degradation of Co3O4@NP-CuO; (c) EPR spectra of ·OH generated under different conditions for the Co3O4@NP-CuO.
Figure 8. (a) PL spectra of Co3O4@NP-CuO, CuO, and Co3O4; (b) Effect of different free radicals on photocatalytic degradation of Co3O4@NP-CuO; (c) EPR spectra of ·OH generated under different conditions for the Co3O4@NP-CuO.
Nanomaterials 12 02850 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sun, Y.; Wang, C.; Qin, S.; Pan, F.; Li, Y.; Wang, Z.; Qin, C. Co3O4 Nanopetals Grown on the Porous CuO Network for the Photocatalytic Degradation. Nanomaterials 2022, 12, 2850. https://doi.org/10.3390/nano12162850

AMA Style

Sun Y, Wang C, Qin S, Pan F, Li Y, Wang Z, Qin C. Co3O4 Nanopetals Grown on the Porous CuO Network for the Photocatalytic Degradation. Nanomaterials. 2022; 12(16):2850. https://doi.org/10.3390/nano12162850

Chicago/Turabian Style

Sun, Yuntao, Can Wang, Shengyao Qin, Fengda Pan, Yongyan Li, Zhifeng Wang, and Chunling Qin. 2022. "Co3O4 Nanopetals Grown on the Porous CuO Network for the Photocatalytic Degradation" Nanomaterials 12, no. 16: 2850. https://doi.org/10.3390/nano12162850

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