3.1 XRD analysis
In order to investigate the crystalline structure of the prepared composites, X-ray diffraction (XRD) examination of Zn0.25Cd0.75S QDs, Zn0.25Cd0.75S QDs/ZnO, and Pt/Zn0.25Cd0.75S QDs/ZnO were carried out in this paper, and the results are shown in Fig. 1. As can be seen from Fig. 1(a), Zn0.25Cd0.75S QDs do not have any diffraction peaks that exactly correspond to CdS (JCPDS#41-1049)[35] and ZnS (JCPDS#05-0566)[36]. However, based on the comparison of peak intensities, the diffraction peaks of ZnS and CdS exist in Zn0.25Cd0.75S QDs, and these diffraction peaks shift regularly. The diffraction peaks of Zn0.25Cd0.75S QDs shift to higher angles in comparison to CdS, while shift to lower angles compared with ZnS[37]. This suggests that Zn0.25Cd0.75S QDs material is not a mixture of ZnS and CdS, but belongs to ZnxCd1−xS solid-solution quantum dots.
Meanwhile, XRD image of Zn0.25Cd0.75S QDs/ZnO given in Fig. 1(b) shows that the diffraction peaks of the composite Zn0.25Cd0.75S QDs/ZnO at 27.5°, 45.3°, and 54.2° correspond to those of Zn0.25Cd0.75S QDs. Moreover, the diffraction peaks at 32.0°, 34.5°, 36.5°, 47.6°, 56.6°, 62.9°, 66.4°, 68.0°, 69.1°, 72.6°, 81.4°, and 89.6° correspond to the standard PDF cards of ZnO (JCPDS#36-1451)[38], indicating that the composite material consists of Zn0.25Cd0.75S QDs and ZnO. In addition, the composite has very high diffraction peaks at 32.0°, 34.5°, and 36.5° consistent with the (100), (002), and (101) crystal planes of ZnO, proving that ZnO has a good crystallinity and exists as a hexagonal fibrillar zincite structure in the composites[39].
Furthermore, XRD results of Pt/Zn0.25Cd0.75S QDs/ZnO show that the diffraction peaks of Pt/Zn0.25Cd0.75S QDs/ZnO correspond to those of Zn0.25Cd0.75S QDs/ZnO without obvious shift (Fig. 1(c)), indicating that Zn0.25Cd0.75S QDs/ZnO is present in the composite Pt/Zn0.25Cd0.75S QDs/ZnO. By comparing the intensities of diffraction peaks between Pt/Zn0.25Cd0.75S QDs/ZnO and Zn0.25Cd0.75S QDs/ZnO in Fig. 1(d), the intensities of ZnO in Pt/Zn0.25Cd0.75S QDs/ZnO are weaker, whereas the diffraction peaks of Zn0.25Cd0.75S QDs in Pt/Zn0.25Cd0.75S QDs/ZnO are sharper, which suggests that the presence of Pt promotes the growth and crystallisation of Zn0.25Cd0.75S QDs. However, the diffraction peaks of Pt are not observed in Fig. 1(c), which may be due to the small amount of its deposition.
3.2 UV–vis/DRS analysis
As can be seen from Fig. 2(a), the monomer ZnO synthesized in this work has an absorption edge wavelength of 401 nm with a strong absorption in the UV region[40]. Meanwhile, the absorption edge wavelength of the monomer Zn0.25Cd0.75S QDs is 526 nm, displaying a certain absorption in the visible region. The absorption edge of Zn0.25Cd0.75S QDs/ZnO shows an obvious redshift compared with ZnO, with the wavelength of 457 nm. After the deposition of noble metal Pt, the absorption range of Pt/Zn0.25Cd0.75S QDs/ZnO composite is further broadened and the absorption edge is redshifted to 497 nm.
The band gap energies of ZnO, Zn0.25Cd0.75S QDs, Zn0.25Cd0.75S QDs/ZnO and Pt/Zn0.25Cd0.75S QDs/ZnO were calculated from Fig. 2(b) using the Kubelka–Munk formula (see Formula S1)[41], and the corresponding results are given in Table S1. The bandgap value of the composite material decreases after combining the wide bandgap semiconductor material ZnO with narrow bandgap Zn0.25Cd0.75S QDs. Meanwhile, the loading of Pt further reduces the bandgap value of the composite, preliminarily suggesting that both the heterostructure and the loading of noble metals can change the bandgap value of the material.
Meanwhile, the absorption edge wavelength of the monomer Zn0.25Cd0.75S QDs is 526 nm, displaying a certain absorption in the visible region. The absorption edge of Zn0.25Cd0.75S QDs/ZnO shows an obvious redshift compared with ZnO, with the wavelength of 457 nm. After the deposition of noble metal Pt, the absorption range of Pt/Zn0.25Cd0.75S QDs/ZnO composite is further broadened and the absorption edge is redshifted to 497 nm.
The band gap energies of ZnO, Zn0.25Cd0.75S QDs, Zn0.25Cd0.75S QDs/ZnO and Pt/Zn0.25Cd0.75S QDs/ZnO were calculated from Fig. 2(b) using the Kubelka–Munk formula (see Formula S1)[41], and the corresponding results are given in Table S1. The bandgap value of the composite material decreases after combining the wide bandgap semiconductor material ZnO with narrow bandgap Zn0.25Cd0.75S QDs. Meanwhile, the loading of Pt further reduces the bandgap value of the composite, preliminarily suggesting that both the heterostructure and the loading of noble metals can change the bandgap value of the material.
3.3 SEM and SEM-EDS analysis
In order to investigate the surface morphology, Pt/Zn0.25Cd0.75S QDs/ZnO composite was analyzed by scanning electron microscopy (SEM), and SEM images at different scales are shown in Fig. 3. It can be seen from Fig. 3(a–c) that the prepared composites are made up of interconnected nanoparticles, forming a skeleton-like morphology.
At the same time, the composite material presents pore structure with a relatively loose state. This structure is not only conducive to the transfer of substances, but also easy to adsorb dyes. From the magnified SEM photograph in Fig. 3(b), it can be observed that the nanoparticles have a rough surface and a uniform size of about 10 nm. And EDS analysis results of Pt/Zn0.25Cd0.75S QDs/ZnO in Fig. 3(d–h) shows that the elements of Zn, O, Cd, S and Pt are uniformly distributed on the surface of the material, indicating the presence of the above-mentioned elements in the composite material with the content of each element shown in Fig. S2.
3.4 TEM and HR-TEM analysis
For investigating the morphological structure and composition of the Pt/Zn0.25Cd0.75S QDs/ZnO composites, TEM analyses were carried out. In Fig. 4(a–c), it can be clearly observed that the prepared composites consist of several nanoparticles agglomerated together with a size of about 10 nm, which is consistent with the SEM results. Fig. 4(d), Fig. S3 and Fig. 4(e–j) show the HR-TEM maps and the Fast Fourier Transform (FFT) of Pt/Zn0.25Cd0.75S QDs/ZnO, respectively. From Fig. 4(d) and Fig. S3, it can be seen that the average particle size of Zn0.25Cd0.75S QDs is about 2–3 nm, which proves that the solid solution Zn0.25Cd0.75S exists in the form of quantum dots on the surface of ZnO[42]. In addition, the lattice spacing in Fig. 4(e) is n=0.278 nm, which belongs to the (100) crystal faces of ZnO[43]. The lattice spacing in Fig. 4(f) is n=0.296 nm, corresponding to the (101) crystallographic plane of the solid solution ZnxCd1-xS QDs[44]. The lattice spacing in Fig. 4(g) is n=0.203 nm, which corresponds to the (200) crystal plane of Pt, demonstrating the presence of noble metal Pt monomers in the composite[45]. The above lattice spacing results indicate the presence of hexagonal fibrillar zincite phase ZnO, metallic Pt, and Zn0.25Cd0.75S QDs in the prepared composites.
3.5 XPS analysis
In this paper, the elemental distribution and valence states on the surface of the composite material Pt/Zn0.25Cd0.75S QDs/ZnO were obtained by XPS detection and the results are shown in Fig. 5. From Fig. 5(a), it can be seen that the surface of the composite material contains Zn, Cd, Pt, S and O elements. Among them, Zn 2p at 1044.5 eV and 1021.4 eV in Fig. 5(b) corresponds to Zn 2p3/2 and Zn 2p1/2, respectively, indicating the presence of Zn2+ on the surface of the composite[46]. Figure 5(c) shows the XPS energy spectrum of Cd, where the binding energies of 411.0 eV and 404.3 eV correspond to Cd 3d3/2 and Cd 3d5/2, respectively, indicating that the Cd element in the composites is present at + 2 valence[47]. The binding energies of 71.7 eV and 74.7 eV correspond to Pt 4f5/2 and Pt 4f7/2 in Fig. 5(f), respectively, indicating that the Pt element exists in the form of Pt0 monomers on the surface of the composite[48]. Meanwhile, the binding energies of S 2p at 162.2 eV and 161.0 eV in Fig. 5(e) are consistent with those of S 2p3/2 and S 2p1/2, indicating that S exists in the − 2 valence state[49]. In addition, there are characteristic peaks of O 1s in Fig. 5(f) at 530.8 eV and 532.3 eV, corresponding to the O atoms in Zn–O formed between O2− and Zn2+ ions and in the form of hydroxyl groups on the surface of the composite, respectively[50].
3.6 N2 adsorption–desorption analysis
In order to investigate the surface physicochemical properties of the composites, N2 adsorption–desorption tests were carried out on ZnO, Zn0.25Cd0.75S QDs/ZnO, and Pt/Zn0.25Cd0.75S QDs/ZnO, and the results are shown in Fig. 6 and Table S2. As can be seen from Fig. 6, according to the IUPAC definition[51], the N2 adsorption–desorption isotherms of the above materials belong to type IV adsorption curve with H3-type hysteresis loops, indicating that all materials have mesoporous structure.
According to the data in Table S2, it can be seen that the specific surface area of ZnO is small, and after the loading of Zn0.25Cd0.75S QDs, the specific surface area becomes approximately five times larger than the monomer ZnO, which further proves that solid solution Zn0.25Cd0.75S exists on the surface of the semiconductor ZnO in the form of quantum dots, resulting in the rough surface and increased specific surface area. However, after loading with Pt, the specific surface area of Pt/Zn0.25Cd0.75S QDs/ZnO only has a small increase of 0.31 m2·g− 1 compared with Zn0.25Cd0.75S QDs/ZnO, which may be due to the low loading amount of Pt.
3.7 Photoluminescence analysis
In order to investigate the photogenerated electron–hole recombination rate of the composites, photoluminescence was used to examine the composites in this paper, and the results are shown in Fig. 7. The photoluminescence spectra of ZnO, Zn0.25Cd0.75S QDs/ZnO, and Pt/Zn0.25Cd0.75S QDs/ZnO samples were obtained under the excitation of wavelength of 300 nm. As can be seen in Fig. 7, ZnO has the highest emission peak, indicating a high electron–hole recombination rate. And after combining with Zn0.25Cd0.75S QDs, the intensity of the emission peak is significantly reduced, which indicates that the construction of heterostructures in the composites can improve the efficiency of photogenerated electron–hole separation. In addition, when the noble metal Pt is loaded on the surface of Zn0.25Cd0.75S QDs/ZnO composite, Pt/Zn0.25Cd0.75S QDs/ZnO composite shows the lowest fluorescence intensity, and thus Pt/Zn0.25Cd0.75S QDs/ZnO has a good carrier separation performance, which is because the noble metal Pt has the ability to transfer electrons on the effect of surface plasmon resonance (SPR), thereby increasing the electron transfer path and effectively suppressing the recombination of electrons and holes[52].
3.8 Electrochemical testing
In order to investigate the photogenerated carrier separation efficiency of ZnO, Zn0.25Cd0.75S QDs, Zn0.25Cd0.75S QDs/ZnO, and Pt/Zn0.25Cd0.75S QDs/ZnO, transient photocurrent response tests and electrochemical impedance tests were carried out, and the results are shown in Fig. 8. The transient photocurrent response can further confirm the electron–hole transfer ability of ZnO, Zn0.25Cd0.75S QDs, Zn0.25Cd0.75S QDs/ZnO, and Pt/Zn0.25Cd0.75S QDs/ZnO. The higher the photocurrent intensity, the faster the carrier migration on the surface of the photocatalyst[53]. As shown in Fig. 8(a), the transient photocurrent intensity of the composite Zn0.25Cd0.75S QDs/ZnO is almost four times higher than that of ZnO, which is because the formation of heterostructure after the combination of ZnO and Zn0.25Cd0.75S QDs promotes the photogenerated current intensity, as the result of generating a large number of electrons with directional motion in Zn0.25Cd0.75S QDs/ZnO. After the deposition of the noble metal Pt, the transient photocurrent intensity of the composite Pt/Zn0.25Cd0.75S QDs/ZnO is greatly enhanced, which is about 1.5 times higher than that of Zn0.25Cd0.75S QDs/ZnO and 5 times higher than that of the monomer ZnO, indicating that the SPR effect of the noble metal Pt effectively inhibits the recombination of photogenerated electrons and holes, and thus improving the survival rate of photogenerated carriers.
In addition, the charge transfer efficiencies of all above materials were tested as shown in Fig. 8(b). The electrochemical impedance results are as follows: Pt/Zn0.25Cd0.75S QDs/ZnO > Zn0.25Cd0.75S QDs/ZnO > Zn0.25Cd0.75S QDs > ZnO. In general, the material with smaller radius of electrochemical impedance has lower charge transfer resistance and higher transfer efficiency[54]. The figure shows that Pt/Zn0.25Cd0.75S QDs/ZnO has the smallest radius, indicating that it has the fastest electron transfer rate and the lowest charge transfer resistance.
3.9 Photocatalytic experiment results of Pt/Zn0.25Cd0.75S QDs/ZnO composite
In order to investigate the photocatalytic performance of Pt/Zn0.25Cd0.75S QDs/ZnO series of materials, multi-mode photocatalytic experiments were carried out using methyl orange (MO) as a model molecule to compare the differences in photocatalytic performance under different light sources, and the relevant results are shown in Fig. 9. According to the results of Fig. 9(a), it can be seen that the degradation efficiency of Pt/Zn0.25Cd0.75S QDs/ZnO is obviously improved, which is basically in line with the results of the above-mentioned UV–vis diffuse absorption spectroscopy analysis. The high photocatalytic degradation efficiency of Pt/Zn0.25Cd0.75S QDs/ZnO is due, on the one hand, to the synergistic effect between Zn0.25Cd0.75S QDs composite and the monomer ZnO with strong absorption ability in the UV region, on the other hand, to the modification of Pt. In order to have a clearer understanding of the effect and influence of different catalysts on the degradation of MO under UV light irradiation, the reaction kinetics graphs are given in this paper as Fig. 9(b), which is calculated as -ln(Ct/C0) = kt [55], and the calculated rate constants of the reaction, k, are shown in Table S3. From Fig. 9(b) it can be seen that -ln(Ct/C0) is essentially linear with the reaction time t, indicating that the degradation of the dye (MO) by the composites follows quasi-primary reaction kinetics.
In addition, in order to further investigate the photocatalytic performance of the composite Pt/Zn0.25Cd0.75S QDs/ZnO, experiments were carried out on the degradation of MO by the composites under different light source conditions, and the results are shown in Fig. 9(c) and Fig. 9(d). The degradation efficiency of Pt/Zn0.25Cd0.75S QDs/ZnO on MO is about 4.5 times and 1.5 times that of the monomer ZnO under visible light and simulated sunlight irradiation, respectively, indicating that Pt/Zn0.25Cd0.75S QDs/ZnO are able to significantly improve the utilization of sunlight during the photolysis process and consequently has a good degradation activity.
Furthermore, in this paper, the hydrogen production performance of ZnO, Zn0.25Cd0.75S QDs, Zn0.25Cd0.75S QDs/ZnO, and Pt/Zn0.25Cd0.75S QDs/ZnO was investigated by using a mixture solution of 0.1 M sodium sulfide (Na2S) and 0.1 M sodium sulfite (Na2SO3) as a sacrificial agent, and the results are shown in Fig. 10. Under the irradiation of a 300 W xenon lamp for 8 h, Zn0.25Cd0.75S QDs show good hydrogen-producing activity compared with ZnO. Meanwhile, the hydrogen production of Zn0.25Cd0.75S QDs/ZnO composites is about 14 times higher than that of the monomer ZnO, suggesting that the construction of heterojunction structure has an obvious enhancement on the hydrogen production performance of ZnO.
In addition, further deposition of the noble metal Pt on the composite Zn0.25Cd0.75S QDs/ZnO results in a hydrogen production of 33.67 mmol·g− 1, which is 33 times higher than that of the monomer ZnO. This is because the loading of Pt effectively inhibits recombination of photoinduced electrons and holes, leading to a significant increase in the hydrogen production capacity of Pt/Zn0.25Cd0.75S QDs/ZnO composite.
In order to speculate the possible reaction mechanism of the composite Pt/Zn0.25Cd0.75S QDs/ZnO in the photocatalytic reaction, reactive group trapping experiments were performed under UV light conditions with MO as a molecular model using different trapping agents: p-benzoquinone (BQ), disodium ethylenediaminetetraacetic acid (EDTA-2Na) and isopropanol (IPA), and the results are shown in Fig. S4. As can be seen from Fig. S4, the photocatalytic degradation of MO by the composite Pt/Zn0.25Cd0.75S QDs/ZnO is significantly reduced after the addition of various scavengers, which is attributed to the decline in the number of reactive groups in the reaction system, leading to the decreased photocatalytic activity in the composite Pt/Zn0.25Cd0.75S QDs/ZnO. It can be seen that there are three types of reactive groups in the composites: ·O2−, h+ and ·OH, and their quantitative relationship is ·O2−> h+> ·OH.
3.10 Possible photocatalytic reaction mechanism of Pt/Zn0.25Cd0.75S QDs/ZnO
According to Eq. (2) and Eq. (3) (see Formula S2), the values of conduction band and valence band were calculated and the results are given in Table S4. Based on the above results and related experimental data, the possible photocatalytic reaction mechanism of Pt/Zn0.25Cd0.75S QDs/ZnO composites was speculated. As shown in Fig. 11, under simulated sunlight irradiation, Pt can be used as an electronic transducer to improve the efficiency of solar energy usage due to the difference in Fermi energy levels with ZnO. According to the conventional energy band structure theory, since the conduction band value of Zn0.25Cd0.75S QDs (-0.41 eV) is more negative than that of ZnO (-0.31 eV), the photogenerated electrons transfer from Zn0.25Cd0.75S QDs to the conduction band of ZnO, and the electrons in the conduction band of ZnO interact with the dissolved oxygen in water to generate ·O2−, which can degrade the dye molecules; and in the process of photolysis of water to produce hydrogen, the electrons in the conduction band of ZnO reduce water to H2. In addition, the valence band of ZnO (2.89 eV) is more positive than that of Zn0.25Cd0.75S QDs (2.05 eV), so that photogenerated holes transfer from ZnO to the valence band of Zn0.25Cd0.75S QDs. Since the valence band value of Zn0.25Cd0.75S QDs is 2.05 eV, which is less than the redox potential for the formation of ·OH (2.4 eV vs NHE), Zn0.25Cd0.75S QDs are not capable of generating ·OH, which does not match with the captured active species, and therefore the photocatalytic mechanism demonstrated in Fig. 11(a) does not agree with the results of photocatalytic capture experiments. Consequently, this study prefers the "Z-type"energy band structure theory based on Pt as the electronic transducer between ZnO and Zn0.25Cd0.75S, as shown in Fig. 11(b).
At first, under simulated sunlight irradiation, due to the relatively low redox capacity of the electrons in the conduction band of ZnO and the holes in the valence band of Zn0.25Cd0.75S QDs, Pt serves as an electron mediator to recombine the electrons and holes produced by the two and thus facilitates the separation of photogenerated carriers. Meanwhile, the photocorrosion of Zn0.25Cd0.75S QDs is effectively suppressed, resulting in the retention of electrons in the conduction band of Zn0.25Cd0.75S QDs and holes in the valence band of ZnO. The e− in the conduction band of Zn0.25Cd0.75S QDs has a large reducing capacity to produce ·O2− from O2 and mineralize MO to CO2 and H2O;In addition, the h+ in the valence band of ZnO is also strongly reducing, and one part reacts with H2O to produce hydroxyl radicals, followed by participating in the mineralization of pollutant molecules to produce CO2 and H2O; the other part reduces H2O to H2 in the presence of a sacrificial agent. The composite material transfers photogenerated carriers through the "Z-type" energy band structure theory and undergoes photodegradation and hydrogen production from photolytic water by the combined action of three different active species, which corresponds to the results of the capture experiments presented above.