Open Access Article
Qiu Zhang
*abc,
Yuekun Zhangb,
Chunxiao Zhanga,
Xiuyan Jianga and
Xuemei Fua
aSchool of Chemical Engineering, Shandong Institute of Petroleum and Chemical Technology, Dongying 257061, China. E-mail: 2023042@sdipct.edu.cn
bShandong Key Laboratory of Green Electricity & Hydrogen Science and Technology, Shandong Institute of Petroleum and Chemical Technology, Dongying, 257061, China
cDongying Key Laboratory of New Energy Materials and Devices, School of Chemical Engineering, Shandong Institute of Petroleum and Chemical Technology, Dongying, 257061, China
First published on 27th October 2025
Quantum dot-sensitized solar cells (QDSSCs) have emerged as promising photovoltaic devices, in which the counter electrode (CE) plays a crucial role in the catalytic reduction of Sn2− and charge transfer. Based on a precursor-directed strategy, this study reports a simple and cost-effective solvothermal method for the synthesis of morphology-controlled CuCo2S4 nanomaterials without templates and structure-directing agents, including flower-like (f-CuCo2S4), nanosheet-like (n-CuCo2S4), nanoparticle-like (p-CuCo2S4), and microsphere-like (m-CuCo2S4) structures. The effects of CuCo2S4 CEs with different morphologies on the photovoltaic performance of QDSSCs were also systematically investigated. Among them, the f-CuCo2S4 CE exhibited the highest specific surface area and the best catalytic performance, resulting in a power conversion efficiency (PCE) of 7.42% for QDSSCs, which is 55% higher than that of materials with other morphologies. Electrochemical analysis confirmed that it delivered the lowest charge transfer resistance (Rct = 0.076 Ω) and the highest electrocatalytic activity. This work highlights the importance of morphology control for optimizing the performance of CEs for efficient QDSSCs.
Classic QDSSCs have a sandwich structure similar to dye-sensitized solar cells (DSSCs),10,11 mainly including four basic elements: quantum dot sensitizer, photoanode, electrolyte and counter electrode (CE).12 Among them, QDs are used to absorb sunlight to generate electron–hole pairs. The photoanode not only acts as an electron transfer bridge, but also bears the role of loading QDs.13 The role of the electrolyte is to achieve the regeneration of the sensitizer.14 As an indispensable component of QDSSCs, the CE plays a pivotal role. Like an efficient catalyst, it is responsible for collecting and transferring electrons from the external circuit and catalyzing the reduction of oxidized electrolytes, ensuring that the chemical reactions inside the cell can proceed smoothly.15,16 Thus, each individual component significantly influences the diverse parameters of the cell. Throughout the existing research, the focus is mainly on the photoanode and electrolyte, while the reports on the CE are relatively few.
As a commonly used counter electrode material, the QDSSCs assembled with traditional brass foil-based Cu2S-CE exhibit relatively higher power conversion efficiency (PCE).17 However, the polysulfide electrolyte often corrodes the brass substrate continuously, resulting in Cu2S film falling off.18 Additionally, its relatively low specific surface area limits the rapid catalytic reduction reaction, then resulting in continuous decline in the mechanical stability and optoelectronic performance of the QDSSCs. Although the catalytic performance and stability of CuxS-based counter electrodes can be enhanced through optimization of the preparation process, the performance of CE materials composed of single metal sulfides is usually limited by the unsatisfactory properties of the single component,19,20 which makes the performance of QDSSCs devices unable to be further improved. Compared with single metal sulfides, bimetallic transition metal sulfides, benefiting from the synergistic effect among different types of mixed-valence metal ions, can generate abundant redox active sites, providing the required low activation energy for electron transfer, thereby significantly enhancing the overall catalytic performance of the material.21,22
In recent years, owing to their high intrinsic electrical conductivity, excellent structural stability, and remarkable synergistic catalytic performance, transition metal sulfides AB2S4 with spinel structure, such as CuCo2S4, NiCo2S4, MnCo2S4, CuFe2S4, etc., have been widely applied in the fields of supercapacitors, lithium-ion batteries, electrochemical sensing, oxygen evolution reactions, and photovoltaic cells.23–27 Among them, CuCo2S4 has the advantages of abundant reserves and non-toxicity.28 Moreover, copper-based materials are of low cost and possess excellent electrical conductivity, while cobalt-based materials exhibit excellent catalytic activity similar to that of noble metals.29 Therefore, in recent years, as a typical representative of spinel sulfides, CuCo2S4 has been extensively studied as the CE material for DSSCs.30,31 Sambandam Anandan et al. prepared CuCo2S4 nanosheets by ultrasonic synthesis method under low-temperature environmental conditions, and applied them to DSSC as CE catalyst, obtaining a PCE of 10.21%.32 Bakhytzhan Baptayev used the nanoflower-like CuCo2S4 prepared by the solvothermal method as the CE of DSSC, achieving a PCE of 7.56%.33 However, to the best of our knowledge, there are still few studies on applying CuCo2S4 as a CE in QDSSCs. Thus, the preparation of spinel Cu and Co metal sulfides into high-efficiency electrode in QDSSCs still has great development space.
Furthermore, the catalytic performance of CEs is closely related to the morphology of the material.34 The design of specific morphologies can not only increase the exposure of the active surface area and the number of catalytic sites, but also facilitate efficient electron transfer and electrolyte diffusion.11,35 As a result, reducing the charge transfer resistance and optimizing the redox kinetics. Nevertheless, in QDSSCs, there are still relatively few studies systematically exploring the influence of the morphology on the performance of CEs. Although a few studies have reported the synthesis methods of CuCo2S4 with definite morphology, template agents or structure-directing agents are usually required during the synthesis process.36,37 This makes the preparation method complicated, to a certain extent, restricts the widespread application of CuCo2S4. Thus, this study demonstrates a straightforward precursor-directed strategy for preparing CuCo2S4-CEs with diverse morphologies, and it is discovered that the nanoflower-like CuCo2S4 (f-CuCo2S4) has a large specific surface area and a three-dimensional porous structure, which can amplify the synergistic effect between Cu and Co ions, and exhibit excellent electrocatalytic activity towards polysulfide electrolytes. The QDSSCs based on the f-CuCo2S4-CE can achieve a PCE of 7.42% with Jsc = 25.42 mA cm−2, Voc = 0.596 V, (FF = 0.49), which is 24%, 39%, and 55% higher than that of the CuCo2S4-CE in the form of nanosheets (n-CuCo2S4), nanoparticles (p-CuCo2S4), and microspheres (m-CuCo2S4) respectively.
In addition, 5 mmol of Cd(NO3)2·4H2O was prepared into a 0.1 M methanol solution as a Cd2+ source, and 5 mmol of Na2S·9H2O was dissolved in a mixed solution of methanol and DI (volume ratio 1
:
1) to obtain a 0.1 M S2− source. The calcined photoanode was vertically immersed in the 0.1 M Cd2+ source for 2 min, then rinsed with methanol and dried with N2. Subsequently, the above film was vertically placed in a 0.1 M S2− source and immersed for 2 min, then rinsed with DI and dried with N2. This process is called a successive ion-layer adsorption and reaction (SILAR) cycle. After 6 cycles, FTO/TiO2/CdS can be prepared. CdSe-QDs were obtained by chemical bath deposition (CBD). A certain amount of CdSO4·8/3H2O and N(CH2COONa)3 (NTA) was weighed to prepare 0.03 M CdSO4·8/3H2O solution and 0.2 M NTA solution, respectively. According to the volume ratio of 1
:
1
:
1, the prepared Na2SeSO3 (the preparation method is presented in the SI), CdSO4·8/3H2O and NTA solution were mixed evenly and stirred continuously at room temperature for 20 min. Subsequently, the prepared FTO/TiO2/CdS photoanode was vertically immersed in the above mixed solution and allowed to stand for deposition in a dark environment at 5 °C for 6 h. After being taken out, it was fully rinsed with DI and dried with N2 to obtain FTO/TiO2/CdS/CdSe. Finally, the SILAR method was used to deposit the ZnS passivation layer. In short, the FTO/TiO2/CdS/CdSe photoanode was vertically immersed in the prepared 0.1 M Zn(CH3COO)2·2H2O solution for 2 min. After the electrode was taken out, it was rinsed with DI and dried with N2. Then the above electrode was vertically placed in 0.1 M Na2S·9H2O and immersed for 2 min. When the deposition was completed, the surface residue was fully rinsed with DI and dried with N2. After alternating for 3 cycles, the FTO/TiO2/CdS/CdSe/ZnS photoanode was received.
:
3, and then ultrasonically dissolved for 1 h. After complete dissolution, an orange-yellow polysulfide electrolyte was obtained.
Next, the FTO/TiO2/CdS/CdSe/ZnS photoanode and various CEs were fixed together face to face using a dovetail clamp. The prepared electrolyte (40 μL) was injected into the gap between the two electrodes using a pipette, then a sandwich-type QDSSCs was constructed. The symmetric cell was used to test the electrochemical performance of the counter electrode. Except for replacing the same two counter electrodes, the rest of the assembly process was consistent with the assembly of QDSSCs. The effective active area of the QDSSC was designed to be 0.12 cm2.
To better prove the elemental composition and chemical state of the sample, X-ray photoelectron spectroscopy (XPS) was further used to measure the flower-like CuCo2S4 sample. The XPS survey spectrum presented in Fig. 3(a) shows the presence of signal peaks of three elements (Cu, Co and S) in the sample, further proving the successful preparation of the sample. Additionally, Fig. 3(b) is a high-resolution XPS spectrum of Cu 2p. The binding energy at 932.8 eV and 952.7 eV correspond to the signal peaks of Cu 2p3/2 and Cu 2p1/2, respectively, confirming the presence of Cu+ in the CuCo2S4 sample.39 There are two weak signal peaks at 935.8 eV and 955.5 eV, indicating the presence of Cu2+ in the sample. Fig. 3(c) depicts the high-resolution XPS spectrum of Co 2p. The peak in the binding energy range of 779–782 eV points to Co 2p3/2, while the characteristic peak signal in the binding energy range of 793–798 eV points to Co 2p1/2. Simultaneously, the peaks at 779.1 eV and 794.2 eV binding energy correspond to the characteristic peaks of Co3+, the peak signals at 781.2 eV and 797.8 eV binding energy indicate the presence of Co2+, and the signal peaks at 784.8 eV and 802.9 eV are usually due to the satellite peaks of Co2+. Fig. 3(d) shows the XPS spectrum of S 2p. The S 2p signal peak of the CuCo2S4 sample can be further deconvoluted into two main peaks (S 2p3/2, S 2p1/2) at 162.1 eV and 163.5 eV. Furthermore, the low-intensity broad peak near 169.3 eV binding energy may be due to the partial oxidation of the sample surface to form S–O bonds. The above test results are consistent with the XRD results, further proving the successful synthesis of the CuCo2S4 sample.
![]() | ||
| Fig. 3 (a) XPS survey spectrum; (b) Cu 2p; (c) Co 2p; (d) S 2p high resolution XPS spectra of flower-like CuCo2S4. | ||
The microstructure and morphology of the CuCo2S4 samples were further characterized using SEM and TEM. As shown in Fig. 4(a), the CuCo2S4 synthesized using metal chlorides (CuCl2·2H2O and CoCl2·6H2O) as precursors exhibited a flower-like nanostructure with a diameter of approximately 1.2 μm. Simultaneously, Fig. 4(b)–(d) respectively show the nanosheet-like CuCo2S4 (from Cu(NO3)2·3H2O and Co(NO3)2·6H2O), loosely aggregated nanoparticle-like CuCo2S4 (from CuSO4·5H2O and CoSO4·7H2O), and microsphere-like CuCo2S4 (from Cu(CO2CH3)2·xH2O and Co(CO2CH3)2·4H2O). Obviously, compared with the other morphologies, the flower-like CuCo2S4 with its open, porous three-dimensional structure is more favorable for increasing the specific surface area of the material, thereby enhancing the number of active catalytic sites and promoting electrolyte infiltration. This indicates that the microscopic morphology of CuCo2S4 is influenced by the type of metal precursor, and that precise morphological control can be achieved by selecting different precursors, ultimately leading to improved counter electrode performance. Fig. 5(a)–(c) clearly depicts the TEM and high-resolution TEM (HRTEM) images of nanoflower CuCo2S4. As shown in Fig. 5(a) and (b), the interior of the nanoflower CuCo2S4 presents an open porous structure, which is consistent with the SEM results. The HRTEM image shown in Fig. 5(c) observed lattice fringes with a lattice spacing of 0.285 nm, corresponding to the (113) crystal plane of the nanoflower CuCo2S4. Additionally, Fig. 5(d) shows the SEM cross-sectional view of the nanoflower CuCo2S4 film obtained by screen printing technology. It can be clearly seen that the morphology of the film is relatively dense and the thickness is about 2.80 μm. Furthermore, the film is in close contact with the FTO substrate, which can reduce the series resistance (Rs) of the cell, thereby improving the photovoltaic performance of the QDSSCs.
![]() | ||
| Fig. 4 SEM images of (a) flower-like; (b) nanosheet-like; (c) nanoparticle-like; (d) microsphere-like CuCo2S4. | ||
![]() | ||
| Fig. 5 (a and b) TEM images; (c) HRTEM images; (d) SEM cross-sectional view; (e–h) EDS element distribution images of flower-like CuCo2S4. | ||
Elemental mapping scanning and EDS were used to further analyze the elemental composition of the CuCo2S4 material. The EDS mapping scan results of the CuCo2S4 are depicted in Fig. 5(e)–(h), and the Cu, Co and S elements are evenly distributed in the sample. Besides, the signal peaks of the EDS spectrum (Fig. S3) indicate the presence of Cu, Co and S elements in all the samples, and each of the elemental analysis report shows that the atomic ratio of Cu, Co and S is close to its stoichiometric ratio. This corresponds to the EDS mapping and further proves the successful synthesis of the CuCo2S4.
Analyzing the specific surface area and pore structure of the synthesized materials is of great significance for evaluating their performance. As revealed in Fig. 6, the specific surface area and pore structure of the materials were characterized by N2 adsorption–desorption isotherms. Apparently, the CuCo2S4 materials with different morphologies all showed IV-type adsorption isotherms accompanied by H3-type hysteresis loops, confirming the existence of mesoporous structures in the materials. The rich mesoporous structure can significantly optimize the catalytic process of the counter electrode, promote the penetration of the electrolyte, and thus improve the overall performance of QDSSCs. The detailed BET data obtained are summarized in Table S1. It can be seen that compared with CuCo2S4 materials with other morphologies, f-CuCo2S4 has the largest specific surface area (SBET = 44.86 m2 g−1). In the catalytic reduction process, the large specific surface area can provide more catalytic active sites. Therefore, QDSSCs composed of f-CuCo2S4 counter electrodes exhibit excellent photovoltaic performance.
To further explore the electron transfer ability and electrocatalytic activity of different CEs, EIS tests were performed on different CE materials using a symmetrical cell model in the frequency range of 10−1–105 Hz and an amplitude of 10 mV. According to the equivalent circuit diagram in Fig. 7(b), the curves after fitting with Z-SimDemo software are displayed in Fig. 7(a) (the inset is an enlarged curves view of f-CuCo2S4 and n-CuCo2S4). Obviously, the Nyquist plots of different CEs are composed of two semicircles. Therein, in the semicircle located in the high-frequency region, the starting point of the semicircle is usually defined as the series resistance (Rs), and the radius of the small semicircle represents the charge transfer resistance (Rct), which can be used to assess the charge transfer process at the CE/electrolyte interface. Besides, the semicircle in the low-frequency region is generated by the diffusion of the electrolyte, which is called the diffusion resistance (Rw). The relevant fitting parameters of various CEs are summarized in Table 1. Generally, the smaller the Rs, the higher the electron transfer efficiency.40 The electrochemical performance parameters in Table 1 demonstrate that the Rs values of f-CuCo2S4, n-CuCo2S4, p-CuCo2S4, and m-CuCo2S4 CEs are 2.35 Ω, 2.77 Ω, 2.94 Ω, and 3.30 Ω, respectively. Clearly, the f-CuCo2S4 CE exhibits the lowest Rs value, which indicates that the f-CuCo2S4 electrode has the best electron transport performance and electrocatalytic activity. Moreover, among the different CEs, the f-CuCo2S4 shows the smallest Rct value (0.076 Ω). Normally, the Rct value is negatively correlated with the electrocatalytic performance of the CEs. The smaller the Rct, the higher the electrocatalytic activity and conductivity of the CEs. Therefore, the f-CuCo2S4 electrode presents optimal catalytic performance. Simultaneously, the corresponding Bode curves of different CEs are plotted in Fig. 7(b). The electron lifetime (τe) at the CE/electrolyte interface can be obtained by the formula τe = 1/2πfmax, where τe represents the time required for Sn2− to be reduced to S2−, and fmax is the maximum characteristic peak frequency in the low-frequency region. Apparently, the τe of various CEs can be estimated as f-CuCo2S4 > n-CuCo2S4 > p-CuCo2S4 > m-CuCo2S4, meaning that the f-CuCo2S4 CE has the fastest catalytic reduction reaction rate for Sn2−, proving that compared with other CuCo2S4 CEs, the f-CuCo2S4 CE has the most outstanding catalytic activity.
| Counter electrode | Rs (Ω) | Rct (Ω) | Rw (Ω) | J0 (mA cm−2) |
|---|---|---|---|---|
| f-CuCo2S4 | 2.35 | 0.076 | 9.78 | 5.37 |
| n-CuCo2S4 | 2.77 | 0.383 | 11.36 | 2.40 |
| p-CuCo2S4 | 2.94 | 0.514 | 41.42 | 1.51 |
| m-CuCo2S4 | 3.30 | 0.730 | 77.87 | 1.01 |
Tafel test was used to further investigate the electrocatalytic performance of different CEs. Fig. 7(c) depicts the Tafel polarization curves of CEs with different morphologies, which can reflect two important parameters: the exchange current density (J0) proportional to the catalytic reaction rate and the limiting current density (Jlim) proportional to the ion diffusion coefficient in the electrolyte. The value of J0 can be obtained by the intercept of the cathode branch and the zero potential intersection. Jlim can be attained by the intercept of the vertical axis. The detailed values are summarized in Table 1. Obviously, the values of J0 (5.37 mA cm−2) and Jlim of the f-CuCo2S4 CE are greater than those of the CEs with other morphologies, which indicates that the f-CuCo2S4 CE has higher electrocatalytic activity and faster electrolyte diffusion rate. And the larger the J0 value, the better the corrosion resistance of the CE in the polysulfide electrolyte. Therefore, from the value of J0, it can be judged that f-CuCo2S4 CE has better corrosion resistance than n-CuCo2S4 (2.40 mA cm−2), p-CuCo2S4 (1.51 mA cm−2) and m-CuCo2S4 (1.51 mA cm−2). The above conclusions are consistent with the analysis results of EIS and Bode curves. In addition, to further illustrate the advantages of f-CuCo2S4 CE, the conductivity (G) of different CEs was characterized by linear sweep voltammetry (LSV), and the results are shown in Fig. 7(d). According to the formula G = 1/R = I/V, the G of each CE can be obtained by fitting the slope of the curve. After calculation, the G values of f-CuCo2S4, n-CuCo2S4, p-CuCo2S4 and m-CuCo2S4 are 0.035 S, 0.021 S, 0.017 S and 0.012 S, respectively. Distinctly, the G value of f-CuCo2S4 CE is significantly higher than that of other CEs, suggesting that the nanoflower-like morphology can more effectively assist the CE to transfer electrons during the operation of QDSSCs, enhance the conductivity of the CE, and thus facilitate the photovoltaic performance of the cell.
To better evaluate the ability of the counter electrode to catalyze the reduction of Sn2− in polysulfide electrolytes, CV tests were performed on different CEs in an electrolyte consisting of 2 M Na2S, 2 M sulfur powder and 0.2 M KCl, employing Pt, saturated calomel electrode (SCE) and CuCo2S4 with different morphologies as the CE, reference electrode and working electrode, respectively. It is generally recognized that a larger reduction peak current density (JRed) and a smaller peak-to-peak separation (Epp) contribute to promote the catalytic reduction of CEs. As depicted in Fig. 8, the values of the cathodic peak current density of CuCo2S4 CEs with different morphologies follow the order of f-CuCo2S4 > n-CuCo2S4 > p-CuCo2S4 > m-CuCo2S4. Additionally, f-CuCo2S4 CE exhibits the smallest Epp value, indicating that f-CuCo2S4 CE has better catalytic activity in polysulfide electrolytes than other CEs. Simultaneously, to further study the stability and corrosion resistance of different CEs in polysulfide electrolytes, multi-cycle continuous CV scan tests were carried out using a symmetric cell model. It can be clearly seen from Fig. 9 that after 50 cycles of scanning, f-CuCo2S4 showed better repeatability compared with n-CuCo2S4, p-CuCo2S4 and m-CuCo2S4 CEs. Besides, f-CuCo2S4 CE also showed the highest current density value, and its attenuation was negligible. However, the current density of n-CuCo2S4, p-CuCo2S4 and m-CuCo2S4 CEs all revealed a significant downward trend. The above results demonstrate that f-CuCo2S4 CE has relatively outstanding stability and electrocatalytic activity in polysulfide electrolytes.
![]() | ||
| Fig. 9 Continuous CV curves of symmetric cell model based on (a) f-CuCo2S4; (b) n-CuCo2S4; (c) p-CuCo2S4; (d) m-CuCo2S4 counter electrodes. | ||
Under AM 1.5G, 100 mW cm−2 simulated sunlight, different counter electrodes and FTO/TiO2/CdS/CdSe/ZnS photoanodes were assembled into “sandwich” type complete QDSSCs and the photovoltaic performance was characterized. To ensure the stability and repeatability of the experimental data, 10 QDSSCs were used for each test, and the detailed photovoltaic performance parameters (Jsc, Voc, FF, PCE) obtained are listed in Table 2. Fig. 10 depicts the photocurrent density–voltage (J–V) curves of QDSSCs equipped with different CEs. As is evident from the figure, the performance parameters of the QDSSC based on f-CuCo2S4 CE are enhanced (Jsc = 25.42 mA cm−2; Voc = 0.596 V; FF = 0.49; PCE = 7.42%), especially the PCE. Compared with QDSSCs assembled with n-CuCo2S4 (PCE = 5.98%), p-CuCo2S4 (PCE = 5.33%) and m-CuCo2S4 CEs (PCE = 4.78%), the PCE of QDSSCs based on f-CuCo2S4 CE increased by about 24%, 39% and 55%, respectively, which was mainly attributed to the large specific surface area and open three-dimensional porous structure of f-CuCo2S4, which increased the active sites for catalytic reduction reactions, promoted the effective penetration of electrolytes, and thus accelerated the electron transfer process. Simultaneously, the corresponding error analysis diagram of the PCE of QDSSCs based on four CEs was depicted, and 10 cells were selected as a group for each CE for testing. Fig. S4(a) shows that the average PCE of QDSSCs based on f-CuCo2S4 CE is 7.40%, demonstrating relatively good reproducibility and stability. For comparison, we also tested the photovoltaic performance of QDSSCs based on conventional Cu2S and Pt CEs under the same test conditions. The test results in Fig. S5 and Table S2 show that the PCE of QDSSCs based on f-CuCo2S4 CE is greatly improved compared with Cu2S and Pt CEs. Additionally, the incident photon-to-electron conversion efficiency (IPCE) was used to further test the photovoltaic performance of the cells. As shown in Fig. S6, the QDSSCs have a light absorption range of 400–700 nm, and the QDSSCs based on f-CuCo2S4-CE exhibited the highest IPCE. Furthermore, the integrated short-circuit current density (Jsc) values calculated from the IPCE are listed in Table S4. Clearly, they agree well with the Jsc measured from the J–V curves.
| Counter electrode | Jsc (mA cm−2) | Voc (V) | FF | PCE (%) |
|---|---|---|---|---|
| f-CuCo2S4 | 25.42 | 0.596 | 0.49 | 7.42 |
| n-CuCo2S4 | 22.92 | 0.585 | 0.45 | 5.98 |
| p-CuCo2S4 | 21.58 | 0.592 | 0.42 | 5.33 |
| m-CuCo2S4 | 20.83 | 0.587 | 0.32 | 4.78 |
For further research the photovoltaic performance of QDSSCs based on various CEs, the charge recombination phenomenon at the photoanode/electrolyte interface of different QDSSCs was characterized by open circuit voltage decay (OCVD) test. First, a stable open circuit voltage was obtained under standard sunlight, and then the light source was turned off to further obtain the test curves, which are presented in Fig. 11(a). Researchers generally agree that a slower open circuit voltage decay implies a slower charge recombination rate and a longer electron lifetime. Therefore, it can be seen from Fig. 11(a) that the QDSSC based on f-CuCo2S4 CE exhibits the slowest decay rate, which indicates that its excellent electrocatalytic activity inhibits the charge recombination at the photoanode/electrolyte interface, thereby improving the conversion efficiency of QDSSCs. Moreover, the startup rate of QDSSCs devices is crucial for the practical application of the device. The light response speed of diverse QDSSCs was studied by on-off experimental measurement, and the results are presented in Fig. 11(b). Obviously, under illumination, the current density of all QDSSCs showed a rapid increase and then stabilized. In the dark state, the current density of the cell displayed a rapid decline. More critically, compared with other QDSSCs, the QDSSCs equipped with f-CuCo2S4 CE revealed the highest current density, and its current density showed a small attenuation trend after 10 on–off experimental tests. This shows that QDSSCs based on CuCo2S4 CE have a strong sensitivity to light, and QDSSCs with f-CuCo2S4 CE have the lowest probability of charge recombination at the photoanode/electrolyte interface.
Reproducibility tests were conducted to further assess the photovoltaic performance of QDSSCs integrated with distinct CuCo2S4 CEs under simulated AM 1.5G solar irradiation. Fig. 12(a–d) illustrates the efficiency statistical histograms of QDSSCs based on f-CuCo2S4, n-CuCo2S4, p-CuCo2S4, and m-CuCo2S4 CEs, respectively, with 20 cells used as samples in each test. Notably, QDSSCs incorporating various CEs exhibit superior reproducibility, and the devices assembled with f- CuCo2S4 CE show higher PCE. All the above analysis results reveal that the nanoflower-like hierarchical architecture of f-CuCo2S4 is beneficial to the electrocatalytic activity, and the large specific surface area and open three-dimensional porous structure help to enhance the stability of the electrode, thereby improving the photovoltaic performance and repeatability of QDSSCs.
![]() | ||
| Fig. 12 Efficiency statistics histogram of QDSSCs equipped with (a) f-CuCo2S4; (b) n-CuCo2S4; (c) p-CuCo2S4; (d) m-CuCo2S4 counter electrodes. | ||
Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra06421j.
| This journal is © The Royal Society of Chemistry 2025 |