Shixun Wang
and
Jianjun Tian
*
Institute of Advanced Materials Technology, University of Science and Technology Beijing, Beijing, 100083, China. E-mail: tianjianjun@mater.ustb.edu.cn
First published on 16th September 2016
Over the past few decades, various types of solar cells provided alternative ways for solar energy conversion. Among them, quantum dot-sensitized solar cells (QDSCs) have gained significant interest due to the advantages of quantum dots (QDs) including easy fabrication, multiple exciton generation, band-gap energy controllability and high absorption coefficient. A QDSC consists of a metal oxide photoanode, QDs, electrolyte and a counter electrode (CE). In comparison with the photoanode and QDs, the CE has not been paid much attention. As an essential part of QDSCs, the CE plays an important role in the charge transport and collection of the device. Here, the recent progress in the development of CEs is reviewed, and the key issues for the materials, structures and performance evaluation of CEs are also addressed.
The representative structure of QDSC consists of a metal oxide semiconductor (MOS) photoanode (usually TiO2 or ZnO), QDs (sensitizer), an polysulfide electrolyte and a CE, as shown in Fig. 1(a).27 Owing to an enlarged surface area to load massive sensitizer for the enhancement of light absorption and provide electrically smooth pathways for electrons transferring to the end of photo-electrode, nanostructured MOS usually used as the supporting frame of QDs. Commonly, metal oxide semiconductors such as TiO2 and ZnO are extensively used in QDSCs.28–30 The MOS photoanode is fabricated by coating nanostructured wide bandgap semiconductor with an optimum thickness about 11 μm on the conducting glass.31 Then, the quantum-dot sensitizers are adsorbed onto the surface of MOS to harvest photons. The widely used electrolytes in DSCs are organic electrolytes with I−/I3− redox couple. However, I−/I3− redox couple is corrosive to most QDs and leading to the deterioration of QDSCs performance. Thus, it is critical to explore an appropriate non-corrosive electrolyte for stable QDSCs. Electrolyte with polysulfide redox couple (S2−/Sx2−) is introduced to provide a stable environment for QDs. And the composition of polysulfide electrolyte was tailored by methanol to improve penetration of the electrolyte into photoanode.32,33 Quasi-solid-state polysulfide electrolyte with the assistant of polymer was also reported to improve the cell stability.34,35 In order to improve the performance of QDSCs, several electrolytes have been reported based on different redox couple such as the Fe2+/Fe3+, Fe(CN)63−/Fe(CN)64−, Co2+/Co3+ and (CH3)4N2S/(CH3)4N2Sn.34–42 QDSCs assembled with electrolytes based on the redox system above got relative high open circuit voltage than cells packaged polysulfide electrolyte. However, QDSCs employing polysulfide electrolyte have shown the highest short circuit current and power efficiency among them.43,44 A Pt-based CE has been widely studied in a typical DSCs device due to its excellent catalytic ability to reduce I3− to I−.45 Nevertheless, Pt-based CEs are unsuitable in a polysulfide based QDSC device for their strong chemisorption with polysulfide redox couple (S2−/Sx2−), leading to a reduced catalytic activity of CEs.46–48 Recently, the copper chalcogenides are used as CEs for the highly efficient QDSCs due to excellent catalytic performance and good compatibility with polysulfide electrolyte.49,50
![]() | ||
Fig. 1 (a) Schematic illustration of a typical QDSC device; (b) interfacial charge transfer processes following a laser pulse excitation. Reproduced with permission from ref. 27. Copyright 2015 American Chemical Society. |
Fig. 1(b) shows photoinduced charge transfer process. When the cell is exposed to continuous illumination, QDs will absorb light within a certain range of wavelengths based on their band gap energy and (①) excite electrons from the valence band (VB) of QDs to their conduction band (CB). (②) Electrons in the CB of QDs are injected into the CB of MOS, and then transferred along the TiO2 nanocrystallites and the external circuit to the CE. At the same time, (③) the holes are released by redox couples in the electrolyte. (④) The oxidized electrolyte formed as the oxidation proceeds is reduced at the CE. Eventually, the reduced electrolyte will be transported to the photoanode to combine with the holes again.27,51 With the driving force provided by the energy-level difference between the aligned Fermi levels at the QDs/MOS and the redox potential of the electrolyte, the electrons are theoretically able to circulate the cell over and over again. However, (⑤) recombination of electrons from QDs and the oxidized form of the redox couple, and (⑥) interfacial recombination of electrons from MOS (TiO2 in Fig. 1(b)) and the oxidized form of the redox couple will be a major factor in limiting the overall efficiency. Thus, many researchers devote themselves to the modification of the photoanode to reduce undesired recombination. A tremendous amount of work has been done to optimize the performance of QDSCs including exploiting various QD sensitizer,25,26,52–54 electrolytes,32,55,56 and counter electrodes (CEs).50,57–59 Among them, CEs have been the least studied.
Actually, CE plays an extremely important role in the structure of QDSC to collect electrons from the external circuit and actuate catalysis of the reduction of oxidized species in electrolyte. That is, CE should catalyze the regeneration of the redox in electrolyte after its action to take back the oxidized sensitizers to the ground state and thus keep the device active and stable. Generally, CEs with high catalytic activity, large surface area, chemical durability and low production cost are indispensable to enhance the performance of QDSC. Besides, the conductivity of CEs corresponding to the kinetics of charge transfer at the CE/electrolyte interface will significantly affect the fill factor (FF) and short circuit current. Thus, the research of alternative catalytic materials with high conductivity to substitute conventional Pt-based CE is an urgent way to improve the conversion efficiency of QDSCs. In brief, the optimization of CEs with developed photoanode may contribute to a prominent result in increasing PCE and stability of QDSCs in the foreseeable future. This short review will concentrate on an overall retrospection of QDSCs counter electrode.
![]() | ||
Fig. 2 (a) The schematic diagram of synthesis of (Pt NPs, Au NPs, and AuPt-BNPs) CEs using DPR method and a QDSC with a CdSe/CdS/ZnO-nano wire photoanode and NPs/FTO CE; (b) the transmission spectra in a range of 300–800 nm of the different CEs; (c) current density–voltage (I–V) characterization of QDSC assembled with different CEs. Reproduced with permission from ref. 60. Copyright 2014 Elsevier B.V. All rights reserved. |
![]() | ||
Fig. 3 (a) The SEM images of PEDOTHDC and PEDOTLDC CEs; (b) I–V characterization of CdS based QDSCs assembled with different CEs. Reproduced with permission from ref. 68. Copyright 2014 Elsevier Ltd. All rights reserved. Attention: the curves in (b) should exchange the colors with each other. |
![]() | ||
Fig. 4 I–V characterization of QDSC assembled with CNT-based CE and CuS CE. The schematic diagram of QDSC assembled with CNT-based CE is inserted. Reproduced with permission from ref. 78. Copyright 2016 Elsevier B.V. All rights reserved. |
Gopi et al. take a series of researches to investigate the reason of the superior performance of CNT-based CE. As shown in Fig. 5, the cyclic voltammetry (CV) was firstly carried out. The negative and positive peaks correspond to the reduction of Sn2− and the oxidation of nS2−, respectively. The CE in the QDSCs serves as a catalyst for reducing the redox couple in the electrolyte. Thus, the cathodic reduction peak current in CV curves can be applied to evaluate the electrocatalytic ability of the CEs. From Fig. 5(a), CNT-based CE obtained the largest cathodic reduction peak current, which indicates the superior performance of it. Besides, the peak-to-peak separation (Epp) value should also be taken into consideration. CNT-based CE has lower value of Epp (1.16 V) than CuS CE (1.27 V). It can mainly be attributed to the large surface area which supplied enough seepage sites for electrolyte into the carbon nanotube layer of CNT-based CE besides the good conductivity, electrocatalytic ability and corrosion–inertness towards polysulfide electrolyte. Accordingly, CNT-based CE has better electrochemical-catalytic activities in comparison of conventional CuS CE.
![]() | ||
Fig. 5 (a) Cyclic voltammetry (CV) of polysulfide redox couple on CNT, CuS and Pt CEs; (b) enlarged part of the CNT and CuS electrodes CV plots. Reproduced with permission from ref. 78. Copyright 2016 Elsevier B.V. All rights reserved. |
Cu2S, a p-type semiconductor, with bandgap of 1.1–1.4 eV has been widely used in QDSCs as CE materials. Generally, Cu2S CEs are fabricated by dipping the etched brass foils into polysulfide electrolyte to form compact Cu2S film.13 But, the poor stability and relatively low surface of the Cu2S compact film still affect its performance as a CE. On account of those problems, novel Cu2S CE preparation methods were developed in order to achieve CE with better stability, conductivity and catalytic activity. Meng et al. applied electroplating method to deposit a thin layer of copper on the surface of FTO and then immersed the copper coated glass in polysulfide solution to prepare Cu2S CE.87,88 As shown in Fig. 6, the final flower-like Cu2S film has a thickness between 100 and 200 nm. The favorable porous structure makes the film better catalytic ability. Fig. 7 shows that the interface charge transfer resistance of electrolyte and Cu2S film is calculated to be a relatively low value of 5.7 Ω cm2 and the CdS based QDSC assembled with this CE obtained a promising PCE of 2.6%.
![]() | ||
Fig. 6 SEM images of (a) bare FTO glass, (b) copper coated FTO glass, (c) Cu2S coated FTO glass and (d) cross-sectional SEM image of Cu2S coated glass. Reproduced from ref. 88 with permission from the Royal Society of Chemistry. |
![]() | ||
Fig. 7 (a) Nyquist plots of symmetric cells based on different CEs; (b) I–V curves of cells assembled with different kinds of CEs. Reproduced from ref. 88 with permission from the Royal Society of Chemistry. |
According to the idea of increasing the surface area of copper sulfides to provide more active catalytic sites through novel preparation method, copper sulfide with various morphologies were reported including porous hierarchical structure,89 nanoplates,90 nanosheets,91 nanoparticals92 and so on. Hu et al. reported a novel ITO@Cu2S based CE for QDSC with the ability of divergent thinking.93 It is another way to increasing the contact area between electrolyte and CE. The preparation process is shown in Fig. 8: (a) gold NPs coated FTO glasses were firstly prepared as catalysis via a sputtering coater and then (b) high-purity metallic indium and tin powders were inducted to synthesis ITO nanowire arrays; (c) chemical bath deposition (CBD) technique was used to grow CdS shell on the surface of ITO nanowires; (d) a solution based cation exchange was then carried out to convert CdS into Cu2S; (e) the obtained ITO@Cu2S nanowire arrays was used as CE to assemble with CdS/CdSe/ZnS photoanode. The formation of high-quality tunnel junctions with short carrier transport path (<100 nm) between core and sheath significantly reduced the sheet resistant of CE and enhanced electrons transfer from ITO to Cu2S. The CdS/CdSe based QDSCs assembled with ITO@Cu2S CE which have ITO nanowires of 8–10 μm achieved enhanced PCE of 4.06% in comparison with Au-based CE (2.20%).
![]() | ||
Fig. 8 (a) Scheme for preparing ITO@Cu2S nanowire arrays on FTO glass; (b) SEM image of ITO nanowires; (c) low-magnification and (d) high-magnification SEM images of ITO@Cu2S. Reproduced with permission from ref. 93. Copyright 2014 American Chemical Society. |
Hu et al. further developed the preparation method of ITO@Cu2S CE.94 Hierarchically assembled ITO@Cu2S nanowire arrays were fabricated as efficient CdSexTe1−x based QDSCs CEs. The main differences of the preparation process lie in the following aspects: the hierarchical ITO nanowire arrays were prepared by repeating the procedure (a) two times (ITO-II) or three times (ITO-III). And the performance of QDSCs assembled with ITO-I@Cu2S, ITO-II@Cu2S, and ITO-III@Cu2S CE were throughout investigated in their work. As shown in Fig. 9, the smaller third-generation branches of ITO-III were sequentially grown on the surfaces of the second-generation branches of ITO-II via an additional (chemical vapor deposition) CVD process. After the CBD and cation exchange process, the smooth surfaces of ITO nanowire arrays' stems and branches became rough. The length of ITO nanowire stem and the branches of ITO-II and ITO-III were 8–10, 4–5, and 0.5–1 μm. As depicted in Fig. 10(a), the QDSC with ITO@Cu2S nanowire CEs exhibited higher short circuit current density (Jsc) and open circuit voltage (Voc) than that with the brass/Cu2S CE. The monochromatic incident photo-to-electron conversion (IPCE) measurements in Fig. 10(b) indicated the enhanced IPCE between wavelengths of 450–700 nm for QDSCs with ITO@Cu2S CEs compared with the QDSCs assembled with brass/Cu2S CE. The increased Jsc can be attributed to the improved utilization of the incident light because of the light scattering caused by the three-dimensional nanowire arrays of the ITO@Cu2S CEs.
![]() | ||
Fig. 9 SEM imaged of hierarchical ITO nanowire arrays before and after the coaxial growth of the Cu2S shell: SEM image of (a) ITO-II, (b) ITO@Cu2S-II, (c) ITO-III, and (d) ITO@Cu2S-III. Reproduced with permission from ref. 94. Copyright 2015 American Chemical Society. |
![]() | ||
Fig. 10 (a) I–V characterization and (b) IPCE spectra of the QDSCs with various ITO@Cu2S and brass/Cu2S CEs. Reproduced with permission from ref. 94. Copyright 2015 American Chemical Society. |
To further evaluate the properties of ITO@Cu2S CEs, resistance–voltage (R–V) measurement were performed. The results of R–V measurement indicate the decrease in the series resistance (Rs) which reduced the loss of photovoltage and the increase in the shunt resistance (Rsh) which reduced the loss of photocurrent of QDSC assembled with brass/Cu2S CE. According to the eqn (1) below, the increase in the light generated current caused the increase in the Voc. Thus, the QDSCs with ITO@Cu2S-III CE achieved increased Voc compared with that with the brass/Cu2S.
![]() | (1) |
To evaluate the catalytic activity of the various ITO@Cu2S CEs, CV measurement were carried out. QDSCs with ITO@Cu2S-III CE obtained highest peak current at reductive potentials which indicated better electrocatalytic ability of it. Besides, 10 times cycling was performed at a scan rate of 50 mV while little change was observed in the curve shape or current density demonstrating the excellent chemical ability of ITO@Cu2S CEs. A Tafel polarization measurement was further conducted on various CEs. The ITO@Cu2S-III CE symmetric cells got the largest exchange current density (J0) which is inversely proportional to Rct. Thus, the smallest Rct was achieved for ITO@Cu2S-III CE indicating the superior catalytic ability. Moreover, ITO@Cu2S-III CE had higher value of the limiting current density (Jlim) among them which shown larger diffusion coefficient in the electrolyte and better catalytic ability. According to the electrochemical impedance spectroscopy (EIS) measurements on symmetrical cells, ITO@Cu2S-III CE exhibited relatively small intrinsic sheet resistance (Rh, 5.05 Ω) and Rct (1.30 Ω). They can be attributed to the formation of conductive networks which enhanced the efficient electronic transmissions in the three-dimensional nanowire array structure as well as to the formation of an effective tunneling junction between the n-type degenerate ITO nanowires and the p-type degenerate Cu2S nanocrystals.93 Therefore, QDSC assembled with ITO@Cu2S-III CE achieved higher PCE of 6.12% compared with brass/Cu2S based QDSCs (5.05%).
However, the electrical conductivity ability and erosion problem caused by polysulfide electrolyte still limit the further development of QDSCs assembled with copper sulfide CE. Recently, researchers report that copper selenide (CuxSe) shows the potential to be the CEs of QDSCs due to the excellent catalytic activity and lower transfer resistance than that of copper sulfide.95–98 In the work reported by Zhong et al.,50 CuxSe nanoparticles were prepared through aqueous solution method and further made into CuxSe pastes to cover on the surface of FTO (to fabricate CuxSe/FTO CE) via screen-printing method. The experimental conditions for the preparation of CuxSe/FTO CE were further optimized through the resultant optimal values of 1:
4 for Cu
:
Se ratio, annealing at 450 °C for 30 min and 2.6 μm for film thickness. The achieved maximum PCE CuxSe/FTO based QDSCs are 6.49%, higher than CuxS/brass CE based cells (6.35%). As depicted in Fig. 11(a), the statistical results show the excellent performance CuxSe/FTO CEs based QDSCs with PECs generally locating at 5.8–6.1%. Besides, Nyquist plots and Tafel polarization obtained using symmetric dummy cells ascertained that the catalytic activity of CuxSe/FTO CE were better than CuxS/brass CE. The outstanding stability of CuxSe/FTO CE based QDSCs were also presented in Fig. 11(d). Devices assembled with CuxSe/FTO still maintained their initial activity due to the stability of CEs. The stability of CE mainly reflects in its corrosion-inertness towards polysulfide electrolyte. In recent years, researchers have devoted themselves to producing CE material with large specific surface area which have the ability to maintain its chemical stability towards the corresponding electrolyte.
![]() | ||
Fig. 11 (a) Statistical results of PCEs for CdSe QDSCs assembled with CuxSe/FTO and CuxS/brass CEs; (b) Nyquist plots of symmetric cells using CuxSe/FTO and CuxS/brass CEs; (c) Tafel polarization curves of symmetric cells used in EIS measurements; (d) high stability of CdSe QDSCs assembled with CuxSe/FTO CE characterized by PCE and photocurrent. Reproduced with permission from ref. 50. Copyright 2016 Elsevier Ltd. All rights reserved. |
Unlike the aqueous solution method introduced above, our group reported a Cu3Se2 nanostructured CE which directly grow on FTO to form double-layer morphologies through successive ionic layer adsorption reaction (SILAR) and CBD process (Fig. 12(a)).99 The CBD duration which would directly affect the formation of the double-layer structure was firstly studied. The electrochemical characteristics of Cu3Se2 CE were investigated by EIS analysis and compared with CEs prepared by different CBD hour. In Fig. 12(d), the semicircle corresponds to the electron transfer at the TiO2/QDs/electrolyte interface and transport in TiO2 film (Rct). QDSCs assembled with Cu3Se2-3 h CE achieved the highest Rct value (172.9 Ω) among them. The Bode plots in Fig. 12(e) shown the considerable electron lifetime of QDSCs with Cu3Se2-3 h CE (10.07 ms). The Tafel spectra of symmetrical dummy cells in Fig. 12(f) depicts the Jo value of Cu3Se2-3 h CEs is the highest among them, which means the excellent electrocatalytic ability of them. The measurement results can be ascribed to the high crystallinity, proper crystal defects and large surface area and outstanding electron transfer pathway in counter electrodes.
![]() | ||
Fig. 12 (a) The preparation processes of Cu3Se2; (b) SEM image and (c) cross-sectional SEM image of Cu3Se2-based CE treated by 3 h of CBD process; (d) electrochemical impedance spectra and (e) Bode phase plot of the QDSCs based on the three different CEs under forward bias (−0.6 V) and dark conditions; (f) Tafel polarization curves for the different CEs. Reproduced from ref. 99 with permission from the Royal Society of Chemistry. |
In the system of CdS/CdSe based QDSCs, the performance of Cu3Se2-based CE was further studied in comparison with brass/CuxS based CE. Fig. 13(a) depicted that the device assembled with Cu3Se2 CEs achieved higher fill factor, larger short circuit current and open circuit voltage than that with CuxS CE. The possible reasons are ascribed to the large surface area, proper crystal defects, high conductivity and good electrocatalytic activity of Cu3Se2 nanostructure, which further accelerate the electron transfer and the reaction of redox in polysulfide electrolyte, and finally result in low charge recombination and high conversion efficiency (best PCE of 5.05%). Generally, both approaches above mentioned are effective in case of Cu3Se2 CEs preparation. The performance of QDSC using Cu3Se2 CE has been studied in our work, in which the PCE shown excellent stability in comparison with CuxS/brass based devices. As shown in Fig. 13(b), the PCE of QDSCs with Cu3Se2 CE changed from 4.95% to 4.81% with a negligible decline of about 2.8%. However, the QDSCs assembled with CuS CE shown relatively low stability and their PCE changed from the initial 4.03% to the final 2.24% with an obvious decline of about 44%. It is evident that Cu3Se2/FTO CE which has larger specific surface area shows remarkable stability.
![]() | ||
Fig. 13 (a) J–V curves of QDSCs assembled with different CEs, the inset is the photographs of CuxS and Cu3Se2/FTO CEs; (b) stability of QDSCs assembled with different CEs. Reproduced from ref. 99 with permission from the Royal Society of Chemistry. |
As depicted in Fig. 14, the PbS film outperforms Pt and CuS film even though CuS has a much higher catalytic ability in polysulfide electrolyte than PbS. It can be attributed to the photoactive characteristics of PbS CE which increased the photocurrent of the resulting QDSCs. The PbS film indicated its absorption of the entire range of visible light and even into the IR region to compensate the shortage of photoanode on light absorption. Thus, the PCE of cells assembled with PbS based CE increased due to the marked increment of Jsc and Voc.
![]() | ||
Fig. 14 (a) Schematic energy diagram representing the entire charge-transfer process in the QDSCs assembled with the PbS CE; (b) SEM image of PbS/FTO CE deposited by 5 SISCR cycles; (c) Nyquist impedance plots of different CEs; (d) I–V curves of PbS, CuS and Pt based QDSCs. Reproduced from ref. 80 with permission from the Royal Society of Chemistry. |
Semiconducting metal selenides, as an essential class of chalcogenides, have attracted enormous attention on account of their distinctive electronic properties and conspicuous chemical behaviors. Recently, metal selenides have emerged as a new class of CE electrocatalysts to replace conventional CE materials of QDSCs. Choi et al. performed a throughout investigation on binary metal selenides to assess their feasibility as CE materials for CdS/CdSe based QDSCs.102 Eight different types of binary selenides (MnSe, CoSe2, NiSe2, Cu1.8Se, MoSe2, WSe2, PbSe, and Bi2Se3) were randomly chosen as electrocatalyst candidates for CE materials in their study. And the photoelectric transition property of solar cells assembled with these metal selenides were compared with conventional Pt CE based QDSCs. As shown in Fig. 15, QDSCs assembled with Cu1.8Se or PbSe CEs shows better overall photoelectric property than that with Pt, while inferior performances were obtained when using other CEs according to the I–V curves. The study on Cu1.8Se and PbSe CEs were further carried out to investigate their physical and electrocatalyst properties. Cu1.8S has a considerable stability while PbSe was partly oxidized to PbO and SeO2 in air. Besides, EIS measurement was also performed to approve Cu1.8S CE had better catalytic ability and electrochemical stability when functioning in polysulfide electrolyte in comparison with PbSe CE. PCE of 5.01% was achieved, applying Cu1.8S CE.
![]() | ||
Fig. 15 I–V curves of QDSCs assembled with various metal selenide CEs under 1 sun illumination. Reproduced with permission from ref. 102. Copyright 2014 American Chemical Society. |
It is known to us that carbon composites have the advantages of large surface area, excellent electrical conductivity and relatively low costs. And Mo compounds are widely used as catalysts in scientific research about hydro-processing which proves their excellent electrocatalytic ability.110 Seol et al. successfully combined the advantages of both CNT and RGO and applied the composite in QDSCs as CE.109 The CNT–RGO complex offered an excellent electron pathway due to the high electrical conductivity and a large surface area by preventing stacking of RGO and bundling of CNT. The reported performance of these composite CEs achieved a PCE of greater than 5% assembled with CdS/CdSe/ZnO nanowire photoanode. The research about Mo-compound/CNT–RGO provided the convincing evidence of the clearly enhanced electrocatalytic activities of the CE. The Mo-compound/CNT–RGO complex was fabricated according to a facile method involving a modified urea-glass route.111 In brief, the Mo-compound (Mo2N, Mo2C, and MoS2) anchored onto the CNT–RGO structure could be prepared by the reaction of controlling the application of urea or thiourea precursor and the ratio of the urea precursor to Mo precursor. As shown in Fig. 16(a) and (b), the achieved multiwalled CNTs are randomly distributed on the RGO layer and the Mo compounds are well-spread over the CNT–RGO support. The 1.5 μm-thick mesoporous film covering on the FTO was used as CE for CdS/CdSe based QDSCs containing polysulfide electrolyte. Besides, the charge transfer process that occurred in the Mo compound/CNT–RGO CE is clearly illustrated by Fig. 16(b). Electrons from external circuit are delivered to the Mo compound along the CNT–RGO composite structure and are used to reduce the oxidized electrolyte. The Mo compound here work as the critical factor in facilitating the catalytic reaction, whereas CNT–RGO composite acts as an excellent electron pathway to reveal its outstanding electron conductivity. Moreover, the larger surface area of CNT–RGO complex provided plenty of load sites and contribute to an increased amount of loading onto Mo compound catalysts.
![]() | ||
Fig. 16 (a) Cross-sectional SEM image of Mo2N/CNT–RGO CE; (b) electron-transfer mechanism of the MoX/CNT–RGO CE; (c) I–V curves and (d) photocurrent stability of QDSCs assemble with Au and the Mo-compound/CNT–RGO CE. Reproduced with permission from ref. 109. Copyright 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. |
To evaluate the performance of these novel composite CE, PCE was firstly carry out by using a solar simulator. The resultant I–V curves are presented in Fig. 16(c). Except for the MoS2/CNT–RGO CE based QDSCs, the other QDSCs all shown improved performance compared to the conventional Au CE based QDSCs. QDSCs assembled with Mo2N/CNT–RGO CE achieved the best performance due to the excellent catalytic ability of Mo2N which greatly contribute to the fill factor and thus considered to be the most promising catalyst for use in QDSC as CE material. Moreover, Fig. 16(d) clearly demonstrate QDSCs with Mo2N and Mo2N maintained their photocurrent even after 5000 s of irradiation due to the noble metal-like catalytic behavior. EIS measurements and Tafel polarization tests were performed after the stability measurement to assure the encouraging effect of the introduction of the Mo2N/CNT–RGO CE.
Recently, researchers further developed the application of CNT as CE in QDSCs. Sayan et al. reported a composite of a 15 wt% graphene oxide nanoribbon (GOR) and Cu1.18S composite CE and achieved a record PCE of 3.55% for CdS based QDSCs compared with Pt CE or brass/Cu2S CE.112 In their work, GOR was produced by oxidative cleavage of the commercial MWCNT whereas graphene oxide (GO) was synthesized by the modified Hummers method. I–V curve depicted in Fig. 17 shows the excellent performances of QDSCs assembled with Cu1.18S–RGO CE and Cu1.18S–GO CE which achieved PCE of 3.38% and 3.22%, respectively. Besides, QDSCs with Cu1.18S–RGO CEs obtained higher FF (0.57) compared with cells assembled with brass/Cu2S CEs (0.50). It indicated the superior catalytic ability and conductivity of Cu1.18S–RGO CEs to some degree. To further evaluate the properties of CEs, EIS measurement, Tafel polarization plots and CV analysis were subsequently carried out which shown excellent agreement with the conclusions above. As shown in Fig. 17(d), the Fermi level (EF) of p-type Cu1.18S is close to the valence band and thus will be in equilibrium with the reduction potential of the polysulfide under dark condition. Since the Highest Occupied Molecular Orbital (HOMO) level of GOR matches well with the reduction potential of polysulfide, it closely coincides with the EF of Cu1.18S. In QDSCs, when the electrons move from the photoanode to the CE to reduce the oxidized electrolyte, if the redox potential for electrolyte is close to the HOMO level of GOR, it will greatly facilitate the electrons migration than Cu1.18S itself. Thus, the particular band alignment will reduce the carrier loss and finally enhance PCE. In addition, the HOMO level of GO is little higher above the redox potential of the electrolyte, and therefore the transfer of electrons will be hindered.
![]() | ||
Fig. 17 (a) I–V characterization of CdS based QDSCs assembled with different CEs; (b) GOR synthesized by unzipping CNT; (c) GO; (d) energy band diagram of Cu1.8S–GOR CE showing electron transfer. The energy levels of GO are also indicated. Reproduced from ref. 112 with permission from the Royal Society of Chemistry. |
CE materials | Preparation | Sensitizer | Isc/mA cm−2 | Voc/V | FF | Efficiency (%) | Rct |
---|---|---|---|---|---|---|---|
Pt113 | Thermal evaporation | CdS/CdSe | 16.2 | 0.49 | 0.49 | 3.90 | |
Au66 | Sputtering | CdS/CdSe | 16.8 | 0.51 | 0.49 | 4.22 | |
AuPt NPs60 | Dry plasma reduction | CdS/CdSe | 16.5 | 0.65 | 0.31 | 3.40 | 54.2 Ω cm2 |
Co–Ru67 | Hydrothermal method | CdS | 9.6 | 0.59 | 0.53 | 3.04 | 21.7 Ω |
PEDOT70 | Electro-polymerization | CdS | 3.2 | 0.59 | 0.72 | 1.37 | 123.1 Ω cm2 |
Polypyrrole (PPy)69 | Electro-polymerization | CdS | 3.9 | 0.39 | 0.27 | 0.41 | 337.5 Ω cm2 |
Polythiophene (PT)69 | Electro-polymerization | CdS | 0.9 | 0.33 | 0.26 | 0.09 | 1267.5 Ω cm2 |
TiO2–PEDOT70 | Electro-polymerization | CdS | 5.5 | 0.42 | 0.67 | 1.56 | 4.3 Ω cm2 |
PVP–CuS108 | Solvothermal | CdS/CdSe | 17.5 | 0.58 | 0.51 | 5.22 | 1.4 Ω |
Activated carbon74 | Doctor blade technique | CdS/CdSe | 10.9 | 0.54 | 0.57 | 3.48 | 0.2 Ω cm2 |
CNT78 | Doctor blade technique | CdS/CdSe | 16.0 | 0.58 | 0.49 | 4.67 | 0.8 Ω |
Carbon nanofibers73 | Nanocasting technology | CdSe | 11.9 | 0.62 | 0.60 | 4.81 | 2.6 Ω cm2 |
Fullerene74 | Doctor blade technique | CdS/CdSe | 12.6 | 0.54 | 0.60 | 4.18 | 84 Ω |
Mesocellular carbon foam75 | Carbonization from silica template | CdS/CdSe | 12.6 | 0.68 | 0.42 | 3.60 | |
Hollow core/mesoporous shell77 | Carbonization from silica template | CdSe | 12.4 | 0.60 | 0.52 | 3.90 | 12 Ω cm2 |
N-doped hollow carbon NPs114 | Detonation-assisted chemical vapor deposition (CVD) | CdS/CdSe | 13.5 | 0.51 | 0.40 | 2.67 | 2.0 Ω cm2 |
AuPt NP/reduced graphene nanoplatelets115 | dry plasma reduction | CdS/CdSe | 15.2 | 0.72 | 0.40 | 4.50 | 34.3 Ω |
Carbon dot/Au nanoraspberries107 | Doctor blade technique | CdS/CdSe | 16.6 | 0.70 | 0.46 | 5.40 | 2.0 Ω cm2 |
Mesoporous carbon/Ti25 | Silica template nanocasting/screen printing | Zn–Cu–In–Se | 25.1 | 0.74 | 0.62 | 11.61 | |
PbS/carbon black105 | Solvothermal/ball mill | CdS/CdSe | 13.3 | 0.50 | 0.58 | 3.91 | 10.3 Ω |
FeS/carbon116 | Electrochemical deposition | CdS/CdSe | 20.3 | 0.44 | 0.51 | 4.58 | 8.6 Ω cm2 |
CuInS2/carbon106 | Hot-injected/ball milling | CdS/CdSe | 14.1 | 0.51 | 0.60 | 4.32 | 18.8 Ω cm2 |
graphene/PbS117 | Hummer's method/SILAR | CdS/CdSe | 11.0 | 0.55 | 0.42 | 2.63 | 4.1 Ω cm2 |
RGO–Cu2S48 | Sonication | CdS/CdSe | 18.4 | 0.52 | 0.46 | 4.40 | 1.6 Ω cm2 |
Cu1.18S–GOR112 | Doctor blade technique | CdS/CdSe | 18.4 | 0.52 | 0.58 | 5.42 | 1.3 Ω |
Mo2N/CNT–RGO109 | Modified urea-glass route | CdS/CdSe | 16.9 | 0.68 | 0.47 | 5.41 | 23.8 Ω |
Mo2C/CNT–RGO109 | Modified urea-glass route | CdS/CdSe | 16.4 | 0.67 | 0.44 | 4.84 | 34.7 Ω |
MWCNT–CZTSe118 | Hot-injection/surface modification | CdSe | 17.0 | 0.53 | 0.51 | 4.60 | 2.5 Ω |
CuxS50 | Etching brass foil by HCL solution | CdS/CdSe | 16.1 | 0.62 | 0.63 | 6.35 | 2.1 Ω |
CuxSe50 | Aqueous solution method | CdS/CdSe | 16.0 | 0.64 | 0.62 | 6.49 | 1.3 Ω |
CoS2 (ref. 81) | Thermal sulfidation | CdS/CdSe | 14.4 | 0.51 | 0.56 | 4.16 | 40.6 Ω cm2 |
Co9S8 (ref. 100) | Template assisted method | CdS/CdSe | 16.8 | 0.48 | 0.46 | 3.72 | 2.3 Ω cm2 |
FeS47 | Sulfidation of carbon steel | CdS/CdSe | 20.4 | 0.42 | 0.40 | 3.34 | 8.9 Ω |
PbS80 | SILAR | CuInS/CdS | 18.3 | 0.58 | 0.45 | 4.70 | 30 Ω cm2 |
PbSe102 | SILAR | CdS/CdSe | 16.7 | 0.59 | 0.48 | 4.71 | 36.3 Ω |
NiSe2 (ref. 102) | Successive ionic layer adsorption and reaction | CdS/CdSe | 14.3 | 0.48 | 0.52 | 3.63 | |
NiCo2S4 (ref. 82) | Chemical synthesis | CdS/CdSe | 13.1 | 0.52 | 0.49 | 3.30 | 2.9 Ω |
Cu2SnSe3 (ref. 84) | Solvothermal | CdSe | 16.3 | 0.56 | 0.54 | 4.93 | 0.3 Ω cm2 |
Cu2ZnSnSe4 (ref. 85) | Hot-injected method | CdSe | 15.5 | 0.54 | 0.52 | 4.35 | 8.9 Ω |
FeS/nickel foam101 | Electrochemistry deposition | CdS/CdSe | 18.7 | 0.41 | 0.58 | 4.39 | 2.0 Ω cm2 |
NiS/PbS104 | CBD | CdS/CdSe | 14.5 | 0.59 | 0.54 | 4.52 | 10.1 Ω |
ITO@Cu2S92 | CVD/CBD/cation exchange | CdSeTe | 15.2 | 0.68 | 0.58 | 6.12 | 1.3 Ω cm2 |
SKE–Cu7S4 (ref. 119) | Template assisted method | CdS/CdSe | 18.5 | 0.51 | 0.47 | 4.43 | 2.0 Ω cm2 |
![]() | (2) |
R = −(dJ/dV)−1 | (3) |
![]() | ||
Fig. 18 (a) The typical I–V curve and (b) the R–V curve of a solar cell with related parameters. The inset is the equivalent circuit. |
The decrease in the value of Rs will reduce the loss of the photovoltage, and the increase in the value of Rsh will reduce the loss of photocurrent, both of which will enhance the value of fill factor (FF). Moreover, the larger value of Rsh will cause an increase in the Voc according to the eqn (4) below:
![]() | (4) |
Furthermore, I–V measurement is not a rigorous research method of the performance of CEs but accurate characterization of the overall performance of solar cells. The most significant analytic methods for investigating the electric and chemical characterization of CEs are briefly introduced below.
![]() | ||
Fig. 19 Nyquist plot obtained from the EIS measurement of symmetric cells. CPE: constant phase element. |
The Nyquist diagram, which is widely used to represent the measurement results of EIS, features typically two semicircles that in the order of increasing frequency are attributed to the Nernst diffusion within the electrolyte, the electron transfer at the QDs/electrolyte interface and the redox reaction at the CE. With the EIS measurement of QDSCs, we can indirectly get the property of CE. In order to directly study the performance of CE, the concept of symmetric cell ought to be brought up. In the system of symmetric cell, various interfacial and internal impedances of photoanode are removed. The existent impedances in the symmetric cell are the charge-transfer impedance at electrolyte/CE interface and the Nernst diffusion impedance (ZN) of the redox couple in the electrolyte. Thus, in the Nyquist diagram, the high-frequency feature is attributed to the charge-transfer resistant (Rct) which provides the most concrete criterion for CEs' performance. The value of Rct which is calculated by the radius of the first semicircle can present the properties of CE.122 And CE with low Rct value means that charge transfer occurs more easily. However, ZN which originated from the diffusion the redox couple in electrolyte is rarely concerned with the performance of CE.
As depicted in Fig. 20, the applied voltage (x axis) and the logarithmic form of current density make the Tafel curve. Theoretically, the Tafel curve can be divided into three different zones. The curve at low potential curve (|U| < 120 mV) is attributed to the polarization zone, the curve at middle potential (with a sharp slope) is attributed to the Tafel zone and the curve at high potential can be attributed to the diffusion zone. Among them, Tafel zone and diffusion zone which shows a horizontal shape are closely relate to the catalytic ability of CE. In the investigation of electrochemical catalytic ability, two critical parameters are required: J0 and Jlim.
J0 here refers to current density at equilibrium which means the oxidized and reduced electrolyte can simultaneously transfer electrons to the electrodes. Thus, the state of equilibrium occurs when the overpotential is zero. Besides, the value of J0 can obtained from the intersection point of the tangent of cathodic branch in the Tafel zone and the equilibrium potential line. The relationship between J0 and Rct can be formulated clearly by eqn (5):
![]() | (5) |
![]() | (6) |
CV is an efficient tool to evaluate electrode kinetics.126 CV is performed by using an electrode connected to a dynamic potentiostat and immersed in electrolytes that contain the redox couple. Then, the redox reaction occurs at the interface between electrolyte and CE in the measurement process. A resultant CV curve is obtained as shown Fig. 21. The reduction process occurs from (a) the initial potential to (d) the switching potential. In this region the potential is scanned negatively to cause a reduction. The resulting current is called cathodic current (Ipc). The Epc, named cathodic peak potential, is reached when all of the oxidized electrolyte at the surface of the electrode has been reduced. After the switching potential has been reached (d), the potential scans positively from (d) to (g). The resulting current is called anodic current (Ipa). The Epa, named anodic peak potential, is reached when all of the reduced electrolyte at the surface of the electrode has been oxidized.
Thus, the cathodic and anodic peak position indicate the rate of charge transfer at the interface of electrolyte/CE. In brief, the lower value of peak-to-peak separation (Epp), the better the CE is at facilitating charge-transfer characteristics under a certain fixed scan rate. Besides, the intensities of the cathodic/anodic peak currents should be evaluated. CE with higher peak current means there is more electrons exist and participate in the redox reaction at the surface of electrode. That is, higher peak current indicates better electrocatalytic ability of CE. And the stability of CE can also be estimated according to the degree of degradation of the curves by repeating the cycles of CV measurement.
This journal is © The Royal Society of Chemistry 2016 |