Strontium titanate nanoparticles as the photoanode for CdS quantum dot sensitized solar cells

Cong Chen, Qilin Dai*, Chuang Miao, Lin Xu and Hongwei Song*
Jilin University, State Key Laboratory on Integrated Optoelectronics, College of Electronic Science and Engineering, Changchun, China. E-mail: qilin_dai@yahoo.com; songhw@jlu.edu.cn

Received 8th October 2014 , Accepted 8th December 2014

First published on 8th December 2014


Abstract

SrTiO3 has the potential to be used as the photoanode for quantum dot sensitized solar cells (QDSSCs) since it has very similar band structures to TiO2 that exhibits the best performance in QDSSCs. In this work, SrTiO3 nanoparticles (NPs) with a cubic crystal structure were prepared by a hydrothermal method. 120, 70 and 30 nm SrTiO3 NPs were obtained by using different starting materials. CdS quantum dots (QDs) were coated on three different sized SrTiO3 NPs using the successive ion layer adsorption and reaction (SILAR) method. QDSSCs based on three different sized SrTiO3 NPs were fabricated and investigated. Due to the higher surface area compared to larger NPs, the device based on 30 nm SrTiO3 NPs shows the best open-circuit voltage (Voc) of 0.76 V and fill factor (FF) of 67% which are relatively high for Voc and FF reported for CdS-QDSSCs. Comparison studies of light-harvesting efficiency, electron transfer processes and band diagrams were performed to analyze the device performance in the SrTiO3-QDSSCs and TiO2-QDSSCs. The 30 nm SrTiO3 NPs can be combined with TiO2 NPs (P25) to improve the performance of the CdS-QDSSCs with an increasing percentage of 12.5%, which represents an alternative way to realize high efficiencies in QDSSCs.


Introduction

QDSSCs have attracted substantial research interest due to the unique optical and electrical properties of the quantum dots (QDs) as light absorbers for solar cell applications.1–3 It’s believed that QDSSCs have the possibility of circumventing the Shockley–Queisser efficiency limit for solar cells based on p–n junctions.4 In addition, QDs with different sizes also have the capability to match the solar spectrum and improve light absorption. Most recently QDSSCs with efficiencies as high as 8.55% have been achieved.5 Various components in QDSSCs can be optimized in the actual process of assembly, the photoanode material as an important area of research has received extensive attention recently. The photoanodes in QDSSCs are usually binary metal oxides, such as TiO2 or ZnO NPs. In comparison with binary metal oxides, ternary oxides, for example, SrTiO3, exhibit better corrosion resistance and offer more freedom in tuning of optical and electrical properties.6,7

SrTiO3 can be roughly considered as a highly doped anatase TiO2 because they have similar structures.8 SrTiO3 with a perovskite structure contains titanium atoms in 6-fold octahedral coordination, which is similar to the titanium arrangement in anatase TiO2. The two materials have exactly the same band gap, but SrTiO3 has a slightly higher flat band potential, which could lead to higher Voc for the QDSSCs based on SrTiO3 compared to that based on TiO2. With higher flat band potential, SrTiO3 has the capability of photoassisted electrolysis of water in comparison with anatase TiO2.9 Higher electron mobility in SrTiO3 (6.35 cm2 V−1 s−1)10 compared to that in TiO2 (0.1 cm2 V−1 s−1) 11 would be favorable for reducing recombination loss and for electron transport. Thus SrTiO3 is a very promising photoanode in QDSSCs. Burnside et al. evaluated the feasibility of SrTiO3 NPs with size range of 10–60 nm in the application of dye sensitized solar cells.8 However, thus far, there have been few reports on SrTiO3 applications in QDSSCs.

In this work, ternary oxide SrTiO3 NPs with different sizes were prepared by a hydrothermal method. Structural characterizations were carried out to examine the size, and crystalline structure of the as-synthesized SrTiO3 NPs. The QDSSCs based on SrTiO3 NPs and CdS quantum dots were fabricated and characterized. 30 nm SrTiO3 NPs show promising device performance. In addition, SrTiO3 NPs can also be used to improve the device performance of TiO2-QDSSCs due to the high Voc and FF of the devices based on SrTiO3. Thus the first role of SrTiO3 in this work is that it can be used as the photoanode in QDSSCs. The second role is that SrTiO3 NPs can also be used to improve the device performance of TiO2-QDSSCs because TiO2-QDSSCs have the best device performance till now.

Experimental

Synthesis of SrTiO3 NPs

SrTiO3 NPs with different sizes were synthesized by a hydrothermal method following previously published procedures.8,12,13 120, 70 and 30 nm SrTiO3 NPs were prepared using titanium dioxide (P25), tetrabutyl titanate, and titanium bis(ammonium lactate) dihydroxide (TALH) as starting materials, respectively. The experimental details can be found in the ESI.

Solar cell preparation

In the preparation of the SrTiO3 photoanodes, SrTiO3 powders were ground in a mortar with a few drops of glacial acetic acid and sufficient amount of ethanol for 20 minutes. Then the ground SrTiO3 powders were dispersed in 30 mL of ethanol and ultrasonically agitated for 30 minutes to make SrTiO3 paste. The detailed steps of SrTiO3 composite paste preparation procedure are described in the ESI (Fig. S1). After that, the SrTiO3 paste was spread onto 0.5 × 0.5 cm2 (active area 0.25 cm2) conducting FTO glass substrate by the blade method and heated to 150 °C then sintered at 500 °C in a furnace for 30 minutes. The thickness of the photoanode is about 4 μm, which can be found from the cross section scanning electron microscope (SEM) image (Fig. S2). CdS QDs were coated on the SrTiO3 photoanode, using the successive ionic layer adsorption and reaction (SILAR) method following the procedure reported previously.14 Specifically, the SrTiO3/FTO substrate was successively immersed into 0.05 M Cd(NO3)2 in methanol and then into 0.05 M Na2S in methanol for 2 minutes each. Following each immersion, the substrate was rinsed in pure methanol for 2 minutes to remove excessive precursors and then was dried before the next dipping to finish one CdS coating cycle. The above process was repeated to coat the QDs with different SILAR cycles to get four to eight layers of CdS. QDSSCs were fabricated by clamping a Cu2S sputtered FTO glass plate onto a CdS QDs coated photoanode and filling the capillary space with the electrolytes (1 M Na2S + 1 M S, 18 MΩ deionized water). For comparison, the devices based on TiO2 and SrTiO3/TiO2 photoanodes were prepared by the same method.

Measurements

X-ray diffraction (XRD) patterns of all the SrTiO3 powders were recorded on a RigakuD/max 2550 X-ray diffractometer, using a monochromatized Cu target radiation source at a scanning rate of 4° min−1. SEM images were obtained from a SIRION field-emission scanning electron microscope. The transmission electron microscope (TEM) data were measured on a JEM-2010 with a working voltage of 200 kV. JV characteristics of the cells were recorded using a Keithley 2400 source meter and a 1.5 AM, 100 mW cm−2 solar simulator lamp. Incident photon to current conversion efficiency (IPCE) was recorded using a computerized setup consisting of Solar Cell Quantum Efficiency—SolarCellScan100. Absorption spectra were recorded with the Shimadzu UV-1600 Spectrophotometer. Electrochemical impedance spectroscopy (EIS) measurements were performed using an impedance measurement unit (ZAHNER-elektrik IM6) in the frequency range of 0.1–105 Hz, and the applied bias voltage and AC amplitude were set as the open-circuit voltage of the cells and 10 mV between the counter electrode and the working electrode, respectively. The nitrogen adsorption–desorption isotherms were measured at −196 °C with a Gemini VII surface area and porosity system. The specific surface area was estimated by the Brunauer–Emmett–Teller (BET) method.

Results and discussion

Structure and morphology

The crystal structure of the SrTiO3 NPs was characterized by XRD patterns, as shown in Fig. 1. All the XRD patterns of the SrTiO3 NPs can be indexed to the cubic structure (JCPDS card 35-734) with well-defined (100), (110), (111), (200), (210), (211), (220) and (310) diffraction peaks. No other crystalline by-products can be observed in these XRD patterns.
image file: c4ra11960f-f1.tif
Fig. 1 XRD patterns of SrTiO3 NPs with different sizes.

The XRD data of SrTiO3 nanoparticles with different sizes look identical, which is consistent with the literature about 20–60 nm SrTiO3 nanoparticles.8,12 Only sub-10 nm SrTiO3 nanoparticles showed the broad XRD peaks.13 Thus we believe that the XRD peaks may exhibit the features of broad peaks when the size of SrTiO3 nanoparticles is below 10 nm.

Fig. 2 shows SEM and TEM images of the as-synthesized SrTiO3 NPs prepared with different starting materials and the inset shows the size distribution of the corresponding SrTiO3 NPs. The sizes of the NPs distribute in a wide range for all the samples. From Fig. 2a and b it can be seen that as P25 and Sr(OH)2 were chosen as starting materials, the products are yielded as cube shaped NPs. The SrTiO3 NPs have a diameter of 120 ± 32 nm (see inset of Fig. 2a). Fig. 2c shows the SEM image of the SrTiO3 NPs prepared with tetrabutyl titanate, Sr(NO3)2 and PVA. It can be seen that the shapes of the NPs are irregular, and the size of the NPs is 70 ± 8 nm. Fig. 2d shows the corresponding TEM image of Fig. 2c. From the boundary of the 70 nm NPs, it can be observed that they actually consist of a large number of smaller NPs with average size of 10 nm. The size and irregular shape of the as-prepared NPs can be correlated to the additive PVA. As a water-soluble polymer, PVA has been widely used to control the morphology and size in nanomaterial growth.15,16 Fig. 2e and f present, respectively, the SEM and TEM images of the SrTiO3 NPs when TALH and Sr(OH)2·8H2O are used in the hydrothermal reaction. It is obvious that all the as-prepared NPs are composed of well-defined and relatively regular-shaped particles. The size of SrTiO3 NPs is 30 ± 8.5 nm, which is comparable to that of TiO2 (P25).


image file: c4ra11960f-f2.tif
Fig. 2 SEM and TEM images of SrTiO3 NPs with different sizes. (a) and (b) 120 nm SrTiO3 NPs, P25 and Sr(OH)2 were chosen as starting materials; (c) and (d) 70 nm SrTiO3 NPs, synthesized using tetrabutyl titanate and Sr(NO3)2 and PVA; (e) and (f) 30 nm SrTiO3 NPs, prepared using TALH and Sr(OH)2·8H2O.

Device performance based on CdS QDs (four to eight cycles)

Current density vs. voltage (JV) characteristics of the devices sensitized by CdS QDs (four to eight cycles) based on different SrTiO3 NPs (120, 70 and 30 nm) were obtained under AM 1.5 G illumination with a light intensity of 100 mW cm−2 and in the dark, as shown in Fig. 3a–c, respectively. For the QDSSCs based on 120 nm SrTiO3 NPs, as shown in Fig. 3a, the short-circuit photocurrent (JSC) increases from 0.28 to 0.56 mA cm−2 and Voc increases from 0.54 to 0.72 V as the number of coating cycles increases from four to seven. However, JSC and Voc decrease as the number of coating cycles increases to eight. The same trend is observed for the QDSSCs based on 70 and 30 nm SrTiO3 NPs (Fig. 3b and c). JSC increases from 0.52 to 0.94 mA cm−2 and Voc increases from 0.42 to 0.74 V as the number of coating cycles increases from four to seven for the QDSSC based on 70 nm SrTiO3 NPs. JSC increases from 0.76 to 1.53 mA cm−2 and Voc increases from 0.50 to 0.76 V as the number of coating cycles increases from four to seven for the QDSSC based on 30 nm SrTiO3 NPs. All devices exhibit the optimum results in the seventh cycle and appear decreased permanently in the eighth cycles. This phenomenon could be explained by the varied IPCE of the QDSSCs and the varied bandgap with CdS particle sizes, which will be discussed in the following text.
image file: c4ra11960f-f3.tif
Fig. 3 (a) JV characteristics of QDSSCs based on CdS QDs prepared with different coating cycles and 120 nm SrTiO3 NPs. (b) JV characteristics of QDSSCs based on CdS QDs prepared with different coating cycles and 70 nm SrTiO3 NPs. (c) JV characteristics of QDSSCs based on CdS QDs prepared with different coating cycles and 30 nm SrTiO3 NPs.

Fig. 4a shows the IPCE spectra of QDSSCs based on CdS QDs with different coating cycles on 70 nm SrTiO3 NPs, which could be representative for that based on 120 and 30 nm SrTiO3 NPs.


image file: c4ra11960f-f4.tif
Fig. 4 (a) IPCE spectra of QDSSCs based on 70 nm SrTiO3 NPs and CdS QDs coated using different SILAR cycles. (b) Bandgaps of CdS QDs prepared with different coating cycles and their band alignment with SrTiO3.

A monotonic increase can be observed as the incident light wavelength is scanned from 600 to 400 nm without exhibiting distinctive excitonic features. The maximum intensity of IPCE increases when the number of the coating cycles increases from four to seven, however, the maximum intensity of IPCE decreases as the number of coating cycles increases to eight, which is well in accordance with the JV characteristics. The absence of the CdS excitonic features in the IPCE spectra could be explained by the broad size distribution of the SILAR-based QDs.14 The IPCE spectra also show onset wavelengths shift from 448 nm (four-cycle cell) to 540 nm (eight-cycle cell), which can be attributed to the increased QD size and decreased QD bandgap. According to the literature, the conduction band edge of TiO2 is ∼4.2 eV below the vacuum level, the conduction band edge of bulk CdS is about 3.98 eV below the vacuum level.17,18 According to a previous report, SrTiO3 has the same bandgap of 3.2 eV as TiO2 but a conduction band edge about 0.2 eV lower than that of TiO2,8 so the conduction band edge of SrTiO3 is −4.0 eV vs. vacuum. Combining the literature results and IPCE test data, the band diagrams of the SrTiO3 nanoparticles and CdS QDs are shown in Fig. 4b. Size-dependent band gaps of CdS QDs are estimated by the onset wavelengths of the IPCE spectra (which are due to light absorption of CdS QDs19). It’s believed that the decrease in the CdS bandgap is mainly caused by a downward shift of the conduction band edge due to the large effective mass of the hole compared to that of the electron.20,21 From Fig. 4b, it can be observed that when the number of coating cycles increases from four to eight, the conduction band edge difference between CdS QD and SrTiO3 gradually becomes smaller. The driving force for electron transfer at the QD/SrTiO3 interface is dependent on the band edge difference. A reduced conduction band difference leads to a decreased driving force and charge injection efficiency.21

Actually, as the number of coating cycles increases, on one hand, the light harvesting efficiency of the cell and the quantity of photo-generated electrons gradually increase, leading JSC to increase; on the other hand, the driving force and charge injection efficiency gradually decrease, leading JSC to decrease gradually. These two competing mechanisms induce the appearance of an optimum coating cycle (seven cycles) for the JV response. Note that as the number of coating cycles increases to seven, the light harvesting efficiency of the cell does not increase with the increase of cycle number further, because the size confinement hardly takes efforts any more. In this case, the decreased driving force contributes dominantly to the solar cell performance.

Device performance based on SrTiO3 NPs with different sizes

Because the QDSSCs based on CdS QDs prepared with seven cycles demonstrate the optimum device permanence, in the following experiments the CdS QDs are all fixed at seven cycles. Fig. 5a shows the influence of SrTiO3 particle size on the JV characteristics of seven-cycle cells. It can be seen that when the size decreases from 120 to 30 nm, the short circuit photocurrent density JSC increases almost three times, from 0.56 to 1.53 mA cm−2. Concomitantly, the Voc increases from 0.72 to 0.76 V, and the FF increases from 0.48 to 0.67. The power conversion efficiency increases from 0.19% to 0.78%. The detailed device performance parameters are listed in Table 1. From Fig. 5a, the device with 30 nm SrTiO3 NPs as the photoanode shows the best device performance of 0.78% with the highest Voc and FF reported in typical CdS-QDSSCs.22,23
image file: c4ra11960f-f5.tif
Fig. 5 (a) JV characteristics of QDSSCs based on CdS QDs prepared with seven coating cycles and different sized SrTiO3 NPs. (b) UV-vis absorption spectra of CdS QDs prepared with seven coating cycles on SrTiO3 NPs with different sizes. The inset shows the photographs of CdS QDs prepared with seven coating cycles on 120, 70 and 30 nm SrTiO3 NPs. (c) IPCE spectra of QDSSCs based on SrTiO3 NPs with different sizes and CdS QDs coated using seven SILAR cycles.
Table 1 Parameters obtained from JV measurements of the QDSSCs prepared with seven coating cycles
Architecture Voc (V) JSC (mA cm−2) FF η (%)
SrTiO3 (30 nm) 0.76 1.53 0.67 0.78%
SrTiO3 (70 nm) 0.74 0.94 0.51 0.35%
SrTiO3 (120 nm) 0.72 0.56 0.48 0.19%


In order to reveal the origin of device performance changing with SrTiO3 particle size, the absorption spectra and IPCE measurements based on the three kinds of best performance devices were designed.

The absorption spectra of CdS QDs, which are obtained by subtracting the blank (no QDs) device absorption from a working device absorption, are presented in Fig. 5b. It can be seen that the absorption intensities of CdS QDs increase when the size of SrTiO3 NPs decreases, indicating that smaller SrTiO3 NPs have higher QD coverage. A similar phenomenon was also observed for the QDSSCs based on TiO2.24 The inset of Fig. 5b shows the photographs of different sized SrTiO3 NPs coated with seven-cycle CdS QDs. The coating of CdS QDs over SrTiO3 is accompanied by color changes visible to the naked eye. The color becomes darker as smaller SrTiO3 NPs are utilized for CdS QD coating, which confirms higher QD coverage on smaller SrTiO3 NPs. Higher QD coverage leads to better device performance due to improved light harvesting efficiency of the cell. Fig. 5c shows the IPCE spectra of seven-cycle cells based on 120, 70 and 30 nm SrTiO3 NPs. The maximum IPCE values increase from 5.7% to 15.6% as the size of NPs decreases following the same trend of JV of QDSSCs (Fig. 5a). The IPCE spectra of CdS QDs are consistent with the corresponding absorption spectra, as shown in Fig. 5b. The onset wavelengths are very similar for devices based on the SrTiO3 NPs of different sizes, which can be observed in both IPCE and absorption spectra (Fig. 5b and c), indicating that the sizes of the QDs are the same for seven-cycle QDs. Although the average size of SrTiO3 NPs changes, however, the size of QDs is only related to the number of SILAR coating cycles. This explains why different sized SrTiO3-based QDSSC devices have the same ideal performance appearing in the seventh cycles.

Different intensities in IPCE and absorption spectra can be attributed to different QD adsorptions on SrTiO3 NPs, and it is the dominant factor to influence the overall performance of the devices.

TiO2/SrTiO3 composite-based QDSSC device performance

It was reported that incorporating different materials and using their respective advantages had a great influence on the device performance.25–27 In order to take advantage of the high Voc and FF characteristics of the device based on 30 nm SrTiO3 NPs and the high current characteristics of TiO2 based QDSSCs. QDSSCs based on the TiO2/SrTiO3 composite system at different ratios of TiO2 to SrTiO3 were fabricated and compared. Fig. 6a shows JV curves of QDSSCs based on the TiO2/SrTiO3 composites under AM 1.5 G illumination with a light intensity of 100 mW cm−2. The corresponding solar cell parameters are summarized in Table 2. Compared with pure TiO2, the QDSSC exhibits higher Voc and FF but slightly small JSC when the ratio of TiO2 to SrTiO3 is 9[thin space (1/6-em)]:[thin space (1/6-em)]1 leading to the energy conversion efficiency being enhanced from 1.6% to 1.8% with an improvement of 12.5%. As the ratio of TiO2 to SrTiO3 decreases further, Voc and FF increase but JSC decreases, resulting in decreased conversion efficiency.
image file: c4ra11960f-f6.tif
Fig. 6 (a) JV characteristics of QDSSCs based on CdS QDs prepared with seven coating cycles and TiO2/SrTiO3 composites with different ratios. (b) Nitrogen adsorption–desorption isotherms of 30 nm SrTiO3 NPs and TiO2.
Table 2 Parameters obtained from JV measurements of the QDSSCs based on seven-cycle CdS QDs and TiO2/SrTiO3 composites. The average and standard deviation (s.d.) values are obtained from a batch of 6 identically processed TiO2/SrTiO3 composite-based QDSSCs
Architecture Voc (V) JSC (mA cm−2) FF η (%)
TiO2 (100%) SrTiO3 (0%) 0.58 ± 0.03 6.6 ± 0.30 0.41 ± 0.01 1.6% ± 0.15
TiO2 (90%) SrTiO3 (10%) 0.60 ± 0.02 6.0 ± 0.24 0.48 ± 0.02 1.8% ± 0.18
TiO2 (80%) SrTiO3 (20%) 0.62 ± 0.02 5.0 ± 0.13 0.48 ± 0.01 1.5% ± 0.16
TiO2 (50%) SrTiO3 (50%) 0.70 ± 0.03 3.0 ± 0.15 0.49 ± 0.05 1.0% ± 0.20
TiO2 (20%) SrTiO3 (80%) 0.73 ± 0.05 2.0 ± 0.22 0.56 ± 0.02 0.82% ± 0.15
TiO2 (0%) SrTiO3 (100%) 0.76 ± 0.06 1.53 ± 0.11 0.67 ± 0.03 0.78% ± 0.10


Mao et al. reported the enhanced photovoltaic response of the TiO2/SrTiO3 composite compared to pure SrTiO3 and TiO2.28 Improved separation efficiency of photogenerated carriers due to the matched energy levels of TiO2 and SrTiO3 were considered as one possible reason for the enhanced photovoltaic response. Reduced recombination rate of the composite was also proposed to explain the improved photovoltaic response, which was manifested by photoluminescence spectra.28 In our case, the electrons will be injected into both SrTiO3 and TiO2 as SrTiO3 was incorporated into TiO2. The flat band potential of SrTiO3 is higher than that of TiO2, which leads to the higher Voc of the device based on TiO2/SrTiO3 composite. In addition, according to Mao’s results, the recombination rate in the TiO2/SrTiO3 composite can be reduced compared to pure TiO2, which results in larger FF. The power conversion efficiency is determined by JSC, FF and Voc. The JSC of the QDSSCs based on the TiO2/SrTiO3 composite is smaller than that for pure TiO2. However, FF and Voc can be increased due to the band alignment and reduced recombination rate. Thus it is very possible that the power conversion efficiency of QDSSCs based on TiO2 can be improved as an appropriate amount of SrTiO3 is used in the composite.

Meng et al. investigated CdS/CdSe-sensitized solar cells controlled by the structural properties of compact porous TiO2 photoelectrodes, and achieved a high FF (61–65%). They explained that was due to extending the light path length and decreasing the electron recombination caused by the scattering layer.29 It was also reported that a high FF (56%) could be obtained by using ordered multimodal porous carbon as the counter electrode in QDSSCs, and that was attributed to the low charge transfer resistance and Nernst diffusion impedance of the counter electrodes.30 Emin et al. obtained a high FF of 60% and attributed that to the low series resistances of the FTO/TiO2 films.3 In our study (see Table 3), the series resistance of FTO/SrTiO3 (35.1 Ω) is very similar to that of FTO/TiO2 (35.6 Ω). However, the charge transfer resistance of the device based on SrTiO3 is 924 Ω which is smaller than that of TiO2 (1149 Ω). Thus we think the low charge transfer resistance leads to high FF for the device based on SrTiO3 in comparison with the TiO2-based device. Kang et al. reported a high Voc of 0.77 V by reducing the recombination process via a novel photoanode architecture of “pine tree” ZnO nanorods on a Si nanowires hierarchical branched structure.31 Self-assembled monolayers were used as recombination barriers to realize high efficiency of the devices and the high Voc.32 In our recent work, we used rare earth ions to modulate the band gap of the photoanode in dye sensitized solar cells to increase Voc.33 We think the main reason for the high Voc of the devices based on SrTiO3 is higher flat band potential.

Table 3 Fitting results of the Nyquist plot
Photoanode Rs (Ω) R1 (Ω) CPE1 R2 (Ω) CPE2
SrTiO3 35.1 924 1.073 × 10−5 1528 2.318 × 10−4
TiO2 35.6 1149 1.052 × 10−5 3874 5.956 × 10−4


JSC is mainly determined by light-harvesting efficiency and charge injection efficiency. Light-harvesting efficiency is correlated with the adsorption capacity of the photoanode, and adsorption capacity is determined by higher specific surface area. Nitrogen adsorption measurements are designed to measure the specific surface area of samples, the nitrogen adsorption–desorption isotherms based on SrTiO3 (30 nm) and TiO2 (P25) samples are plotted in Fig. 6b. The BET surface areas for the SrTiO3 and TiO2 are reported to be 40 and 51 m3 g−1 respectively by the nitrogen adsorption–desorption measurements. Thus the adsorption ability of 30 nm SrTiO3 is lower than that of P25 due to the difference in surface areas. The surface areas of 120 nm and 70 nm SrTiO3 NPs are 7.1 and 15.7 m3 g−1 respectively. This further explains the lower JV and IPCE of the devices based on 120 nm and 70 nm SrTiO3 NPs compared to that based on 30 nm SrTiO3 NPs shown in Fig. 5a and c.

To study the adsorption ability of SrTiO3 compared to TiO2, oxide photoanodes coated with CdS QDs by the SILAR process were put into 50 mL dilute nitric acid for two hours to dissolve the CdS QDs and to produce Cd2+. A VA Computrace (TEA 4000) was used to detect the concentration of Cd2+, and the results are listed in Table S1. It can be observed that the concentration of Cd2+ increases when the ratio of TiO2 to SrTiO3 increases. The concentration of Cd2+ taken by SrTiO3 is lower compared to that of TiO2, which indicates that SrTiO3 has lower adsorption ability. Thus the low JSC of the QDSSC based on SrTiO3 could be partially attributed to the low adsorption ability of SrTiO3 compared to TiO2.

EIS study of SrTiO3-QDSSCs and TiO2-QDSSCs

A comparison study of the electron transfer processes in the SrTiO3-QDSSCs and TiO2-QDSSCs was carried out by EIS measurements. Typical Nyquist plots of SrTiO3-QDSSCs and TiO2-QDSSCs are shown in Fig. 7. The inset shows the equivalent circuit which is used to fit the Nyquist plot according to ref. 34. A smaller semicircle in the high frequency region (around 100 Hz) and a larger semicircle in the low frequency region (around 1 Hz) can be clearly observed. Each semicircle indicating a charge transfer process exhibited by a resistance capacitance (RC) parallel circuit. For a more precise fitting, the capacitance element is replaced by a constant phase element (CPE). Rs, R1 and R2 values are listed in Table 3, which are determined by fitting according to the equivalent circuit. Rs is the device ohmic series resistance, which is contributed by the sheet resistance of the substrates, resistivity of the electrolyte, and electrical contacts of the cell. R1 and CPE1 stand for the charge transfer resistance and double layer capacitance at the electrolyte/counter electrode interface.34 R2 and CPE2 represents the recombination charge transfer resistance and chemical capacitance at the photoanode/QDs/electrolyte interface.29 The Rs value for the SrTiO3-QDSSCs is 35.1 Ω, which is similar to that of TiO2-QDSSCs (35.6 Ω), this indicates that SrTiO3-QDSSCs and TiO2-QDSSCs have the same device configuration. The R1 value for the SrTiO3-QDSSCs is 924 Ω, which is a little smaller compared to that of TiO2-QDSSCs (1149 Ω). The R2 values of the SrTiO3-QDSSC and TiO2-QDSSC are 1528 Ω and 3874 Ω respectively. The higher R2 implies a lower recombination of electrons between the electrolyte and the conduction band of the photoanode than is expected.29 This result reveals that the recombination of electrons in SrTiO3-QDSSCs is higher than TiO2-QDSSCs, leading to a lower JSC in SrTiO3-QDSSCs.
image file: c4ra11960f-f7.tif
Fig. 7 Nyquist plots of QDSSCs. The inset is the equivalent circuit. The scattered points are experimental data and the solid lines are the fitting curves.

The driving force analysis between SrTiO3/TiO2 and CdS QD

Fig. 8 shows the band diagrams and the estimated conduction band edges of SrTiO3, TiO2 and CdS QDs (seven cycles). According to previous discussion and described in Fig. 4b, the conduction band edge of the CdS QDs is estimated as ∼0.1 and ∼0.3 eV higher than that of the SrTiO3 and TiO2 respectively, which will make the charge injection from QDs to SrTiO3 or TiO2 possible. The conduction band edge of SrTiO3 is 0.2 eV higher than that of TiO2, leading to the higher position of the Fermi level in SrTiO3, which is related to the larger Voc of the QDSSCs based on SrTiO3 according to the Voc analysis in theory as follows.
image file: c4ra11960f-f8.tif
Fig. 8 Band alignment of SrTiO3, TiO2 and CdS QDs prepared with seven cycles. (Note that band positions are for reference only and not drawn to precise scale.)

According to experimental results and theoretical calculations, although the Voc is determined by many factors inside the cell, the open circuit voltage of QDSSCs can be estimated by:35,36

Voc = EgΔ1Δ2,
where Eg is the bandgap of CdS QDs, Δ1 represents the energy difference between the conduction band of the CdS QDs and the conduction band of the semiconductor oxide (TiO2 or SrTiO3); Δ2 is the energy difference between the oxidized redox potential of the electrolyte and the valence band of the CdS QDs. Compared with the conventional TiO2, SrTiO3 has a reduced Δ1, this will lead to improvement in the Voc of the device. Therefore, the QDSSCs based on SrTiO3 have a higher open circuit voltage than that based on TiO2. However, Δ1 can determine the driving force for the charge injection from QDs to SrTiO3. So it will also result in a small driving force for the charge injection from QDs to SrTiO3. In addition, the charge recombination process happens synchronously when Δ1 decreases. These two factors would partially contribute to the low device performance compared to TiO2-QDSSCs. In a word, the low JSC of SrTiO3-QDSSCs can be attributed to the smaller surface area of the SrTiO3 NPs, the smaller driving force due to the higher flat band potential and the smaller recombination charge transfer resistance of the device. Thus both higher Voc and lower JSC originate from the flat band potential of SrTiO3 which leads to the lower device performance.

Conclusions

In summary, in this work, SrTiO3 NPs with different sizes were synthesized by a hydrothermal method and utilized as the photoanode in QDSSCs. CdS QDs were deposited on the NPs using the SILAR approach. Device performance of the QDSSCs was studied, and the highest Voc of 0.76 V and FF of 67% were observed for the cells prepared with seven coating cycles on 30 nm SrTiO3 NPs. IPCE measurements were carried out to inspect CdS bandgaps and alignment with SrTiO3 NPs. Through nitrogen adsorption measurements, it was proven that 30 nm SrTiO3 NPs have a surface area of 40 m3 g−1, which is smaller than that of P25 (51 m3 g−1). Improved device performance of TiO2-QDSSCs was observed as 30 nm SrTiO3 NPs were incorporated into TiO2, which can be attributed to the high Voc of SrTiO3-QDSSCs caused by their band structure and larger FF due to a reduced recombination rate. That points to a new method to improve the solar cell efficiency. EIS measurements for SrTiO3-QDSSC and TiO2-QDSSC were also carried out to study electron transport and recombination processes in the QDSSCs. Smaller recombination charge transfer resistance was observed for the SrTiO3-QDSSC compared to the TiO2-QDSSC, which can result in lower device performance. The first scientific significance of QDSSC fabrication with the SrTiO3 is SrTiO3 nanoparticles can be used as the photoanode for QDSSCs, and the devices based on SrTiO3 show high Voc and FF compared to other QDSSCs. The second one is that TiO2/SrTiO3 composites were developed to improve the TiO2-QDSSC device performance by using the high Voc and FF of SrTiO3-QDSSCs.

Acknowledgements

This work was supported by the Major State Basic Research Development Program of China (973 Program) (no. 2014CB643506), the National Natural Science Foundation of China (Grant no. 11374127, 11304118, 61204015, 81201738, 81301289, 61177042, and 11174111), Program for Chang Jiang Scholars and Innovative Research Team in University (no. IRT13018). The China Postdoctoral Science Foundation Funded Project (2012M511337 and 2013T60327).

Notes and references

  1. D.-M. Li, L.-Y. Cheng, Y.-D. Zhang, Q.-X. Zhang, X.-M. Huang, Y.-H. Luo and Q.-B. Meng, Sol. Energy Mater. Sol. Cells, 2014, 120, 454–461 CrossRef CAS PubMed .
  2. G. Seo, J. Seo, S. Ryu, W. Yin, T. K. Ahn and S. I. Seok, J. Phys. Chem. Lett., 2014, 5, 2015–2020 CrossRef CAS .
  3. S. Emin, M. Yanagida, W. Peng and L. Han, Sol. Energy Mater. Sol. Cells, 2012, 101, 5–10 CrossRef CAS PubMed .
  4. O. E. Semonin, J. M. Luther, S. Choi, H.-Y. Chen, J. Gao, A. J. Nozik and M. C. Beard, Science, 2011, 334, 1530–1533 CrossRef CAS PubMed .
  5. C.-H. M. Chuang, P. R. Brown, V. Bulović and M. G. Bawendi, Nat. Mater., 2014, 13, 796–801 CrossRef CAS PubMed .
  6. B. Tan, E. Toman, Y. Li and Y. Wu, J. Am. Chem. Soc., 2007, 129, 4162–4163 CrossRef CAS PubMed .
  7. Y. Zhang, H. Zhang, Y. Wang and W. Zhang, J. Phys. Chem. C, 2008, 112, 8553–8557 CAS .
  8. S. Burnside, J.-E. Moser, K. Brooks, M. Grätzel and D. Cahen, J. Phys. Chem. B, 1999, 103, 9328–9332 CrossRef CAS .
  9. H. O. Finklea, Semiconductor electrodes, Elsevier Science Ltd, 1988 Search PubMed .
  10. R. Moos, W. Menesklou and K. H. Härdtl, Appl. Phys. A, 1995, 61, 389–395 CrossRef .
  11. P. Tiwana, P. Docampo, M. B. Johnston, H. J. Snaith and L. M. Herz, ACS Nano, 2011, 5, 5158–5166 CrossRef CAS PubMed .
  12. X. Wei, G. Xu, Z. Ren, C. Xu, G. Shen and G. Han, J. Am. Ceram. Soc., 2008, 91, 3795–3799 CrossRef CAS PubMed .
  13. K. Fujinami, K. Katagiri, J. Kamiya, T. Hamanaka and K. Koumoto, Nanoscale, 2010, 2, 2080–2083 RSC .
  14. H. Lee, M. Wang, P. Chen, D. R. Gamelin, S. M. Zakeeruddin, M. Graetzel and M. K. Nazeeruddin, Nano Lett., 2009, 9, 4221–4227 CrossRef CAS PubMed .
  15. H. Yu, X. Xu, X. Chen, T. Lu, P. Zhang and X. Jing, J. Appl. Polym. Sci., 2007, 103, 125–133 CrossRef CAS .
  16. D. Guin, S. V. Manorama, J. N. L. Latha and S. Singh, J. Phys. Chem. C, 2007, 111, 13393–13397 CAS .
  17. H. Lee, H. C. Leventis, S.-J. Moon, P. Chen, S. Ito, S. A. Haque, T. Torres, F. Nueesch, T. Geiger, S. M. Zakeeruddin, M. Graetzel and M. K. Nazeeruddin, Adv. Funct. Mater., 2009, 19, 2735–2742 CrossRef CAS .
  18. T. Lv, L. Pan, X. Liu, T. Lu, G. Zhu, Z. Sun and C. Q. Sun, Catal. Sci. Technol., 2012, 2, 754–758 CAS .
  19. Q. Dai, J. Chen, L. Lu, J. Tang and W. Wang, Nano Lett., 2012, 12, 4187–4193 CrossRef CAS PubMed .
  20. K. Tvrdy, P. A. Frantsuzov and P. V. Kamat, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 29–34 CrossRef CAS PubMed .
  21. G. Rothenberger, D. K. Negus and R. M. Hochstrasser, J. Chem. Phys., 1983, 79, 5360–5367 CrossRef CAS PubMed .
  22. T. Bora, H. H. Kyaw and J. Dutta, Electrochim. Acta, 2012, 68, 141–145 CrossRef CAS PubMed .
  23. Q. Zhang, Y. Zhang, S. Huang, X. Huang, Y. Luo, Q. Meng and D. Li, Electrochem. Commun., 2010, 12, 327–330 CrossRef CAS PubMed .
  24. A. Kongkanand, K. Tvrdy, K. Takechi, M. Kuno and P. V. Kamat, J. Am. Chem. Soc., 2008, 130, 4007–4015 CrossRef CAS PubMed .
  25. B. C. O’Regan, S. Scully, A. C. Mayer, E. Palomares and J. Durrant, J. Phys. Chem. B, 2005, 109, 4616–4623 CrossRef PubMed .
  26. A. Kay and M. Gratzel, Chem. Mater., 2002, 14, 2930–2935 CrossRef CAS .
  27. K. Tennakone, G. K. R. Senadeera, V. P. S. Perera, I. R. M. Kottegoda and L. A. A. De Silva, Chem. Mater., 1999, 11, 2474–2477 CrossRef CAS .
  28. Y. T. Zheng, Z. L. Zhang and Y. L. Mao, J. Alloys Compd., 2013, 554, 204–207 CrossRef CAS PubMed .
  29. Q. Zhang, X. Guo, X. Huang, S. Huang, D. Li, Y. Luo, Q. Shen, T. Toyoda and Q. Meng, Phys. Chem. Chem. Phys., 2011, 13, 4659–4667 RSC .
  30. S.-Q. Fan, B. Fang, J. H. Kim, B. Jeong, C. Kim, J.-S. Yu and J. Ko, Langmuir, 2010, 26, 13644–13649 CrossRef CAS PubMed .
  31. P. Sudhagar, T. Song, D. H. Lee, I. Mora-Seró, J. Bisquert, M. Laudenslager, W. M. Sigmund, W. I. Park, U. Paik and Y. S. Kang, J. Phys. Chem. Lett., 2011, 2, 1984–1990 CrossRef CAS .
  32. P. Ardalan, T. P. Brennan, J. R. Bakke and S. F. Bent, in Photovoltaic Specialists Conference (PVSC), 2010 35th IEEE, 2010, pp. 000951–000954 Search PubMed .
  33. C. Miao, C. Chen, Q. Dai, L. Xu and H. Song, J. Colloid Interface Sci., 2015, 440, 162–167 CrossRef CAS PubMed .
  34. Z. Huang, G. Natu, Z. Ji, P. Hasin and Y. Wu, J. Phys. Chem. C, 2011, 115, 25109–25114 CAS .
  35. H. J. Snaith, Adv. Funct. Mater., 2010, 20, 13–19 CrossRef CAS .
  36. J. Zhang, H.-B. Li, S.-L. Sun, Y. Geng, Y. Wu and Z.-M. Su, J. Mater. Chem., 2012, 22, 568–576 RSC .

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra11960f

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.