DOI:
10.1039/C5RA21822E
(Paper)
RSC Adv., 2015,
5, 103095-103104
Concordantly fabricated heterojunction ZnO–TiO2 nanocomposite electrodes via a co-precipitation method for efficient stable quasi-solid-state dye-sensitized solar cells†
Received
19th October 2015
, Accepted 12th November 2015
First published on 13th November 2015
Abstract
This manuscript is concerned with the successful attempts we have made to fabricate nanostructured spheres composed of mixed metal oxides. ZnO/TiO2 nanocomposites supported on an FTO substrate are used as the photoanode electrode for quasi-solid-state dye-sensitized solar cells (QS-DSSCs). The phase purity of the ZnO and TiO2 phases of the composite shell has been studied by X-ray diffraction peak analysis. A novel gel polymer electrolyte based on a poly(acrylamide)–poly(ethylene glycol) composite and a binary organic solvent was prepared. The polymer gel electrolyte based on the composite of poly(acrylamide)–poly(ethylene glycol), the binary organic solvent of ethylene carbonate and propylene carbonate and the additive of 4-tert-butylpyridine has been employed to fabricate a quasi-solid-state dye-sensitized solar cell. The conversion efficiency of the dye-sensitized solar cells with nanocomposites is 6.5% which is more than double compared with that of bare ZnO nanoparticle photoanodes (3.8%). We believe that this improvement comes from the synergetic effect between ZnO and TiO2, which increases dye absorption, electron transport and electron lifetime, as discussed with the EIS and loaded absorption results. From the current–voltage and incident photon-to-current conversion efficiency (IPCE) measurement, the conductivity of the nanocomposite ZnO–TiO2 was shown to be higher compared to the nanostructured ZnO itself. This simple method can be universally adopted for all quasi-solid-state electrolyte-based DSSCs in order to improve their performance and durability.
Introduction
After the discovery of the photoelectrochemical properties of nanostructured titanium oxide (TiO2), it has been recognized as one of the most promising wide bandgap semiconducting materials for photocatalysis, dye-sensitized/quasi-solid-state dye-sensitized solar cells (DSSCs/QS-DSSCs) and lithium ion batteries.1–6 The DSSC is a molecular approach to photovoltaic solar energy conversion technology. This is one of the emerging photovoltaic technologies that offers the potential to reduce the cost of photovoltaic electricity production. During the past two decades, nanoporous polycrystalline titania has been extensively used in DSSCs, which have demonstrated to be a promising alternative to silicon-based solar cells due to their relatively high solar-to-electric power conversion efficiency at a low cost.
ZnO has similar energy band levels, superior electron mobility,3 and a functioning high specific surface area, which results in it being an effective dye-adsorption and light-scattering layer. To achieve this we have developed a co-precipitation wet-chemical deposition synthesis route for ZnO–TiO2 nanocomposite structures with a well-defined shape and size. Such novel nanocomposite powders provide not only an effective surface area but are also helpful for effective light harvesting in DSSCs.27–29
The light-to-electric energy conversion efficiencies of DSSCs based on liquid electrolytes using organic compounds, such as acetonitrile, propylene carbonate and ethylene carbonate as solvent and the iodide/triiodide (I−/I3−) redox couple as electrolyte, have reached 10–11% under irradiation of AM 1.5.2–7 However, this type of liquid-junction cell still has some problems including low long-term stability, which is caused by organic solvent evaporation and leakage of liquid electrolytes, high temperature instability, and difficulties in sealing the devices.8 To overcome these problems, much effort has been made to replace the liquid electrolytes with solid or quasi-solid type charge transport materials.9–12 Compared with other kinds of charge transport materials, the gel polymer electrolytes have some advantages including high ionic conductivities, which are achieved by trapping the liquid electrolyte in polymer cages formed in a host matrix, and good contacting and filling properties with the nanostructured electrode and counter electrode. Up to now, several types of gel electrolytes based on different kinds of polymers have already been used in quasi-solid-state dye-sensitized solar cells.12–25 Poly(acrylamide) possess a carbonyl group and an amine group on its molecular chain,26 and also poly(ethylene glycol) has many hydroxyl groups on its molecular chain. It is expected that the interaction between the sensitized dye and the matrix of the gel polymer can be improved based on the hydrogen bond interaction between the carbonyl group on sensitized dye and the carbonyl group, amine group and hydroxyl group on the matrix of the gel polymer electrolyte. On the other hand, the ionic conductivity of the gel polymer electrolyte can be enhanced according to the complexation from the carbonyl group, amine group and hydroxyl group on poly(acrylamide) and poly(ethylene glycol) to K+ ions in the electrolyte. Consequently, the overall conversion efficiency of the DSSC can be enhanced. The present study is focused on using poly(ethylene glycol) as both a reactant and plasticizer, amd a novel homogeneous poly(acrylamide)–poly(ethylene glycol) composite without phase separation was synthesized. The key innovation in the present study is to use the composite as a matrix, with 4-tert-butylpyridine as an additive, and the binary organic compounds ethylene carbonate and propylene carbonate as solvent, and a gel polymer electrolyte with a quasi-solid-state was prepared. Furthermore, a dye-sensitized solar cell was fabricated by sandwiching the gel polymer electrolyte. Meanwhile, the structure and band analysis were examined by X-ray diffraction patterns (XRD) and X-ray photoelectron spectroscopy (XPS) analysis. The structure and band analysis was examined using XRD and XPS. The morphological study of the obtained powders was carried out using FESEM and TEM. Optical characterization was performed using UV-vis-NIR spectroscopy. The electrochemical properties, such as the charge transport in the ZnO–TiO2-based photoanodes, and the ZnO–TiO2 /electrolyte interfacial properties, were investigated in detail using current–voltage characteristics and IPCE measurements.
Experimental section
Poly(ethylene glycol) with an average molecular weight of 400 (PEG-400), acrylamide monomer, ammonium persulfate, potassium iodide, iodine, ethylene carbonate (EC), propylene carbonate (PC) and γ-butyrolactone (γ-BL) were all of A.R. grade and all purchased from Sigma-Aldrich Chemicals. PEG-400 was dried at 120 °C for 12 h before use. Other reagents were used without further treatment before use.
Conducting glass plates (FTO glass, with a fluorine-doped tin oxide over-layer, and a sheet resistance of 15 Ω sq−1, purchased from Hartford Glass Co., USA) were used as a substrate for precipitating the TiO2 porous film on. The sensitizing dye cis-[(dcbH2)2Ru(SCN)2] was purchased from Solaronix, SA.
Preparation of pure ZnO nanoparticles and ZnO–TiO2 nanocomposites
In a typical procedure for co-precipitation,30 zinc nitrate (ZnNO3·5H2O) (Sigma-Aldrich) as a source of Zn2+ ions in the case of pure ZnO nanoparticles, or zinc nitrate and titanium trichloride (TiCl3) (Sigma Aldrich) as a source of Zn2+ and Ti2+ ions in the case of ZnO–TiO2 nanocomposites, are dissolved with an aqueous sodium hydroxide solution (5 M) (Fluka) and the pH is adjusted to the desired value of 10. Furthermore, the preparation of the TiO2 nanoparticles is mentioned in the ESI.† The resulting milky white precipitate in each case was collected, filtered, washed with distilled water several times to remove impurities, and then dried at 80 °C for 24 h. The as-prepared ZnO nanopowders and ZnO–TiO2 nanocomposites were annealed at 600 °C for 2 h. After cooling down, the thermally treated powders were collected, packed and kept in a desiccator for further physicochemical investigations.
Synthesis of the poly(acrylamide)–poly(ethylene glycol) in situ composite
The poly(acrylamide)–poly(ethylene glycol) in situ composite was synthesized by adding 3 g PEG-400 into 6 g acrylamide monomer. The mixture was heated at 70–75 °C to melt the monomer and to mix the two components homogeneously. The mixture was marked as (A). A polymerization initiator (ammonium persulfate, 1 wt% of acrylamide monomer) and PEG-400 (1 g) were mixed and stirred until the mixture dissolved entirely at room temperature, and the mixture was marked as (B). Under vigorous stirring and being kept at a temperature of 70–75 °C in a water bath heater, mixture (B) was added into mixture (A) slowly, the polymerization reaction took place, and a homogeneous mixture formed; the mixture was marked as (C). It is noticeable that as soon as the polymerization reaction begins, mixture (C) should be taken out of the water bath heater and cooled down to room temperature to prevent implosion because the reaction is exothermic. When mixture (C) became a viscous gel in the ambient environment, it was heated again at 60–65 °C for 30 min to complete the polymerization reaction. After that, the poly(acrylamide)–poly(ethylene glycol) in situ composite was synthesized.
Preparation of the gel polymer electrolyte
A mixture of organic solvent was made by mixing ethylene carbonate (EC), propylene carbonate (PC), and γ-butyrolactone (γ-BL). Potassium iodide (KI = 10 wt% of the total weight of the composite and the mixture solvent) and iodine (I2 = 10 wt% of KI) were dissolved in the mixture of organic solvent to form a liquid electrolyte. A suitable amount of the poly(acrylamide)–poly(ethylene glycol) composite was added into the liquid electrolyte under continuous stirring at room temperature to form a gel polymer electrolyte with a quasi-solid-state.
Assembly of the quasi-solid-state dye-sensitized solar cell
Photoanodes and solar cells were prepared as reported elsewhere.31–35 Pastes of ZnO, TiO2 or ZnO–TiO2 were coated onto the FTO substrate (15 Ω sq−1) through mixing with distilled water and absolute ethanol. The coated substrates were heated at 450 °C for 30 min. After cooling, the coated substrates (electrodes) were immersed in a dye bath containing 0.3 mM N719 dye (Solaronix) in ethanol solution for 24 h to absorb the dye adequately, the other impurities were washed up with anhydrous ethanol, and they were dried in moisture-free air. After that, a porous film electrode of ZnO, TiO2 or ZnO–TiO2 with absorbed dye was prepared. A quasi-solid-state dye-sensitized solar cell was assembled by dropping a drop of the gel polymer electrolyte into the aperture between the ZnO, TiO2 or ZnO–TiO2 porous film electrode (anode electrode) and a Pt plated conducting glass sheet (cathode electrode, prepared by adding a 50 nm layer of platinum (Pt) on the ITO surface using electrodeposition). The two electrodes were clipped together and a cyanoacrylate adhesive was used as a sealant to prevent the electrolyte solution from leaking.
Surface/structure characterization
The crystallite phases were identified by X-ray diffraction (XRD) on a Bruker axis D8 diffractometer. The particle morphologies were examined using field emission scanning electron microscopy (FESEM, JSM-6700FT, JEOL). Transmission electron microscopy (TEM) and high resolution transmission electron microscopy (HRTEM) were performed with a JEOL-JEM-1230 microscope. The UV-vis transmission spectrum was measured by a UV-vis-NIR scanning spectrophotometer (Jasco-V-570 Spectrophotometer, Japan). The Fourier transform infrared (FTIR) absorption spectrum was obtained using a JASCO 3600 spectrophotometer in the wave number range 200–4000 cm−1. XPS (ESCA-3400, Shimadzu) was used to analyze the chemical composition of the surfaces. Photocurrent–voltage J–V characteristic curve measurements were performed using solar simulation and a Keithley 2601 multimeter. IPCE analyses were carried out using a QE/IPCE measurement system from Oriel at 10 nm intervals between 300 and 700 nm, where a monochromator was used to obtain the monochromatic light from a 300 W Xe lamp. A calibrated photodiode (S1227-1010BQ from Hamamatsu) was used before each IPCE analysis.
Results and discussion
Growth and characterization of the pure ZnO and TiO2 nanoparticles and the ZnO–TiO2 nanocomposites
ZnO nanoparticles were fabricated on FTO. The XRD data (Fig. 1a) show that the nanoparticles grew as crystalline ZnO with the hexagonal wurtzite structure (space group P63mc (186); a = 0.3249 nm, c = 0.5206 nm). The data are in agreement with the Joint Committee on Powder Diffraction Standards (JCPDS) card for ZnO (JCPDS 070-8070). The prominence of the peak assigned to the (002) plane of ZnO is consistent with predominant ZnO growth along the c-axis, perpendicular to the substrate, giving the nanorod morphology. Furthermore, TiO2 nanoparticles were synthesized by an autoclave hydrolysis method and ultimately dispersed in water. On the basis of a Scherrer analysis of the (010) peak at 26° of the X-ray powder diffraction pattern (as shown in Fig. 1b), the crystallite size of the TiO2 nanoparticles was found to be 16 nm, and only the anatase phase was observed. The particle size is generally similar to the crystallite size for this synthesis method.36 Full profile pattern fittings were applied to confirm the crystal phases of the samples as well as to determine the crystallite sizes. No impurity peaks are detected, indicating that the grown nanostructures were of high purity and good crystallinity.
 |
| Fig. 1 Left: XRD patterns. Right: 3D simulation of the pure ZnO, the pure TiO2 and the ZnO/TiO2 nanocomposite. | |
More complicated features are observed in the spectrum of the mixed oxide system (Fig. 1). The reflections, due to the hexagonal ZnO crystal phase, vanished. The spectrum form suggests a high degree of amorphous fraction. The broad band at 29.6° and the clearly seen maxima at 30.9°, 44°, 53.2°, 69.5°, 73° and 77° are assigned to ZnTiO3 (cubic symmetry; JCPDS 00 025 1164). The other two peaks at 34.3° and 36.1° can be referred to as anatase ZnO. The peak located at 64.2°, which is most probably related to cubic TiO2 for the XRD line at the same position, appeared in the spectrum of the ZnO film (Fig. 1).
Morphological properties
FTIR studies of the ZnO/TiO2 film show the characteristics of the formation of a high purity material. The FTIR spectrum in Fig. S2 (see ESI†) clearly shows the peaks corresponding to ZnO and TiO2. Peaks located in the area of 1000–1250 cm−1 and 1750 cm−1 correspond to the vibration of Ti–O and the Ti–O–O bridging stretching mode.37 A sharp strong band around 669 cm−1 was observed corresponding to the stretching mode of the Zn–O nanoparticles. On the other hand, the absence of peaks at 1400 cm−1 proves the absence of peroxo groups. It is noted that the examination of the film in IR was made on silicon wafer. It is also clear from the data in Fig. S3 that there are no IR absorption peaks corresponding to impurities like organic residues, –CH and –CH2 at 1400–2900 cm−1 and C–O–C at 1000–1500 cm−1. Furthermore, the FTIR spectrum firmly suggests the presence of –OH groups, which is absolutely necessary for dye adsorption in DSSC fabrication. The samples show peaks corresponding to the stretching vibration of O–H and bending vibrations of adsorbed water molecules at around 3300–3450 cm−1 and at 1640 cm−1 respectively.38 Thus, the as-prepared ZnO/TiO2 films after washing with pure water consisted of pure ZnO and TiO2 particles without any significant organic contaminants, which is an important advantage in the evaluation of the DSSC efficiency made with these ZnO/TiO2 electrodes.
The FESEM images in Fig. 2a and b show that the ZnO and ZnO/TiO2 films exhibit a uniform surface morphology without any cracks. The figures show an approximately spherical morphology with the agglomeration of particles, which is the core material. In the case of adding ZnO to the TiO2 core material, the size of the particles is slightly increased which is related to the shell thickness as presented in Fig. 2b. The chemical compositional analysis of the samples is essential for knowing the exact concentration of elements, added dopants, and defects, if any, present in the samples. Fig. 2c shows a cross sectional view of a QS-DSSC consisting of an FTO layer covered with a porous, 1.2 μm thick 20 nm crystallite TiO2 film. N719 Ru dye and polymer gel electrolyte fill the pores of the film and form a 110 nm thick over-layer. Moreover, Fig. 2d shows a schematic diagram of the QS-DSSC shown in Fig. 2c. Furthermore, it is interesting to note (according to Fig. S1, left, ESI†) that the TiO2 material is highly crystalline, as expected, and the particles are connected between them. There is no indication of isolated particles, thus we believe that the formation of the TiO2 film is accompanied by good mechanical stability. Apparently, there is no measurable fraction of the amorphous TiO2 phase because the quantity of titanium(IV)-iso-propoxide (TTIP) used was relatively low. The nanoparticles had an average size of 25–30 nm, confirming the X-ray pattern behavior.
 |
| Fig. 2 FESEM patterns of the pure ZnO (a), the ZnO/TiO2 nanocomposite (b), and the cross section of a full cell (c), and a schematic diagram of a dye-sensitized solar cell (d). | |
The shape, crystallinity and size of the particles were examined with HRTEM. Fig. 3a–d show the TEM images of the ZnO and ZnO/TiO2 films. Both of them show a spherical morphology with good monodispersity of the nanoparticles. Fig. 3b of ZnO/TiO2 infers that ZnO was effectively reacted as a composite with the core material TiO2. The high resolution image of each sample is shown in Fig. 3c and d. The lattice-resolved image in Fig. 3c exhibits a typical spacing of about 2.40 Å, corresponding to the (101) plane of the wurtzite ZnO structure. For ZnO–TiO2 composite systems, along with TiO2, a lattice spacing of about 2.35 Å corresponding to the (101) plane of wurtzite ZnO and a lattice spacing of 3.45 Å corresponding to the (101) plane of TiO2 anatase was observed, as shown in Fig. 3d. Here, it is possible to observe the interference between the lattice corresponding to ZnO and TiO2, which may be due to the good formation of the nanocomposites between both of them.
 |
| Fig. 3 TEM patterns of the pure ZnO (a) and the ZnO/TiO2 nanocomposite (b), and HRTEM patterns of the pure ZnO (c) and the ZnO/TiO2 nanocomposite (d). | |
UV-vis spectroscopy, bandgap energy and chemical components
Fig. 4 (left) shows transmittance UV-vis spectroscopy of the ZnO nanoparticles and the ZnO/TiO2 nanocomposites in the wavelength range of 400–800 nm. There is an obvious red shift compared with the ZnO nanoparticles, possibly due to differences in the surface state of the ZnO/TiO2 composites. In addition, the absorbance of the spectrum increases in the case of ZnO compared with that of TiO2 (as shown in Fig. S1, right, in the ESI†), due to the surface states of ZnO. To obtain the bandgap, the spectra shown in Fig. 4 (left) are replotted in Fig. 4 (right). The theory of optical absorption defines the relationship between the absorption coefficient and the photo-energy. The relevant equation for a direct transition semiconductor is shown as follows:39 α = (hν − Eg)1/2/hν, where α = A/t, and α is a coefficient, A is absorbance, t is the film thickness and hν is the photo-energy. Using this equation, curves of hν versus (αhν)2 were plotted and the extrapolated liner portion corresponded to the Eg value. The estimated bandgap energy was 2.58 eV for the ZnO nanoparticles and 3.00 eV for ZnO/TiO2. To the authors’ best knowledge, the band energy of ZnO is 3.37 eV and that for anatase TiO2 is 3.20 eV. The band energy of the ZnO/TiO2 composites is lower than that for the ZnO and TiO2 nanoparticles. That is because the presence of ZnO can modify the optical properties, to extend the range of the excited spectrum and favor the absorption of solar energy in the visible region. It also has been found that a red shift occurs as the ZnO deposition time increased, causing an increase in the amount of Zn. The chemical components and chemical changes due to the modification of the amount of ZnO on the surface of the TiO2 nanoparticles were examined by XPS, as shown in Fig. 5. The ZnO/TiO2 nanocomposites were selected to analyze changes of the surface state. Fig. 5(a) shows the spectra of the TiO2 nanoparticles and the ZnO/TiO2 composites over a wide scan range. Zn was observed for the composites, because of the ZnO coating on the TiO2 nanoparticles. A high resolution XPS spectrum of Zn on the surface of the ZnO/TiO2 composites is shown in Fig. 5(c). The Zn 2p3/2 peak at approximately 1021.6 eV is assigned to the Zn–O bonds,40 which indicates the formation of a ZnO composed with TiO2 nanoparticles by a hydrothermal method. The narrow scan of the Ti 2p peaks for the TiO2 nanoparticles is shown in Fig. 5(b). A slight shoulder peak is observed at approximately 459.8 eV. This is due to the formation of Ti3+. The strong chemical activity of Ti3+ can transfer excess electrons to the carboxylate group of the dye molecules, which form coordination bonds and may increase the amount of dye adsorption.41 Moreover, the formation of Ti3+ can increase oxygen vacancies, which support the transport of electrons and holes between the dye molecules and the photoanode. The Ti(2p) peak of the ZnO/TiO2 composites shifts to a lower binding energy. It also shows that the composed ZnO can change the energy band of TiO2. In Fig. 5(d), it can be seen that a good symmetric O 1s peak was obtained for the pure TiO2 nanoparticles, while the O 1s peak for the ZnO/TiO2 composites is asymmetric.
 |
| Fig. 4 UV-vis T% (left) and bandgap (right) of the pure ZnO (a) and the ZnO/TiO2 nanocomposite (b). | |
 |
| Fig. 5 XPS of (a) a wide scan range, (b) and (c) a narrow scan range of the TiO2 and ZnO nanoparticles, respectively, and (d) the binding energy of oxygen particles. | |
Photoelectrochemical properties: J–V characteristics
For the quasi-solid-state DSSCs with the low fluidity electrolyte, the electron collection yield is low because of the short electron diffusion length in the ZnO/TiO2 film.42,43 Reducing the thickness of the ZnO/TiO2 film permits more electrons to be collected by the FTO before recombination by triiodide ions in the electrolyte.44 However, a thinner ZnO/TiO2 film would incur a lower amount of adsorbed dye and limit the absorption of incoming light. In this study, we employ Ru N719 organic dyes with a (PA/PEG) hybrid polymer gel electrolyte to construct quasi-solid-state DSSCs.
The photocurrent density–voltage (I–V) curves of the PA/PEG-based quasi-solid-state solar cells sensitized by Ru N719 using ZnO and/or the ZnO–TiO2 nanocomposites, illuminated at a light intensity of 100 mW cm−2, are displayed in Fig. 6 (left).
 |
| Fig. 6 Comparison of the I–V characteristics (left) and IPCE (right) of the DSSCs made from pure ZnO (a), and the ZnO/TiO2 nanocomposites (b). | |
For comparison, I–V curves of the solar cells sensitized by Ru N719 using TiO2 nanoparticles measured under the same conditions are also presented in Fig S3(a) (discussed in the ESI†). The detailed photovoltaic performance parameters of Voc, Jsc, FF, and η under a certain light intensity are summarized in Table 1. For the ZnO–TiO2 nanoparticle-based cells, the presence of PA/PEG can obviously improve the photovoltaic performance in terms of an increased Voc, Jsc, and η, as compared to that of the pure ZnO or pure TiO2-based cells. As mentioned above, PA/PEG can be adsorbed onto the ZnO, TiO2 and ZnO–TiO2 surface through Ti–O and Zn–O bonds and can suppress back electron transfer from the nanoparticles to I3−,45 consequently raising the quasi-Fermi level of the conduction band electrons in the nanoparticle film and increasing the Voc. As a result, the η value is improved from 3.8% (cell with ZnO) to 6.5% (cell with ZnO–TiO2) and from 5.6% (cell with TiO2 in the ESI†) to 6.5% (cell with ZnO–TiO2). For the cell with ZnO–TiO2 nanoparticles, a more negative lowest unoccupied molecular orbital level of the dye and a lower back reaction rate at the ZnO–TiO2/electrolyte interface between the injected electrons and I3− together result in a higher Jsc as well as Voc (discussed in the ESI†). The considerable efficiency and excellent long-term stability of the cell with the ZnO–TiO2 nanoparticles reveal that the development of this type of ruthenium dye-sensitized quasi-solid-state solar cell is promising for commercial applications.
Table 1 I–V characteristics results of the DSSCs manufactured using the pure ZnO, the pure TiO2 and the ZnO/TiO2 nanocomposite, respectively. Furthermore, a detailed simulated value of recombination resistance (R2) and electron lifetime value (τr) from the EIS spectra, calculated by an equivalent circuit as shown in Fig. 7
Sample |
Voc |
Jsc (mA cm−2) |
FF |
η (%) |
Active area (cm2) |
R2 (Ω) |
τr (s) |
Pure ZnO |
0.812 |
7.2419 |
65.86 |
3.80 |
0.25 |
256.5 |
0.170 |
Pure TiO2 |
0.711 |
9.267 |
67.07 |
5.60 |
0.25 |
158.2 |
0.138 |
ZnO/TiO2 nanocomposite |
0.740 |
13.017 |
67.48 |
6.50 |
0.25 |
96.04 |
0.094 |
The incident photoelectric conversion efficiency and electrochemical impedance spectroscopy (EIS)
The I–V characteristic results are further confirmed by the IPCE values obtained for the DSSCs. The corresponding IPCE values for the DSSCs employing different nanomaterials are mentioned below. The IPCE values obtained for the cells with ZnO–TiO2 and with ZnO are 80% and 54% respectively at 532 nm, as shown in Fig. 6 (right), which are compared to cells with TiO2 that were of 64%, as shown in Fig S3(b).† The highest photocurrent response is attained for the ZnO–TiO2-based DSSCs. The corresponding IPCE values are in good agreement thus confirming the improved pore filling obtained as the liquid electrolyte is utilized together with the gel polymer electrolyte.
Charge transfer dynamics. To obtain better insight into the dynamics of the interfacial charge transfer process within the DSSCs, electrochemical impedance spectroscopy (EIS) was performed. Fig. 7 shows the Nyquist plots of the DSSCs based on the above three photoelectrodes. Generally, two semicircles in the Nyquist plots could be observed. The small semicircles in the high frequency region and large semicircles in the low frequency region46 correspond to the electron recombination at the photoelectrode/dye/electrolyte interface, and this shows that the recombination resistance decreases in the order of pure ZnO > pure TiO2 > ZnO/TiO2 photoelectrode prepared by the co-precipitation route implying a fast recombination reaction (or shorter electron lifetime) for the above photoelectrode order.47–52 The high frequency semicircle (1k–100k Hz) is related to the capacitance (CPE1) and charge transfer resistance (R1) between the counter electrode and electrolyte, while the impedance in the middle frequency region is associated with the capacitance (CPE2) and interfacial charge transfer resistance (R2) between ZnO/TiO2 (or ZnO or TiO2)/dye and the electrolyte. Fig. 7 provides the impedance spectra of the DSSCs with ZnO/TiO2 (or ZnO or TiO2) photoelectrodes. The corresponding R2 values, by fitting the impedance spectra based on the equivalent circuit, are shown in the inset of Fig. 7 and are listed in Table 1. The interfacial charge transfer (recombination) resistance (R2) of the DSSCs decreased significantly from 256.5 Ω to 96.04 Ω in the order of pure ZnO > pure TiO2 > ZnO/TiO2 photoelectrode-based DSSCs, indicating that a faster recombination rate occurred within the DSSCs based on the TiO2 NPs. This may result in a smaller photovoltage which is in agreement with the above photovoltaic data. Specifically, there was a huge difference shown in the large semicircles. This result indicates that the ZnO/TiO2 nanocomposite improves the efficient charge transport in the mesoporous layer.
 |
| Fig. 7 Nyquist diagrams of the impedance spectra of (a) the pure ZnO, (b) the pure TiO2, and (c) the ZnO/TiO2 nanocomposite. Furthermore, the inset refers to the equivalent circuits of the DSSCs. R1: the serial resistance, R2: the charge transfer resistance of FTO ZnO/TiO2 (or ZnO or TiO2) and the counter electrode/electrolyte interfaces, CPE1: the constant phase element of FTO ZnO/TiO2 (or ZnO or TiO2) and the counter electrode/electrolyte interfaces, R3: the electron transfer in the ZnO/TiO2 (or ZnO or TiO2)/dye/electrolyte interfaces, and CPE2: the constant phase element of the ZnO/TiO2 (or ZnO or TiO2)/dye/electrolyte interfaces. | |
Moreover, this result is in good agreement with the changing trends of the PCE and the amount of dye loading as shown in Fig. 8, which is demonstrated by the optical absorbance with the wavelength by dissolving out the adsorbed dye molecules from the photoanode in 10 mL NaOH. The probable reason is that a low dye loading on the ZnO anode may generate a low electron density and interfacial charge transfer rate, causing a high R2 value. With an appropriate coating of TiO2 on ZnO–TiO2, the amount of adsorbed dye increases and significantly reduces the R2 value. This fact may be partly attributed to the penetration limitation of dye molecules due to the blocking effect induced by the coated TiO2 and/or ZnO–TiO2.
 |
| Fig. 8 Optical absorbance of the dissolved-out dyes of the pure TiO2, the pure ZnO and the ZnO/TiO2 nanocomposite, respectively loaded on Ru N719 dye. | |
Since the Voc value of the DSSCs was correlated with the charge recombination in the conduction band of TiO2 (ref. 53) the electron lifetime (τr) values calculated by fitting the equation τr = R2 × CPE2 (CPE2, chemical capacitance) was used to differentiate the difference of Voc in the DSSCs based on various photoanodes. Obviously, τr of the DSSCs decreased with the order of pure ZnO > pure TiO2 > ZnO/TiO2 photoelectrodes, which was in agreement with the significant reduction of Voc displayed in the J–V results. Accordingly, one can conclude that the alternate co-precipitated nanocomposites introduced in the present work markedly augmented the specific surface area of ZnO–TiO2, while on the other hand simultaneously provided additional recombination sites at the surface of the TiO2 NPs for recombination of the photo-generated electrons with I3− in the redox polymer electrolyte. It is worth noting that the electron lifetime for the ZnO–TiO2 cell was longer than that of the TiO2 and ZnO-based DSSCs since the former (nanocomposite) effectively suppresses the charge recombination. ZnO-based DSSCs exhibited the shortest electron lifetime, suggesting that photo-generated electrons within such cells may suffer a series of trapping events or serious recombination with oxidizing species in the electrolyte as they undertook a random network pathway through the mesoporous nanoparticle film.
Conclusions
Quasi-solid-state DSSCs were successfully fabricated using the composite poly(acrylamide)–poly(ethylene glycol)-based polymer gel electrolyte and Ru N719-based organic dyes. The hybrid polymer gel is beneficial to the entrapment of a large volume of liquid electrolyte. The electrolyte gelatinized by the hybrid PA/PEG exhibits considerable ionic conductivity and a triiodide ionic diffusion constant. By optimization of the electrolyte composition and the ZnO, TiO2 and ZnO/TiO2 film thickness, the fabricated quasi-solid-state solar cells sensitized by N719 dye achieved an overall energy conversion efficiency of 3.8, 5.6 and 6.5%, respectively at a light intensity of 100 mW cm−2. Moreover, we have successfully synthesized a composite nanostructure consisting of ZnO nanoparticles composed with TiO2 nanoparticles, using facile wet-chemical methods. The SEM images indicated that the composites reveal a well-ordered and good size distribution of the particles and they were used for photocatalytic dye decolorization. In addition, the nanocomposites showed no aggregation in comparison with other samples. It is important for solar cell activity that the particle size of the photocatalyst should be homogeneous. The XRD study exhibited that there were no impurities in the samples. Further addition of ZnO resulted in an increase in the particle size. This nanocomposite photoanode points to a class of structures that can provide both fast transport and a high surface area, enabling DSSCs that are tolerant of electrolytes with lower recombination rates than those found with the conventional iodide/triiodide couple. The high efficiency and excellent long-term stability of the quasi-solid-state DSSCs assembled with organic dye and the PA/PEG hybrid gel electrolyte indicate that they are promising for industrialization.
Conflicts of interest
The authors declare no competing financial interest.
Acknowledgements
The authors would like to extend their sincere appreciation to Central Metallurgical Research and Development Institute, Egypt for its financial support to pursue the work.
References
- B. O’Regan and M. Grätzel, Nature, 1991, 353, 737–740 CrossRef.
- A. Yella, H.-W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, M. K. Nazeeruddin, E. Wei-Guang Diau, C.-Y. Yeh, S. M. Zakeeruddin and M. Grätzel, Science, 2011, 334, 629–634 CrossRef CAS PubMed.
- E. M. Elsayed, A. E. Shalan and M. M. Rashad, J. Mater. Sci.: Mater. Electron., 2014, 25, 3412–3419 CrossRef CAS.
- M. M. Rashad, A. E. Shalan, Mónica Lira-Cantú and M. S. A. Abdel-Mottaleb, Appl. Nanosci., 2013, 3, 167 CrossRef CAS.
- A. E. Shalan and M. M. Rashad, Appl. Surf. Sci., 2013, 283, 975 CrossRef CAS.
- M. M. Rashad, I. A. Ibrahim, I. Osama and A. E. Shalan, Bull. Mater. Sci., 2014, 37(4), 903–909 CrossRef CAS.
- Q. Zhang, T. P. Chou, B. Russo, S. A. Jenekhe and G. Cao, Angew. Chem., Int. Ed., 2008, 47, 2402–2406 CrossRef CAS PubMed; Q. Zhang, C. S. Dandeneau, S. Candelaria, D. Liu, B. B. Garcia, X. Zhou, Y.-H. Jeong and G. Cao, Chem. Mater., 2010, 22, 2427–2433 CrossRef.
- E. Hosono, S. Fujihara, T. Kimura and H. J. Imai, J. Colloid Interface Sci., 2004, 272, 391–398 CrossRef CAS PubMed.
- Z. Huo, S. Dai, K. Wang, F. Kong, C. Zhang, X. Pan and X. Fang, Sol. Energy Mater. Sol. Cells, 2007, 91, 1959–1965 CrossRef CAS.
- Z. Lan, J. Wu, D. Wang, S. Hao, J. Lin and Y. Huang, Sol. Energy, 2007, 81, 117–122 CrossRef CAS.
- P. Balraju, M. Kumar, Y. S. Deol, M. S. Roy and G. D. Sharma, Synth. Met., 2010, 160, 127–133 CrossRef CAS.
- J. Chen, T. Peng, K. Fan, R. Li and J. Xia, Electrochim. Acta, 2013, 94, 1–6 CrossRef CAS.
- W. Kubo, K. Murakoshi, T. Kitamura, S. Yoshida, M. Haruki, K. Hanabusa, H. Shirai, Y. Wada and S. Yanagida, J. Phys. Chem. B, 2001, 105, 12809–12815 CrossRef CAS.
- S.-Y. Shen, R.-X. Dong, P.-T. Shih, V. Ramamurthy, J.-J. Lin and K.-C. Ho, ACS Appl. Mater. Interfaces, 2014, 6, 18489–18496 CAS.
- S.-H. Park, I. Young Song, J. Lim, Y. Soo Kwon, J. Choi, S. Song, J.-R. Lee and T. Park, Energy Environ. Sci., 2013, 6, 1559 CAS.
- W. Kubo, T. Kitamura, K. Hanabusa, Y. Wada and S. Yanagida, Chem. Commun., 2002, 374–375 RSC.
- Y. Dkhissi, F. Huang, Y.-B. Cheng and R. A. Caruso, J. Phys. Chem. C, 2014, 118, 16366–16374 CAS.
- C. Wang, L. Wang, Y. Shi, H. Zhang and T. Ma, Electrochimica Acta, 2013, 91, 302–306 CrossRef CAS.
- J. Shi, S. Peng, J. Pei, Y. Liang, F. Cheng and J. Chen, ACS Appl. Mater. Interfaces, 2009, 1(4), 944–950 CAS.
- N. Manfredi, A. Bianchi, V. Causin, R. Ruffo, R. Simonutti and A. Abbotto, J. Polym. Sci., Part A: Polym. Chem., 2014, 52, 719–727 CrossRef CAS.
- P. Wang, S. M. Zakeeruddin, J. E. Moser, M. K. Nazeeruddin, T. Sekiguchi and M. Grätzel, Nat. Mater., 2003, 2, 402–407 CrossRef CAS PubMed.
- T. M. W. J. Bandara, W. J. M. J. S. R. Jayasundara, H. D. N. S. Fernado, M. A. K. L. Dissanayake, L. A. A. de Silva, I. Albinsson, M. Furlani and B.-E. Mellander, J. Appl. Electrochem., 2015, 45, 289–298 CrossRef CAS.
- Y. Yang, C.-H. Zhou, S. Xu, J. Zhang, S.-J. Wu, H. Hu, B.-L. Chen, Q.-D. Tai, Z.-H. Sun, W. Liu and X.-Z. Zhao, Nanotechnology, 2009, 20, 105204 CrossRef PubMed (9pp).
- P. M. Sommeling, M. Spath, H. J. P. Smit, N. J. Bakker and J. M. Kroon, J. Photochem. Photobiol. A: Chem., 2004, 164, 137–144 CrossRef CAS.
- A. F. Nogueira, C. Longo and M. A. D. Paoli, Coord. Chem. Rev., 2004, 248, 1455–1468 CrossRef CAS.
- J. Wu, Z. Lan, D. Wang, S. Hao, J. Lin, Y. Wei, S. Yin and T. Sato, J. Photochem. Photobiol. A, 2006, 181, 333–337 CrossRef CAS.
- M. M. Rashad, A. E. Shalan, M. Lira-Cantu and M. S. A. Abdel-Mottaleb, J. Ind. Eng. Chem., 2013, 19, 2052 CrossRef CAS.
- A. E. Shalan, M. Rasly, I. Osama, M. M. Rashad and I. A. Ibrahim, Ceram. Int., 2014, 40, 11619–11626 CrossRef CAS.
- A. E. Shalan, M. M. Rashad, Y. Yu, M. Lira-Cantu and M. S. A. Abdel-Mottaleb, Electrochim. Acta, 2013, 89, 469 CrossRef CAS.
- A. E. Shalan, I. Osama, M. M. Rashad and I. A. Ibrahim, J. Mater. Sci: Mater. Electron., 2014, 25, 303–310 CrossRef CAS.
- A. E. Shalan, M. M. Rashad, Y. Yu, M. Lira-Cantú and M. S. A. Abdel-Mottaleb, Appl. Phys. A, 2013, 110, 111 CrossRef CAS.
- M. M. Rashad and A. E. Shalan, Appl. Phys. A, 2014, 116, 781–788 CrossRef CAS.
- M. M. Rashad, E. M. Elsayed, M. S. Al-Kotb and A. E. Shalan, J. Alloys Compd., 2013, 581, 71 CrossRef CAS.
- I. Tacchini, A. Ansón-Casaos, Y. Yu, M. T. Martínez and M. Lira-Cantu, Mater. Sci. Eng. B, 2012, 177, 19–26 CrossRef CAS.
- I. Gonzalez-Valls, Y. Yu, B. Ballesteros, J. Oro and M. Lira-Cantu, J. Power Sources, 2011, 196, 6609–6621 CrossRef CAS.
- V. Manthina, J. P. C. Baena, G. Liu and A. G. Agrios, J. Phys. Chem. C, 2012, 116, 23864–23870 CAS.
- S. S. Kanmani and K. Ramachandran, Renewable Energy, 2012, 43, 149–156 CrossRef CAS.
- E. Stathatos, Y. Chen and D. D. Dionysiou, Sol. Energy Mater. Sol. Cells, 2008, 92, 1358–1365 CrossRef CAS.
- R. S. Mane, W. J. Lee, H. M. Pathan and S. H. Han, J. Phys. Chem. B, 2005, 109, 24254–24259 CrossRef CAS PubMed.
- S. S. Kim, J. H. Yum and Y. E. Sung, J. Photochem. Photobiol., A, 2005, 171, 269–273 CrossRef CAS.
- R. Liu, W.-D. Yang, L.-S. Qiang and H.-Y. Liu, J. Power Sources, 2012, 220, 153–159 CrossRef CAS.
- F. Gao, Y. Wang, D. Shi, J. Zhang, M. Wang, X. Jing, R. Humphry-Baker, P. Wang, S. M. Zakeeruddin and M. Grätzel, J. Am. Chem. Soc., 2008, 130, 10720–10728 CrossRef CAS PubMed.
- D. Kuang, S. Ito, B. Wenger, C. Klein, J.-E. Moser, R. Humphry-Baker, S. M. Zakeeruddin and M. Gratzel, J. Am. Chem. Soc., 2006, 128, 4146–4154 CrossRef CAS PubMed.
- S. Ito, S. M. Zakeeruddin, R. Humphry-Baker, P. Liska, R. Charvet, P. Comte, M. K. Nazeeruddin, P. Péchy, M. Takata, H. Miura, S. Uchida and M. Grätzel, Adv. Mater, 2006, 18, 1202–1205 CrossRef CAS.
- R. Komiya, L. Han, R. Yamanaka, A. Islam and T. Mitate, J. Photochem. Photobiol., A, 2004, 164, 123–127 CrossRef CAS.
- F. Fabregat-Santiago, J. Bisquert, E. Palomares, L. Otero, D. B. Kuang, S. M. Zakeeruddin and M. Gratzel, J. Phys. Chem. C, 2007, 111, 6550 CAS.
- X. Y. Wang, Y. Liu, X. Zhou, B. J. Li, H. Wang, W. X. Zhao, H. Huang, C. L. Liang, X. Yu, Z. Liu and H. Shen, J. Mater. Chem., 2012, 22, 17531 RSC.
- M. Law, L. E. Greene, J. C. Johnson, R. Saykally and P. D. Yang, Nat. Mater., 2005, 4, 455 CrossRef CAS PubMed.
- I. Montanari, J. Nelson and J. R. Durrant, J. Phys. Chem. B, 2002, 106, 12203 CrossRef CAS.
- M. Wang, P. Chen, R. Humphry-Baker, S. M. Zakeeruddin and M. Grätzel, ChemPhysChem, 2009, 10, 290 CrossRef CAS PubMed.
- W.-Q. Wu, Y.-F. Xu, C.-Y. Su and D.-B. Kuang, Energy Environ. Sci., 2014, 7, 644 CAS.
- L.-C. Chen, S.-F. Tsai, J.-H. Chen and G.-W. Wang, Int. J. Photoenergy, 2013, 417964, 9, DOI:10.1155/2013/417964.
- D. Kuang, C. Klein, Z. Zhang, S. Ito, J. E. Moser, S. M. Zakeeruddin and M. Gratzel, Small, 2007, 3, 2094 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available: Additional SEM images, XRD patterns, optical characterization, a I–V curve and IPCE measurements of various products. See DOI: 10.1039/c5ra21822e |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.