Babasaheb R. Sankapal*a,
Dipak B. Salunkheb,
Sutripto Majumdera and
Deepak P. Dubalc
aNano Materials and Device Laboratory, Applied Physics Department, Visvesvaraya National Institute of Technology, Nagpur – 440 010, M.S., India. E-mail: brsankapal@gmail.com; brsankapal@phy.vnit.ac.in; Fax: +91 712 2223230; Tel: +91 712 2801170
bThin Film and Nano Science Laboratory, Department of Physics, School of Physical Sciences, North Maharashtra University, Jalgaon – 425 001, M. S., India
cCatalan Institute of Nanoscience and Nanotechnology, CIN2, ICN2 (CSIC-ICN), Campus UAB, E-08193 Bellaterra, Barcelona, Spain
First published on 16th August 2016
Mesoporous titanium dioxide (TiO2) nanostructured thin films, formed by a simple chemical route, namely, the successive ionic layer adsorption reaction (SILAR) technique, were efficiently used to grow a cadmium sulphide (CdS) nanoparticle architecture for device grade applications. CdS nanoparticle have a highly symmetric size and shape that could be varied in a controlled manner (<10 nm size) depending on the number of SILAR immersions without the use of an organic linker molecule. XRD and HR-TEM studies showed the formation of CdS nanoparticles on TiO2 nanoparticles. Such a nanomorphology showed good optical coverage over the visible region of light, and it revealed a red-shift with respect to the number of CdS layers, which was confirmed by both optical and external quantum efficiency studies. The photoelectrochemical performance of the TiO2/CdS photoanodes showed that as the particle size of CdS increases with the number of SILAR immersions, the photocurrent density and photovoltage were found to be directly concurrent with the charge transport mechanism, which increases the cell efficiency.
In photoelectrodes, narrow band gap CdS,6–9 CdSe10–15 and CdTe16,17 are used as light harvesters, mostly coupled with wide band gap semiconducting materials, such as TiO2, ZnO and SnO2. One of the subjects investigated recently is the linker-assisted attachment of these metal chalcogenides with wide band gap materials for solar cell applications, such as in the quantum dot sensitized solar cell (QDSSC).6–21 In the current research, the use of bifunctional organic linker molecules for the attachment of nanoparticles to the electrode material reduces the electron tunnelling injection effect.18 Also, different bifunctional linker molecules directly influence the kinetics of charge separation and transportation,11,19–21 and this has been considered a key reason to limit the conversion efficiencies in QDSSC. Hence, there is a prevailing need to find a suitable sensitizer to harvest more visible-light utilizing a linker-free approach. There are many reports where the conformational growth of the CdS quantum dots has been achieved without any linker over different morphologies of TiO2. Among these, Chen et al. demonstrated the growth of CdS QDs by the CBD method, which involved the anchoring of CdS over TiO2 nanotubes.22,23 Recently, it was found that the attachment of the CdS QDs can be done more efficiently through use of the SILAR method. In this perspective, Pawar et al. formed a uniform growth of CdS QDs over TiO2 nanorods.24 Apart from this, Tau et al. synthesized CdS nanoparticles over vertically aligned TiO2 nanorod arrays.25 However, in all of the above-stated linker-free approaches the basic TiO2 film was formed either by anodic oxidation of the titanium foil or by the hydrothermal method. Therefore, to ease the fabrication, an easy, convenient and efficient method for the growth of the TiO2 film is desired. Consequently, in the present study we fabricated TiO2 through a spin-coating method at ambient temperature.
Apart from the fabrication of TiO2, this article also reports on the room temperature controlled growth of a CdS assembly over mesoporous TiO2 by a successive ionic layer adsorption reaction (SILAR) method utilizing a linker-free approach. The direct chemisorption involved in the reaction kinetics facilitates the chemical reaction, resulting in the growth of particles and, hence, the growth can be controlled at an ionic level. van der Waals and electrostatic forces took part in the heterogeneous nucleation, resulting in the growth of quantum-sized CdS nanoparticles on TiO2 film surface by varying the number of SILAR immersions. This method enables the systematic monitoring of the nucleation and growth of accumulation of CdS quantum dots on the TiO2 electrode by an ion-by-ion reaction mechanism. The systematic growth can be modulated in the evolution of nanoparticles, which is depicted in the schematic in Fig. 1 showing the growth at 8, 13, 15 and 20 SILAR immersions, respectively. Additionally, the inset in Fig. 1 clearly demonstrates the TiO2 as bigger-sized particle, while the yellow dots represent the CdS quantum dots formed. The surface morphological studies confirmed that the growth of CdS nanoparticles films resulted in a cauliflower-like morphology with the increasing number of SILAR immersions, as shown in the schematic in Fig. 1. Diverse characterizations were performed to support the size tuning and the results are reported and discussed herein. The synthesized nanostructure (TiO2/CdS) could function as a photoanode for solar cell applications. A sandwich-type solar cell device was compiled and its photovoltaic performance tested under light illumination (AM 1.5G, 50 mW cm−2) and the photovoltaic behaviour is also discussed herein.
Fig. 2 shows the X-ray diffraction (XRD) patterns of FTO, TiO2 and TiO2/CdS nanostructures for different number of immersions. The most intense diffraction peak at 2θ = 25.3° is indexed to the tetragonal crystal structure of TiO2 with an anatase phase (JCPDS card 84-1285) and lattice orientation along the (101) plane. Additionally, some small peaks appear along the (004), (200) and (105) planes related to the seminal crystal structure and phase of TiO2, denoted by the (T) symbol. Due to the high intensity of the TiO2 and FTO peaks, the effect of the CdS nanoparticles coating on TiO2 is diminished, even though, some tiny peaks appear at 27.8°, 36.1° and 43.8° in Fig. 2, confirming the presence of another layer coated on the TiO2 electrode. In the XRD patterns, the diffraction peaks along the (420), (305) and (444) crystal planes correspond to the orthorhombic structure of CdS (JCPDS: 47-1179) and are denoted in the figure by the (C) symbol. The broadness in tiny peaks itself revealed, the nanocrystalline nature of the CdS material deposited on TiO2 surface. Thus, the formation of TiO2 along with CdS was confirmed from the XRD patterns. This result confirms the successful deposition of CdS on the TiO2 film. The other peaks denoted by the triangles (Δ) symbol are due to the FTO substrates. The average crystallite size was calculated by the well-known Debye–Scherrer equation:
![]() | (1) |
![]() | ||
Fig. 2 X-ray diffraction patterns of the FTO, TiO2 and CdS coated FTO/TiO2 for 8, 13, 15 and 20 SILAR cycles. |
Fig. 3 shows the high-magnification SEM images of mesoporous TiO2 electrode at different stages of modifications with a scale bar of 100 nm. These images provide a perspective of the electrode morphology and its ability to anchor the CdS nanoparticles on the mesoporous TiO2 electrode. Fig. 3(a) shows the growth of a few small grains on and in inside 20–30 nm pores on the TiO2 surface for eight immersions. The TiO2 surface was further modified by coating CdS nanoparticles for 13, 15 and 20 immersions, as depicted in Fig. 3(b)–(d), respectively. From Fig. 3(b), it can be seen that the CdS quantum dots (size < 10 nm) are clearly visible on the TiO2 surface after 13 immersions. The inset in Fig. 3(b) shows a bigger white circle corresponding to TiO2, while the small dots on the TiO2 belong to CdS. Further increases in the number of immersions causes the aggregation of CdS nanoparticle on the TiO2 mesoporous network. Most of the TiO2 surface areas are covered by CdS nanoparticles, resulting in a cauliflower-like morphology (Fig. 3(c) and (d)). This may be due to the growth of the CdS grain not being on to the TiO2 surface but rather on pre-grown CdS particles, which act as secondary nucleation centres and finally lead to the cauliflower-like morphology. These results demonstrate that the synthesised mesoporous TiO2 thin film provides a high surface area on which the CdS particles grow, resulting in the nano to quantum structure. The surface area was calculated by using the following formula:
![]() | (2) |
TiO2/CdS SILAR cycles | Jsc (mA cm−2) | Voc (mV) | FF (%) | η (%) | VFB (mV) | Rs (Ω) | Rsh (kΩ) | Surface area (m2 g−1) |
---|---|---|---|---|---|---|---|---|
8 | 0.398 | 266 | 43.4 | 0.091 | −155 | 276 | 1.19 | 33.09 |
13 | 0.707 | 336 | 35.1 | 0.166 | −256 | 265 | 1.21 | 29.87 |
15 | 1.114 | 384 | 34.8 | 0.297 | −301 | 261 | 1.34 | 26.84 |
20 | 2.026 | 402 | 35.8 | 0.579 | −423 | 176 | 1.47 | 17.64 |
For the TEM images, the nanostructure architecture was scratched without disturbing the FTO substrate and placed on a copper-coated carbon grid. Fig. 4(a) shows a typical TEM image of the TiO2 coated with CdS quantum dots (size < 10 nm) for 13 immersions. The surface coating in LHS and the surface coverage with uniformity in the RHS image can be clearly seen with a size of less than 10 nm. The inset of Fig. 4(a) shows the evidence for the presence of Cd, Ti, S, O, C and Cu elements in the deposited material. The presence of Cu and C is due to the copper-coated carbon mesh. Fig. 4(b) depicts the HR-TEM image of the CdS-coated TiO2 from 13 SILAR immersions. The inter-planar spacing (‘d’ value) was estimated as 0.34 nm and was assigned to the anatase phase of tetragonal TiO2 oriented along the (101) plane. This supports well the plane observed at 25.3° in the XRD studies for TiO2 (JCPDS card 84-1285). As a typical example, the clear existence of CdS quantum dots on TiO2 can be seen from the high-quality HR-TEM image shown in Fig. 4(c). The inter-planar distance of 0.24 nm was calculated and assigned to the orthorhombic crystal structure of CdS, which corresponds to the (305) plane as revealed from the XRD analysis (JCPDS card 47-1179). The particle size for 13 immersions was estimated to be 4–6 nm, which is responsible for the red-shift due to the size quantization effect observed in the optical absorption curve. Thus, the successful formation of CdS nanoparticles on mesoporous TiO2 nanoparticles was confirmed.
The absorbance spectra were recorded in the 300 to 700 nm wavelength regime of incident light. The absorption spectra of bare TiO2 and four different-sized CdS nanoparticles-coated TiO2 networks are illustrated in Fig. 5, with FTO as a reference. A sharp increase in the absorption spectra at shorter wavelengths (below 380 nm) can be observed for TiO2, suggesting that the absorption band edge would facilitate its application as a wide band gap material. Fig. 5 depicts the absorption spectra of the CdS-coated TiO2 film for the different number of immersions in the SILAR technique. Swelling in the absorption curves between 450 and 500 nm is clearly evident and is boosted with the increase in the number of SILAR immersions; this clearly indicates a red-shift in the absorption edge. By noting the absorption maxima at 451, 461, 472 and 485 nm for the 8, 13, 15 and 20 SILAR immersions, the respective band edge transition energies were calculated at the respective absorption maxima and were found to be 2.74, 2.68, 2.62 and 2.55 eV.26 The particle sizes of CdS with the different number of SILAR immersions were calculated from the absorption maxima by utilizing the empirical fitting function reported by Yu et al.3 for CdS and were found to be 5.52, 5.66, 6.18 and 6.81 nm for the respective maxima of the absorption peaks. The particle size distribution should result in sharp absorption peaks appearing in the absorption spectra measured for the particles in solvents; however, this is not seen in Fig. 5 and instead small swellings in the absorption spectra were observed. This may be attributed to the fact that the increase in the number of immersions causes the agglomeration of small-sized particles resulting in bigger ones, which are responsible for the shift in the optical properties towards the bulk system.
![]() | ||
Fig. 5 UV-Vis absorption spectra for FTO/TiO2 and CdS coated on the FTO/TiO2 electrode in different numbers of SILAR immersions. |
A sandwich cell was prepared by using TiO2 and TiO2/CdS as a photoanode and a platinum-coated conducting glass electrode as the counter electrode. The two electrodes were separated by a thin polyethylene film (50 μm) as a spacer. The empty cell was tightly held and then the edges were heated to 130 °C to seal the two electrodes together. The composition of the polysulphide electrolyte was 1 M Na2S, 1 M sulphur powder and 1 M NaOH. The electrolyte was introduced into all the fabricated cells by using capillary action. The active surface area of the photoanode was 0.16 cm2. In addition to enhanced light absorption, the formation of a TiO2 and TiO2/CdS nano-heterojunction in the hybrid structure also facilitated separation of the photogenerated electrons and holes, as shown in Fig. 6(a).
The effectiveness of the CdS over TiO2 for different rinsing cycles and the use of bare TiO2 as a photoanode in photoelectrochemical QD-sensitized solar cells were assessed utilizing EQE measurements. On increasing the number of immersion cycles due to the enhanced absorption of a number of photons, the increasing growth of CdS results in a steady growth of the short-circuit current density. This gradually increases the value of the EQE for the different number of immersions of the TiO2/CdS samples, which may be due to the charge injection of highly efficient light harvesting CdS QDs into porous TiO2, as shown in Fig. 6(b), in which the highest EQE value of 16% was obtained for 20 SILAR cycles of CdS on TiO2. This leads to more electron collection, which inhibits electron recombination and enhances the electron scattering ability. Furthermore, this may result in the easy transfer of electrons to the conduction band of TiO2 and to the electrolyte.
The experimental current density–voltage (J–V) characteristics measured from the TiO2 and TiO2/CdS photoelectrode are shown in Fig. 6(c). For all the cases, the photoelectrochemical cells' significant activities in the dark were shown by the illustrated heterojunction properties. In the third quadrant of the figure, a sharp increase in current density can be observed, which might be due to penetration of the liquid electrolyte through the pores of TiO2 and it then touch the back platinum contact. This is rather a good indication of the formation of a nano- or mesoporous TiO2 network, which is necessary to coat CdS particles from the top to the bottom. Under illumination, TiO2 showed a slight photovoltaic effect (photocurrent density of ∼88 μA cm−2) due to the absorption of UV light in the nano- or mesoporous network. For comparison, the photovoltaic curves for CdS with different numbers of layers over TiO2 are illustrated. A significant enhancement in the short-circuit current can be seen with the increase in the number of SILAR immersions of CdS along with an enhancement in the open-circuit voltage. This gives a maximum short-circuit current value of 2.02 mA cm−2 for 20 immersions, with an efficiency of 0.579%. This enhancement is strictly attributed to not only the enhancement in the absorption values with respect to the increase in layer thickness but also to the agglomeration of small-sized particles to form bigger ones, which are responsible for the shift in the optical band gap to lower energies due to the quantum size effect. The more filling of CdS inside the nanoporous TiO2 layer is also seen by the decrease in the tendency of sharp increase in reverse current, which mainly avoids direct contact between the liquid electrolyte and back contact. The increase in the photovoltage is well attributed to the covering of the TiO2 layer with more CdS with the increase in the number of immersions. The photovoltage generated in the semiconductor sensitizer solar cell is characteristic of the recombination centres in the semiconductor sensitizer solar cell. The increase in the thickness of the CdS layer on TiO2 with the increasing number of SILAR immersions simultaneously provides more recombination's centres and hence the photovoltage and photocurrent increase with the number of SILAR immersions.27,28 A straightforward technique was used to estimate Rs and Rsh from the solar cell photovoltaic output curves by using the following relations29 and the results are given in Table 1
![]() | (3) |
![]() | (4) |
The decrease in Rs is direct evidence for the photogenerated charge carriers. The photovoltaic output increases and the obtained values of Rs decrease with the increase in the number of SILAR immersions (i.e. the increasing thickness of the absorber layer), which supports well the separation of charge carriers across the TiO2/CdS interface, whereby CdS can act as a sensitizer to generate the charge carriers. Hence, a boost in device efficiency is observed with increasing the number of SILAR cycles. Also, Rsh is affected by both the charge carrier recombination and short circuiting within the cell. Therefore, a higher coating layer for the heterostructure electrodes can result in a pinhole-free surface, thus providing less access for electrolyte contacts to the FTO substrate. The increase in Rsh values with the increasing number of SILAR immersions i.e. with the coating of the CdS layer, reduces the participation of FTO in PEC performances and increases the number of available recombination centres. The photovoltaic parameters, such as the current density (Jsc), open-circuit voltage (Voc), fill factor (FF), power conversion efficiencies (η%) series resistance (Rs) and shunt resistance (Rsh), are depicted in Table 1. The photovoltaic conversion efficiency was calculated using the following formula:
![]() | (5) |
Beyond 20 SILAR cycles of CdS over TiO2 (as shown in Fig. S1†), it was found that the thickness of the electrode increases, and as a result of the light is transmitted into the depth of the electrode, which leads to a decrease in the light intensity. This phenomenon directly lowers the excessive electron density, resulting in the lowering of Voc, as shown in (Table T1†).30 This then results in the increase in Rs. Now, due to the greater thickness of the electrode than the light penetration depth, the recombination centres increase and electron loss takes place. This effect directly decreases the value of Jsc of the cell. The measurement of capacitance as a function of applied voltage provides useful information about photoelectrodes, such as about the flat band potential (VFB). VFB is an important factor for photoelectrochemistry and facilities the energetic position of semiconductor materials. The value of VFB can be obtained by using the Mott–Schottky relation:
![]() | (6) |
In order to investigate the kinetics of the photoelectrochemical process, electrochemical impedance spectroscopy (EIS) was performed at a forward bias potential of −0.4 V under dark conditions. Under this condition, the QDSSC behaves as a leaking capacitor.34 A frequency scan of 0.01 to 104 Hz was chosen for the investigations. The nature of the photoelectrochemical plots, such as the Nyquist and Bode plots, for bare TiO2 and TiO2/CdS QDSSC was quiet similar those reported by Sudhagar et al. where they demonstrated self-assembled CdS quantum dot sensitised TiO2 nanospheroidal solar cells without using any linker.35 Fig. 7(a) shows that the Nyquist plots were first semicircular arcs in the high frequency range, reflecting the charge transfer resistances (R1) and the constant phase element (Q1) at the Pt/electrolyte interface. The second semicircle towards the higher frequency region can be ascribed to the interfacial charge transfer (R2) and constant phase element (Q2) at the TiO2/CdS/electrolyte interface, while (R) represents the serial resistance, which is determined by the sheet resistance between the TiO2 layer and FTO substrate. Q1 and Q2 basically describe a non-ideal frequency dependent capacitance due to the non-uniform distribution of the current by the material heterogeneity. The inset of the figure is the simple equivalent electric circuit that was used to fit the EIS data to explain the parameters such as the electron lifetime and the recombination of the TiO2/QDs film.36,37 For the TiO2/CdS system, Lee et al. also showed the same equivalent circuit diagram and controlled the rate of recombination by the incorporation of a ZnS layer.38 It can be seen from Table 2 that the value of (R) decreases from the bare TiO2 layer to the TiO2/CdS layer (at 20 SILAR cycles). The value of (R1) at the high frequency region was found to decrease as the number of SILAR cycles increases. Besides this, the second semicircular arc of the Nyquist plot was found to decrease along the radius, which apparently decreases R2, which suggests that the QDs are coated on the surface of the highly porous TiO2 with a high photocurrent density.39 This might effectively decrease the ohmic polarization and impart a better interfacial charge transfer. The lowest value of R2 for TiO2/CdS at 20 SILAR cycles indicates the easy transport and effective charge collection, with a very low chance of recombination and thus an enhanced electron transfer rate and high power conversion efficiency.40
![]() | ||
Fig. 7 (a) Nyquist plot in the dark and (b) Bode plot of bare TiO2 and CdS coated on the FTO/TiO2 electrode in different numbers of SILAR immersions. |
Sl. no. | Label | R1 (kΩ) | R2 (Ω) | τe (ms) | keff (s−1) |
---|---|---|---|---|---|
1 | TiO2 | 10 | 1038 | 126 | 7.93 |
2 | 8 cycles CdS/TiO2 | 8.85 | 946 | 201 | 4.93 |
3 | 13 cycles CdS/TiO2 | 8.49 | 813 | 253 | 3.95 |
4 | 15 cycles CdS/TiO2 | 7.25 | 725 | 323 | 3.09 |
5 | 20 cycles CdS/TiO2 | 5.59 | 676 | 496 | 2.01 |
Fig. 7(b) shows the Bode plot for the bare TiO2 and TiO2/CdS at different SILAR cycles. Interestingly, it was been found that on increasing the number of SILAR cycles over bare TiO2, degradation of the TiO2 surface traps can occur, which passivates the Bode plot towards the low frequency region. Again from Table 2, the values of the electron lifetime (τe = 1/2πfmax) increases, which in turn decreases the effective recombination rate (keff = 1/τe).41 This may be the cause for the increase in the value of the open-circuit voltage (Voc), since the increase in the carrier recombination is the prominent factor for the reduction of the Voc. Further recombination rate increases lead to high FF values, which is in line with the photovoltaic performance data.
Cd(CH3COO)2 + CH3OH → Cd2+ + 2CH3COO− + CH3OH | (7) |
Meanwhile, an anionic solution was prepared by dissolving 2.4 mg of sodium sulphide flakes into 50 ml methanol, giving a resultant solution of pH 7.6. This formed a polar solution containing polar S2− ions in solution, as shown in the following reaction:
Na2S + CH3OH → S2− + 2Na+ + CH3OH | (8) |
The simplified version of the ion-by-ion mechanism of the chemical reaction for the growth of CdS nanoparticles onto the surface of TiO2 is as follows:
Cd2+ + S2− → CdS | (9) |
The growth of the nano-sized CdS nanoparticles on the mesoporous TiO2 electrode was controlled by varying the number of vertical immersions of the TiO2 film into the separate anionic and cationic solutions. In the present work, the growth of CdS nanoparticles, which constitutes the stages of the thin film growth depending on number of SILAR immersions as 8, 13, 15 or 20, is discussed herein.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra14275c |
This journal is © The Royal Society of Chemistry 2016 |