Next Article in Journal
Palladium Comprising Dicationic Bipyridinium Supported Periodic Mesoporous Organosilica (PMO): Pd@Bipy–PMO as an Efficient Hybrid Catalyst for Suzuki–Miyaura Cross-Coupling Reaction in Water
Next Article in Special Issue
One Simple Strategy towards Nitrogen and Oxygen Codoped Carbon Nanotube for Efficient Electrocatalytic Oxygen Reduction and Evolution
Previous Article in Journal
Tuning the Catalytic Properties of Copper-Promoted Nanoceria via a Hydrothermal Method
Previous Article in Special Issue
Ternary N, S, and P-Doped Hollow Carbon Spheres Derived from Polyphosphazene as Pd Supports for Ethanol Oxidation Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis and Characterization of Two Dimensional SnO2-Decorated Graphene Oxide as an Effective Counter Electrode in the DSSC

by
Mohamed S. Mahmoud
1,2,
Moaaed Motlak
3 and
Nasser A. M. Barakat
1,*
1
Chemical Engineering Department, Minia University, El-Minia 61111, Egypt
2
Collage of Applied Science, Department of Engineering, Sohar 311, Oman
3
Department of Physics, Collage of Science, University of Anbar, Al Ramadi 31001, Iraq
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(2), 139; https://doi.org/10.3390/catal9020139
Submission received: 1 December 2018 / Revised: 18 January 2019 / Accepted: 22 January 2019 / Published: 1 February 2019
(This article belongs to the Special Issue Immobilized Non-Precious Electrocatalysts for Advanced Energy Devices)

Abstract

:
SnO2-decorated graphene oxide (SnO2/GO) was synthesized by the modified Hummers’s method, followed by a chemical incorporation of SnO2 nanoparticles. Then, the nanocomposite was used as anon-precious counter electrode in a dye-sensitized solar cell (DSSC). Although GO has a relatively poor electrical conductivity depending essentially on the extent of the graphite oxidation, presence of SnO2 enhanced its structural and electrochemical properties. The Pt-free counter electrode exhibited a distinct catalytic activity toward iodine reduction and a low resistance to electron transfer. Moreover, the decorated GO provided extra active sites for reducing I3 at the interface of the CE/electrolyte. In addition, the similarity of the dopant in the GO film and the fluorine-doped tin oxide (FTO) substrate promoted a strong assimilation between them. Therefore, SnO2-decorated GO, as a counter electrode, revealed an enhanced photon to electron conversion efficiency of 4.57%. Consequently, the prepared SnO2/GO can be sorted as an auspicious counter electrode for DSSCs.

1. Introduction

Basically, the conversion of solar energy into electricity proceeds through direct or indirect routes. Currently, direct conversion of solar radiation-to-electricity is receiving the maximum attention. In this regard, silicon-based photovoltaic cell (PV panel) is the most widely used. However, PV cells are suffering from high fabrication cost ($/kW) because their function counts on the presence of a relatively thick layer of doped silicon to obtain an acceptable photon capture rate. Comparably, The dye sensitized solar cell (DSSC) provides a low-cost transfer route of photons to electrons, due to its superior advantages over the traditional PV cells; it absorbs more solar radiation per surface area than the traditional PV cells [1,2]. Recently, the efficiency of the DSSCs has been increased to 17%, which puts it in the market race with the conventional PV [3]. Mainly, the DSSC consists of a porous layer of titanium oxide nanoparticle coated with a solar sensitive organometallic dye, the counter electrode (Pt), and the working electrolyte (e.g., the redox couple I3/I). Its mechanism of work is discussed in details elsewhere [4,5]. Numerous efforts to enhance the solar energy-to-electricity efficiency of the DSSCs have been reported. These attempts focused on optimizing the properties of the four components of the DSSC:
  • The sensitized dye
  • The material of the photoanode
  • The redox couple electrolyte and
  • The non-precious counter electrode
In this regard, there have been several attempts to develop stable solar-sensitive dyes [3,6] and electrolytes [6,7]. Similarly, developing the photoanode has been investigated by altering the type of the mesoporous film that is coated over the photo anode glass, such as TiO2 nanofibers [8,9], ZnO and ZnO/TiO2 nanocomposite [10,11], SnO2 [12,13], or CdO [14]. These oxides are mixed with the organometallic dye to form the photoanode of the DSSC.
Compared to the photoanode, the counter electrode (CE) is more expensive because it is usually fabricated from precious metals (e.g., Pt), which adds an additional capital cost. Therefore, developing effective counter electrode from cheap materials can strongly enhance the DSSC’s rank. Typically, the CE should assure three attributes [15]: (i) As a catalyst, it should transfer the electron to the oxidized redox couple (I). (ii) As a cathode, it should collect the electrons coming from the outside circuit and get them ready to be transferred to the cell. (iii) As a mirror, it should reflect the transmitted light back to the DSSC to enhance the use of photons [16]. The main characteristics of the optimal CE are summarized as follows [17]: high catalytic activity, high electrical conductivity, maximum reflectivity, low-cost, large surface area, porous nature, optimal thickness, electrochemical and mechanical stability, energy level that matches the potential of the redox couple electrolyte, and high adhesivity with the FTO [17]. As a heterogeneous catalytic reaction is taking place on the surface of the counter electrode, the reduction of the electrolyte (e.g., iodine ion) can be considered a combination of adsorption and electrochemical reaction processes. Accordingly, to exploit their high adsorption capacity, carbon nanomaterials such as nanotubes [18], graphene [19], and nanofibers [20] have been utilized as support for the counter electrode materials. Moreover, some carbonaceous nanostructures show a distinguished activity individually [21,22]. Graphene is branded by its amazing electrical and mechanical properties. Therefore, many researchers show interest in using this nanomaterial in different applications [23]. However, its high hydrophobicity negatively affects the performance in the aqueous media due to the poor contact which adds an additional resistance to the reactants transfer. On the other hand, due to possessing oxygenated groups on the surface, graphene oxide (GO) is more hydrophilic than graphene. However, the newly added active groups decrease the electrical conductivity of GO compared to graphene. The conductivity of GO depends on the C/O ratio in the sample. During the oxidation step, the sp2 carbons are removed and replaced by sp3 ones having oxygen functionalities. This process creates a band gap by pulling the bands apart. Hence, when graphene is fully oxidized, the bands are far apart, and the GO behaves as an insulator. In the middle of this fully oxidized GO and pure graphene, it behaves as a semiconductor. Partially oxidized GO shows the appearance of sp3 regions and sp2 regions [24], and the GO can conduct through these sp2 regions via Klein tunneling [24]. In addition, insertion of a transition metal oxide in the GO layers could play a crucial role in enhancing the electron conduction of the GO. SnO2 is a crucial semiconductor that is used as a transparent conducting oxide (TCO). This TCO normally coated over the glass substrate and used for optical applications. Hitherto, SnO2 is usually considered as an oxygen-deficient n-type semiconductor [25]. Therefore, doping GO with SnO2 nanoparticles may enhance the conductive properties of the GO and promote the electron transfer to the redox chattel in the DSSC. Even though there are intensive studies on alternatives of counter electrode, thus far, to our best knowledge, there is no report regarding the application of SnO2-decorated GO as counter electrode in the DSSC. In this work, we report the synthesis of SnO2-decorated GO, its characterization, and application as a counter electrode in the DSSC.

2. Results and Discussion

Figure 1a shows the transmission electron microscope (TEM) image of the prepared material. It is noticeable that the sample composed of a sheet of GO contains randomly distributed nanoparticles. High resolution TEM (HR-TEM) image shown in Figure 1b affirms that the attached nanoparticles appear as textured crystalline mode, which indicates that these nanoparticles are SnO2-based, highly crystalline compound. However, the distribution of these nanoparticles is random. The higher magnification image indicates that the interlayer distance is 0.26 nm and the particle size is less than 20 nm.
Figure 2 displays the X-ray diffraction (XRD) patterns of both pristine and decorated GO. Initially, from XRD pattern, it is easy to confirm the formation of GO from the utilized precursor. Typically, the graphite reveals a sharp peak in the XRD pattern at 2θ value of ~26°. Due to exfoliation, the graphite-identified peak disappears and another peak is formed at ~10°, while graphene is distinguished by a very broad peak centered at ~24° [26]. As shown, the GO main peak at 2θ values of 10.2°, corresponding to the (001) basal plane of GO with a d-spacing value of (d001 = 0.961 nm), appears in both samples. However, the intensity of such peak in the SnO2/GO sample has a lower intensity than pure GO. Typically, the intensity of the GO main identification peak is 867 and 1957 count/s for the SnO2/GO and pristine GO, respectively. The reason of low peak intensity is probably due to the heat treatment and incorporation of foreign atoms in the GO layer, which was also noticed by R. Krishna et al. [27]. In addition, the strong diffraction peaks at 2θ values of 26.38°, 33.48°, 37.76°, and 51.46° corresponding to (110), (101), (200) and (211) crystal planes (JCPDS:41-1445), respectively, confirm the incorporation of tetragonal crystals of SnO2 inside the GO layer. It is also noticeable that the SnO2 peaks are broadened, indicating the presence of small crystalline size. It is important to determine the crystal structure of the SnO2-decorated GO due to its direct effect on the power conversion efficiency of the DSSC. According to Scherer’s equation (τ = Kλ/βcosθ, where τ is the ordered (crystalline) domains mean size, λ is the wavelength of the utilized X-ray irradiation (0.1504 nm), K is a constant value with a typical value of 0.9 referring to the shape factor, β is the line broadening at half the maximum intensity (FWHM) in radians and θ is the Bragg angle), the average grain size was determined to be 5.1 nm.
Figure 3 displays the Raman spectra of pristine and SnO2-incorporated GO in the range of 100–3000 cm−1. In the GO spectrum, two protuberant peaks corresponding to D and G bands, the characteristic peaks of GO, are located at 1354 and 1591 cm−1 with respective ID/IG intensities ratio of 1.07. The inset confirms the stability of the prepared GO during the achieved GO decoration process by SnO2 nanoparticles. As shown in the inset, there is no considerable change in the ID/IG intensities ratio. Moreover, in the Raman spectrum, the pristine graphene reveals an additional peak (2D), indicating an increase in the average size of sp2 domains at ~2790 cm−1 [28]. As shown, the 2D peak does not appear in the SnO2-GO spectrum, which confirms maintaining the composition of the GO. In other words, the utilized treatment process did not affect the content of the oxygenated groups covering the surface of GO sheets. The crystalline structure of SnO2 is a tetragonal rutile with point group D4h [29]. Typically, three modes identifying SnO2 can be found in the Raman spectrum: 474 cm−1 (Eg), 631 cm−1 (A1g) and 775 cm−1 (B2g) [30]. When the particle size decreases, A1g and B2g modes of SnO2 are shifted to lower wave numbers and Eg mode is shifted to higher wave number [30]. The obtained results show peaks at 615, 435, 245, and 120 cm−1. The band at 435 cm−1 probably corresponds to the (Eg) mode of the oxide. Similarly, the mode at 615 cm−1 can be assigned to symmetric O-Sn-O stretching (A1g) shifts due to the relatively small particle size of SnO2 nanoparticles. The position and the intensity of SnO2 peak in the Raman spectrum depends on the crystal size. The vibrational mode (B1g) peak appears only along with nanomaterials. Thus, in the Raman spectrum, the presence of a peak at 120 cm−1 is associated with non-degenerated B1g mode of SnO2 and its appearance is caused by lowering the surface phonon frequencies that leads to increase in the inter-atomic distance on the surface [30].
Overall, based on the aforementioned analytical technique, it is safe to claim that the proposed material is GO sheets decorated by small and highly crystalline SnO2 nanoparticles.
It is worth mentioning that the function of the redox shuttle (I3/I) is very important to ensure a stable transfer of electrons between the photoanode and the counter electrode of a DSSC. The photon-induced excitation of the dye leads to transfer of electrons to the oxide semiconductor. Then, the electron donor (I3) must rapidly return the excited dye back to its ground state. Afterwards, the electron acceptor (I) should move to the CE to get the missing electrons to regenerate the electron donor (I3). The electron migration through the external circuit will close the circuit [2,4].
To verify the electrocatalytic activity of the as-prepared SnO2-incorporated GO, cyclic voltammetry (CV) analysis of the prepared catalyst was achieved using a conventional three-electrode system. SnO2-doped GO, Pt wire and Ag/AgCl served as working, counter and reference electrodes, respectively. The electrolyte solution consisted of 10 mM LiI, 1 mM I2, and 0.1 M LiClO4 in acetonitrile. Figure 4a shows the cyclic voltammograms (CV) at scan rate of 25 mV/s. It is clear that the introduced SnO2/GO shows both oxidation and reduction peaks, which indicate achieving the following reactions [20].
I3 + 2e → 3I
3I2 + 2e → 2I3
The counter electrode exhibits high redox current densities, indicating that the speed of the redox reaction on the surface of the modified GO is fast. Besides the activity of the SnO2 nanoparticles attached on the surface of the GO, the enhancement in the surface area after the decoration process can be assigned as a second reason for the observed high performance. Typically, the measured surface area of the pristine and decorated GO was 159.7 ± 2 and 210.5 ± 3 m2/g, respectively. The increase in the surface area can be attributed to the high exfoliation rate upon addition of the tin precursor. Therefore, the improved specific surface area of the modified GO can be considered as a stimulator for such boost of the electrocatalytic activity toward the reduction of I3 ions. Incidentally, the modified GO is responsible for the efficient reduction of I3 ions by providing sufficient superficial area and promoting fast electronic transformation, which results in higher Jsc and solar-electrical conversion efficiency in the DSSC.
Figure 4b illustrates CV curves of the SnO2-incorporated GO electrode in the prepared electrolyte solution at different scan rates (10–50 mV/s). It is clear that altering the scan rate increased the current density values. In addition, it slightly shifted the potential peaks of the anodic and cathodic reactions. The inset of Figure 4b shows the influence of the scan rate on the anodic and cathodic peak currents. It is noticeable that they are linearly dependent on the scan rate. Such finding discloses that the rate of diffusion of I3 ion is the hindering step in the oxidation–reduction process [20].
Performing the electrochemical impedance spectroscopy (EIS) analysis of a DSSC helps to get insight about some important performance parameters, such as charge transport through the photoanode and near the counter electrode, charge transfer due to electron back reaction, and capacitive accumulation of the charges at different processes in the cell. It was used to measure the resistances of the individual interfaces of the fabricated DSSC based on the introduced SnO2/GO counter electrode. The merit of EIS spectrum over the I–V curve is that EIS indicates the response of the solar-induced current on the electrical bias voltage values, which cannot be indicated through the I–V curve.
The impedance spectrum of a DSSC typically shows three semicircles in the Nyquist plot. According to the direction of decreasing frequency, the first semicircle corresponds to the charge transfer processes which happens in the CE/electrolyte interface with a characteristic frequency ωCE; the second or middle semicircle corresponds to the diffusion of the electron through the TiO2 layer and the reversible electron reaction with the oxidized state of the redox species at the TiO2/electrolyte boundary; and the third semicircle at the low frequency region corresponds to the diffusion of I3 in the electrolyte, with a characteristic frequency ωD [31]. The characteristic frequency for electron transport or diffusion (ωd) appears in the middle semicircle at the high frequency region, while the peak frequency (ωk) of that semicircle corresponds to the electron reversible reaction. Figure 5 shows the Nyquist and Bode plots of GO and SnO2/GO as counter electrodes in the DSSC. In the case of GO, as shown in Figure 5A, only one semicircle appears, indicating an insignificant charge transfer at the GO/electrolyte interface. For the SnO2/GO plot, two semicircles signify two different kinds of impedance. In the high-frequency region (103–106 Hz), the semicircle denotes the resistance due to the charge transfer (RCE) at the counter electrode/electrolyte interface. In the middle-frequency (1–103 Hz) region, the semicircle denotes the diffusion through the TiO2 layer and electron reversible reaction with the oxidized state of the redox species at the TiO2/electrolyte boundary (RCT). However, in the low-frequency region (0.1–1 Hz), no semicircle of the complete response appears, which could be attributed to the small distance between the working and counter electrodes and the low viscosity of the electrolyte.
These data indicate that the low (RCE) is related to a comparatively high rate of electron transfer at the interface of electrolyte/counter electrode, which leads to improve the catalytic activity toward the redox couple in the electrolyte. In the DSSC cell, the (RCE) value was found to be 43.1 Ω, which indicates that the SnO2/GO revealed an effective catalytic activity toward the reduction of I3 ions in redox electrolyte. Moreover, the cell shows low series resistance (Rs) value of 23.6 Ω, which is attributed to an excellent conductivity of SnO2/GO film and strong contact between the film and FTO substrate [20].
The catalytic activity of the SnO2-incorporated GO counter electrode can be presented using current density (J), which is calculated from the charge transfer resistance (RCT) [32]:
RCT = RT/nFJ
where R is the gas constant (J/mol·K), T is temperature (K), n is the number of involved electrons (n = 2), and F is the Faraday constant.
Figure 6 depicts the J–V curves of DSSCs under metal halide lamp illumination. Table 1 displays the photovoltaic parameters from J–V curves and a brief comparison with other counter electrodes. The data indicate that the cell based on the introduced SnO2/GO exhibited an energy conversion efficiency of 4.57%, which is slightly lower than the standard Pt counter electrode. Such energy conversion efficiency of the cell is linked to the activity of the counter electrode, which synergizes the reduction reaction at the interface of electrolyte/counter electrode, which in turn assists the regeneration of the ground state of the organometallic dye at the interface of electrolyte/photoanode. As shown in Table 1, the other parameters for the fabricated cell are Voc of 0.67 V, Jsc of 14.04 mA−2 cm and FF of 0.49. These results are lower than those reported from GO–Pt composite fabricated by low-temperature electrodeposition process, which revealed an energy conversion efficiency of 7.05% [33]. However, the authors used Pt precious metal, which is not aligned with our goals.

3. Materials and Methods

3.1. Catalyst Preparation

To prepare the electrode, all the involved chemicals, e.g., SnCl2, thiourea, graphite, and KMnO4, were obtained from Sigma Aldrich Co., Seoul, South Korea. The procedure started with slowly dissolving 0.42 g of tin chloride in 20 mL distilled water. Parallelly, 0.15 g of thiourea were also dissolved in 20 mL distilled water individually for 20 min. Then, the two solutions were mixed using a magnetic stirrer for 1 h. The GO was prepared by a modified Hummers’s method. The next step was to disperse 0.2 g of GO in 100 mL distilled water, and then ultra-sonication for 40 min. Afterwards, the two mixtures were mixed for 20 min, and then 200 mL of hydrazine were added just before the hydrothermal step. Next, hydrothermal treatment was done for 10 h at 180 °C. Finally, the sample was filtered and dried at 60 °C overnight.

3.2. The DSSC Fabrication

The fabrication of DSSCs consisted of three steps [18,39]: fabrication of the photoanode, fabrication of the counter electrode and the addition of the working electrolyte. Typically, FTO substrate glass (FTO 10 Ω/sq.) was used to prepare the photoanode. Preparation started by plating nanocrystalline TiO2 (Degussa P-25) thin film over FTO glass by simple doctor blade technique. The prepared photoanode had an active area of 0.25 cm2 with the film thickness of 8–10 μm. Then, the FTO was annealed at 450 °C for 30 min. Afterwards, the photoanodes were immersed in the dye solution consisting of 0.3 mM ruthenium(II) 535 bis-TBA (N-719, Solaronix; Sigma, Seoul, South Korea), rinsed with ethanol and dried for 24 h under nitrogen flow. Similarly, the SnO2-decorated GO was pasted over the FTO glass by simple doctor blade technique. The obtained film was cleaned by ethanol and then dried at 60 °C for 30 min. To complete the fabrication of DSSC, the SnO2-decorated GO counter electrode was placed over the dye-adsorbed TiO2 photoanode and sealing was made using Surlyn sheet (SX 1170-60, Solaronix, 60 μm thick) between the two electrodes. Finally, the electrolyte, consisting of lithium iodide (LiI 0.5 M), Iodine (I2 0.05 mM), and tetra-butyl-pyridine (C9H13N 0.2 M) in acetonitrile, was inserted into the holes in the counter electrode using a syringe.

3.3. Characterization and Application

The TEM images of the prepared catalyst were retrieved using JEOL JEM 2010 transmission electron microscope, working at a voltage of 200 kV, (JEOL Ltd., Tokyo, Japan). Clarification of the distance of the TEM images was acquired by ImageJ 1.47v software. To determine the crystal structure of the prepared catalyst, Rigaku X-ray diffractometer (XRD, Rigaku, Tokyo, Japan) with Cu Kα (λ = 1.5406 Å) was utilized. The Raman spectrum was also acquired using Dispersive Raman spectrometer (BRUKER-SENTERRA, Boerdestr, Warstein, Germany) equipped with an integral microscope (Olympus, Tokyo, Japan). The excitation source was neodymium-doped yttrium aluminum garnet (Nd/YAG) laser (532 nm), and providing a power of 20 mW on the sample, and analyzed by X-ray Diffraction Microanalysis (XPERT-PRO-P-Analysis, The Woodlands, TX, USA). The electrochemical impedance spectroscopy (EIS) and cyclic Voltammetry (CV) measurements were performed using a VersaSTAT 4 (AMETEK, New York, NY, USA) electrochemical analyzer using a conventional three-electrode cell structure. In the utilized cell, glassy carbon, Pt and Ag/AgCl served as working, counter and reference electrodes, respectively. To deposit the proposed GO-based material on the active surface of the working electrode, 2 mg of the function material were added to 20 µL Nafion solution (5 wt %) and 400 µL isopropanol. Afterwards, the suspension was subjected to utltra-sonication process for 20 min. Then, 15 µL of the suspension were poured in three steps on the electrode active area. Finally, the electrode was dried at 80 °C. Current–voltage characteristics of DSSCs were measured by using digital Multimeters (Model 2000, Keithley, Filderstadt, Germany) and a variable load. As a simulation of the solar radiation, a 1000 W metal halide lamp served as a light source, and its light intensity (or radiant power) was adjusted to simulate mornign. Then, 1.5 radiation at 100 mW cm−2 with a Si photo detector fitted with a KG-5 filter (Schott, Cheney, KS, USA) as a reference was calibrated at NREL (New York, NY, USA). The metal uptake, during the decoration process, was determined by estimating the remaining tin ions in the initial solution using the titration procedure. Briefly, the solution was hot treated with hydrochloric acid, and then drops of starch indicator solution were added. Finally, the solution was titrated against iodate–iodide standard solution. The uptake was determined to be around 15 wt %.

4. Conclusions

SnO2 nanoparticles-decorated graphene oxide can be synthesized by hydrothermal treatment of GO prepared by Hummers’s method in presence of tin chloride. The active chemical groups covering the prepared GO upon using the chemical route lead to a homogeneous distribution for the metallic nanoparticles. Moreover, these groups procure to a good attachment of the inorganic oxide with the carbonaceous support. The prepared SnO2/GO catalyst improves the electrocatalytic activity toward iodine redox reaction. Moreover, a decrease in the electron transfer was observed due to the harmony between the metallic nanoparticles in the proposed composite and the tin oxide existing in the utilized transparent glass electrodes (FTO). Accordingly, exploiting the prepared catalyst as a counter electrode in the DSSC showed a relatively good energy conversion with an efficiency of 4.57%. Overall, it can be claimed that the SnO2-decorated GO might be a new effective counter electrode for DSSCs.

Author Contributions

M.S.M. helped in writing the original draft of the manuscript. M.M. helped in the formal analysis. N.A.M.B. helped in supervision and project administration.

Acknowledgments

The authors deeply thank M. Shaheer Akhtar (Chonbuk National University) for measuring the J–V data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mathew, S.; Yella, A.; Gao, P.; Humphry-Baker, R.; Curchod, B.F.; Ashari-Astani, N.; Tavernelli, I.; Rothlisberger, U.; Nazeeruddin, M.K.; Grätzel, M. Dye-sensitized solar cells with 13% efficiency achieved through the molecular engineering of porphyrin sensitizers. Nat. Chem. 2014, 6, 242. [Google Scholar] [CrossRef] [PubMed]
  2. Ning, Z.; Fu, Y.; Tian, H. Improvement of dye-sensitized solar cells: What we know and what we need to know. Energy Environ. Sci. 2010, 3, 1170–1181. [Google Scholar] [CrossRef]
  3. Lee, C.-P.; Li, C.-T.; Ho, K.-C. Use of organic materials in dye-sensitized solar cells. Mater. Today 2017, 20, 267–283. [Google Scholar] [CrossRef] [Green Version]
  4. Gong, J.; Sumathy, K.; Qiao, Q.; Zhou, Z. Review on dye-sensitized solar cells (DSSCs): Advanced techniques and research trends. Renew. Sustain. Energy Rev. 2017, 68, 234–246. [Google Scholar] [CrossRef]
  5. Sharma, S.; Siwach, B.; Ghoshal, S.; Mohan, D. Dye sensitized solar cells: From genesis to recent drifts. Renew. Sustain. Energy. Rev. 2017, 70, 529–537. [Google Scholar] [CrossRef]
  6. Ye, M.; Wen, X.; Wang, M.; Iocozzia, J.; Zhang, N.; Lin, C.; Lin, Z. Recent advances in dye-sensitized solar cells: From photoanodes, sensitizers and electrolytes to counter electrodes. Mater. Today 2015, 18, 155–162. [Google Scholar] [CrossRef]
  7. Wu, J.; Lan, Z.; Lin, J.; Huang, M.; Huang, Y.; Fan, L.; Luo, G. Electrolytes in dye-sensitized solar cells. Chem. Rev. 2015, 115, 2136–2173. [Google Scholar] [CrossRef]
  8. Mahmoud, M.S.; Akhtar, M.S.; Mohamed, I.M.; Hamdan, R.; Dakka, Y.A.; Barakat, N.A. Demonstrated photons to electron activity of S-doped TiO2 nanofibers as photoanode in the DSSC. Mater. Lett. 2018, 225, 77–81. [Google Scholar] [CrossRef]
  9. Joshi, P.; Zhang, L.; Davoux, D.; Zhu, Z.; Galipeau, D.; Fong, H.; Qiao, Q. Composite of TiO2 nanofibers and nanoparticles for dye-sensitized solar cells with significantly improved efficiency. Energy Environ. Sci. 2010, 3, 1507–1510. [Google Scholar] [CrossRef]
  10. Manthina, V.; Baena, J.P.C.; Liu, G.; Agrios, A.G. ZnO–TiO2 Nanocomposite Films for High Light Harvesting Efficiency and Fast Electron Transport in Dye-Sensitized Solar Cells. J. Phys. Chem. C 2012, 116, 23864–23870. [Google Scholar] [CrossRef]
  11. Yang, M.; Dong, B.; Yang, X.; Xiang, W.; Ye, Z.; Wang, E.; Wan, L.; Zhao, L.; Wang, S. TiO2 nanoparticle/nanofiber—ZnO photoanode for the enhancement of the efficiency of dye-sensitized solar cells. RSC Adv. 2017, 7, 41738–41744. [Google Scholar] [CrossRef]
  12. Basu, K.; Benetti, D.; Zhao, H.; Jin, L.; Vetrone, F.; Vomiero, A.; Rosei, F. Enhanced photovoltaic properties in dye sensitized solar cells by surface treatment of SnO2 photoanodes. Sci. Rep. 2016, 6, 23312. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Y.; Qi, T.; Wang, Q.; Zhang, Y.; Wang, D.; Zheng, W. Preparation of SnO2/rGO Photoanode and Its Effect on the Property of Dye-Sensitized Solar Cells. IEEE J. Photovolt. 2017, 7, 399–403. [Google Scholar] [CrossRef]
  14. Bykkam, S.; Kalagadda, B.; Kalagadda, V.R.; Ahmadipour, M.; Chakra, C.S.; Rajendar, V. Effect of Few-Layered Graphene-Based CdO Nanocomposite-Enhanced Power Conversion Efficiency of Dye-Sensitized Solar Cell. J. Electron. Mater. 2018, 47, 620–626. [Google Scholar] [CrossRef]
  15. Thomas, S.; Deepak, T.; Anjusree, G.; Arun, T.; Nair, S.V.; Nair, A.S. A review on counter electrode materials in dye-sensitized solar cells. J. Mater. Chem. A 2014, 2, 4474–4490. [Google Scholar] [CrossRef]
  16. Wu, J.; Li, Y.; Tang, Q.; Yue, G.; Lin, J.; Huang, M.; Meng, L. Bifacial dye-sensitized solar cells: A strategy to enhance overall efficiency based on transparent polyaniline electrode. Sci. Rep. 2014, 4, 4028. [Google Scholar] [CrossRef] [PubMed]
  17. Wu, J.; Lan, Z.; Lin, J.; Huang, M.; Huang, Y.; Fan, L.; Luo, G.; Lin, Y.; Xie, Y.; Wei, Y. Counter electrodes in dye-sensitized solar cells. Chem. Soc. Rev. 2017, 46, 5975–6023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Prasetio, A.; Subagio, A.; Purwanto, A.; Widiyandari, H. Dye-sensitized solar cell based carbon nanotube as counter electrode. AIP Conf. Proc. 2016, 1710, 030054. [Google Scholar]
  19. Hung, K.-H.; Li, Y.-S.; Wang, H.-W. Dye-sensitized solar cells using graphene-based counter electrode. In Proceedings of the 12th IEEE Conference on Nanotechnology (IEEE-NANO), Birmingham, UK, 20–23 August 2012; pp. 1–12. [Google Scholar]
  20. Motlak, M.; Barakat, N.A.; Akhtar, M.S.; Hamza, A.; Kim, B.-S.; Kim, C.S.; Khalil, K.A.; Almajid, A.A. High performance of NiCo nanoparticles-doped carbon nanofibers as counter electrode for dye-sensitized solar cells. Electrochim. Acta 2015, 160, 1–6. [Google Scholar] [CrossRef]
  21. Lee, W.J.; Ramasamy, E.; Lee, D.Y.; Song, J.S. Efficient dye-sensitized solar cells with catalytic multiwall carbon nanotube counter electrodes. ACS Appl. Mater. Int. 2009, 1, 1145–1149. [Google Scholar] [CrossRef]
  22. Wang, X.; Zhi, L.; Müllen, K. Transparent, conductive graphene electrodes for dye-sensitized solar cells. Nano Lett. 2008, 8, 323–327. [Google Scholar] [CrossRef] [PubMed]
  23. Marcaccio, M.; Paolucci, F. Making and Exploiting Fullerenes, Graphene, and Carbon Nanotubes; Springer: Berlin, Germany, 2014. [Google Scholar]
  24. Gao, W. The chemistry of graphene oxide. In Graphene Oxide; Springer: Berlin, Germany, 2015; pp. 61–95. [Google Scholar]
  25. Rahal, A.; Benramache, S.; Benhaoua, B. Preparation of n-type semiconductor SnO2 thin films. J. Semicond. 2013, 34, 083002. [Google Scholar] [CrossRef]
  26. El-Deen, A.G.; Barakat, N.A.; Khalil, K.A.; Kim, H.Y. Development of multi-channel carbon nanofibers as effective electrosorptive electrodes for a capacitive deionization process. J. Mater. Chem. A 2013, 1, 11001–11010. [Google Scholar] [CrossRef]
  27. Krishna, R.; Titus, E.; Okhay, O.; Gil, J.C.; Ventura, J.; Ramana, E.V.; Gracio, J.J. Rapid electrochemical synthesis of hydrogenated graphene oxide using Ni nanoparticles. Int. J. Electrochem. Sci. 2014, 9, 69. [Google Scholar]
  28. Wei, A.; Wang, J.; Long, Q.; Liu, X.; Li, X.; Dong, X.; Huang, W. Synthesis of high-performance graphene nanosheets by thermal reduction of graphene oxide. Mater. Res. Bull. 2011, 46, 2131–2134. [Google Scholar] [CrossRef]
  29. Katiyar, R.; Dawson, P.; Hargreave, M.; Wilkinson, G. Dynamics of the rutile structure. III. Lattice dynamics, infrared and Raman spectra of SnO2. J. Phys. C Solid Stat. Phys. 1971, 4, 2421. [Google Scholar] [CrossRef]
  30. Li, Z.; Shen, W.; Zhang, X.; Fang, L.; Zu, X. Controllable growth of SnO2 nanoparticles by citric acid assisted hydrothermal process. Colloid Surf. Physicochem. Eng. Asp. 2008, 327, 17–20. [Google Scholar] [CrossRef]
  31. Sarker, S.; Ahammad, A.J.S.; Seo, H.W.; Kim, D.M. Electrochemical Impedance Spectra of Dye-Sensitized Solar Cells: Fundamentals and Spreadsheet Calculation. Int. J. Photoenergy 2014, 2014, 17. [Google Scholar] [CrossRef]
  32. Murakami, T.N.; Grätzel, M. Counter electrodes for DSC: Application of functional materials as catalysts. Inorg. Chim. Acta 2008, 361, 572–580. [Google Scholar] [CrossRef]
  33. Yu, Y.-H.; Teng, I.J.; Hsu, Y.-C.; Huang, W.-C.; Shih, C.-J.; Tsai, C.-H. Covalent bond–grafted soluble poly(o-methoxyaniline)-graphene oxide composite materials fabricated as counter electrodes of dye-sensitised solar cells. Org. Electron. 2017, 42, 209–220. [Google Scholar] [CrossRef]
  34. Huo, J.; Wu, J.; Zheng, M.; Tu, Y.; Lan, Z. High performance sponge-like cobalt sulfide/reduced graphene oxide hybrid counter electrode for dye-sensitized solar cells. J. Power Sources 2015, 293, 570–576. [Google Scholar] [CrossRef]
  35. Demir, E.; Savk, A.; Sen, B.; Sen, F. A novel monodisperse metal nanoparticles anchored graphene oxide as Counter Electrode for Dye-Sensitized Solar Cells. Nano-Struct Nano-Obj. 2017, 12, 41–45. [Google Scholar] [CrossRef]
  36. Jang, H.-S.; Yun, J.-M.; Kim, D.-Y.; Na, S.-I.; Kim, S.-S. Transparent graphene oxide–Pt composite counter electrode fabricated by pulse current electrodeposition-for dye-sensitized solar cells. Surf. Coat. Technol. 2014, 242, 8–13. [Google Scholar] [CrossRef]
  37. Satoshi, K.; Daiki, A.; Etsuo, S.; Masahiro, M. Copper Sulfide Catalyzed Porous Fluorine-Doped Tin Oxide Counter Electrode for Quantum Dot-Sensitized Solar Cells with High Fill Factor. Int. J. Photoenergy 2017, 2017, 5461030. [Google Scholar]
  38. Zaaba, N.; Foo, K.; Hashim, U.; Tan, S.; Liu, W.-W.; Voon, C. Synthesis of graphene oxide using modified hummers method: Solvent influence. Procedia Eng. 2017, 184, 469–477. [Google Scholar] [CrossRef]
  39. Rao, S.S.; Gopi, C.V.; Kim, S.-K.; Son, M.-K.; Jeong, M.-S.; Savariraj, A.D.; Prabakar, K.; Kim, H.-J. Cobalt sulfide thin film as an efficient counter electrode for dye-sensitized solar cells. Electrochim. Acta 2014, 133, 174–179. [Google Scholar]
Figure 1. Transmission electron microscope (TEM) (a); and High resolution TEM (HR-TEM) (b) images of SnO2-incorporated GO.
Figure 1. Transmission electron microscope (TEM) (a); and High resolution TEM (HR-TEM) (b) images of SnO2-incorporated GO.
Catalysts 09 00139 g001
Figure 2. XRD patterns for the pristine and SnO2-incorporated graphene oxide.
Figure 2. XRD patterns for the pristine and SnO2-incorporated graphene oxide.
Catalysts 09 00139 g002
Figure 3. Raman spectroscopy analyses for the pristine and SnO2-decorated graphene oxide. The inset displays high magnification of the D and G bands.
Figure 3. Raman spectroscopy analyses for the pristine and SnO2-decorated graphene oxide. The inset displays high magnification of the D and G bands.
Catalysts 09 00139 g003
Figure 4. (a) Cyclic voltammetry analyses of the introduced SnO2/GO in presence of iodine solution with a scan rate of 25 mV/s in the potential range from −1 to 1 V vs. Ag/AgCl; and (b) the cyclic voltammograms at different scan rates.
Figure 4. (a) Cyclic voltammetry analyses of the introduced SnO2/GO in presence of iodine solution with a scan rate of 25 mV/s in the potential range from −1 to 1 V vs. Ag/AgCl; and (b) the cyclic voltammograms at different scan rates.
Catalysts 09 00139 g004
Figure 5. Electrochemical impedance of the DSSC fabricated with GO and SnO2-incorporated GO-based counter electrode: (A) Nyquist plot; and (B) Bode diagram.
Figure 5. Electrochemical impedance of the DSSC fabricated with GO and SnO2-incorporated GO-based counter electrode: (A) Nyquist plot; and (B) Bode diagram.
Catalysts 09 00139 g005
Figure 6. Photovoltaic characteristics (J–V) of fabricated DSSCS using SnO2-incorporated GO based counter electrode measured under the irradiance of AM 1.5 G sunlight of 100 mW cm−2.
Figure 6. Photovoltaic characteristics (J–V) of fabricated DSSCS using SnO2-incorporated GO based counter electrode measured under the irradiance of AM 1.5 G sunlight of 100 mW cm−2.
Catalysts 09 00139 g006
Table 1. The detailed photovoltaic parameters from J–V curves and a brief comparison with other works.
Table 1. The detailed photovoltaic parameters from J–V curves and a brief comparison with other works.
Type of CEη, %FFVOC, VJsc, mA/cm2Ref.
Pt layer5.90.510.71816.12[20]
CoS/rGO9.390.630.76419.42[34]
Mw-Pt NPs@GO7.960.710.716.02[35]
GO/FTO3.990.380.7114.02[36]
GO-ED Pt/FTO7.050.660.7114.98[37]
POMA-FGO (1%)7.260.610.7316.31[38]
CuS5/FTO5 μm1.120.520.375.81[39]
SnO2/GO4.570.490.6714.04This study

Share and Cite

MDPI and ACS Style

Mahmoud, M.S.; Motlak, M.; Barakat, N.A.M. Facile Synthesis and Characterization of Two Dimensional SnO2-Decorated Graphene Oxide as an Effective Counter Electrode in the DSSC. Catalysts 2019, 9, 139. https://doi.org/10.3390/catal9020139

AMA Style

Mahmoud MS, Motlak M, Barakat NAM. Facile Synthesis and Characterization of Two Dimensional SnO2-Decorated Graphene Oxide as an Effective Counter Electrode in the DSSC. Catalysts. 2019; 9(2):139. https://doi.org/10.3390/catal9020139

Chicago/Turabian Style

Mahmoud, Mohamed S., Moaaed Motlak, and Nasser A. M. Barakat. 2019. "Facile Synthesis and Characterization of Two Dimensional SnO2-Decorated Graphene Oxide as an Effective Counter Electrode in the DSSC" Catalysts 9, no. 2: 139. https://doi.org/10.3390/catal9020139

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop