Next Article in Journal
Vital Role of Synthesis Temperature in Co–Cu Layered Hydroxides and Their Fenton-like Activity for RhB Degradation
Next Article in Special Issue
Enhancing Photocatalysis of Ag Nanoparticles Decorated BaTiO3 Nanofibers through Plasmon-Induced Resonance Energy Transfer Turned by Piezoelectric Field
Previous Article in Journal
Decolorization and Oxidation of Acid Blue 80 in Homogeneous and Heterogeneous Phases by Selected AOP Processes
Previous Article in Special Issue
Highly Efficient and Selective Carbon-Doped BN Photocatalyst Derived from a Homogeneous Precursor Reconfiguration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

α-Fe2O3/Reduced Graphene Oxide Composites as Cost-Effective Counter Electrode for Dye-Sensitized Solar Cells

1
School of Materials, Shenzhen Campus of Sun Yat-sen University, No. 66, Gongchang Road, Guangming District, Shenzhen 518107, China
2
Songshan Lake Materials Laboratory, Room 425, C1 Building, University Innovation City, Songshan Lake, Dongguan 523000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(6), 645; https://doi.org/10.3390/catal12060645
Submission received: 16 May 2022 / Revised: 7 June 2022 / Accepted: 8 June 2022 / Published: 13 June 2022
(This article belongs to the Special Issue Advances in Heterojunction Photocatalysts)

Abstract

:
The counter electrode (CE) is an important and vital part of dye-sensitized solar cells (DSSCs). Pt CEs show high-performance in DSSCs using iodide-based electrolytes. However, the high cost of Pt CEs restricts their large-scale application in DSSCs and the development of Pt-free CE is expected. Here, α-Fe2O3/reduced graphene oxide (α-Fe2O3/RGO) composites are prepared as the Pt-free CE materials for DSSCs. A simple hydrothermal technique was used to disseminate the α-Fe2O3 solid nanoparticles uniformly throughout the RGO surface. The presence of the α-Fe2O3 nanoparticles increases the specific surface area of RGO and allows the composites to be porous, which improves the diffusion of liquid electrolyte into the CE material. Then, the electrocatalytic properties of CEs with α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt materials are compared. The α-Fe2O3/RGO CE has a similar electrocatalytic performance to Pt CE, which is superior to those of the pure α-Fe2O3 and RGO CEs. After being fabricated as DSSCs, the current–voltage measurements reveal that the DSSC based on α-Fe2O3/RGO CE has a power conversion efficiency (PCE) of 6.12%, which is 88% that of Pt CE and much higher than that of pure α-Fe2O3 and pure RGO CEs. All the results show that this work describes a promising material for cost-effective, Pt-free CEs for DSSCs.

Graphical Abstract

1. Introduction

Due to their low price, greater energy conversion efficiency, and easy manufacturing technique, dye-sensitized solar cells (DSSCs) have received much interest [1,2]. The choice of counter electrode (CE) material is different for different electrolytes [3]. CE electron transmission from the external circuit to iodine and triiodide (I/I3) in the redox electrolyte is crucial for developing DSSCs [4]. On the CE of DSSCs, thin films of Pt are often utilized as catalysts. However, large-scale production is not possible because Pt is a precious metal and costly. As a result, various attempts have been made to determine potential alternative materials for replacing Pt CE in DSSCs, such as transition metal oxides, which are cheaper, more conductive, and more chemically stable. Among the transitional metal oxides, iron oxides have the features of low cost, no toxicity, elemental abundance [5,6]. Fe2O3 is an important iron oxide found in various forms, such as α-Fe2O3, β-Fe2O3, γ-Fe2O3, and ε-Fe2O3 [7]. Each has its own unique characteristics. For example, α-Fe2O3 has very good electrochemical activity and is the most stable phase of iron oxide under ambient conditions [8,9], which makes it a promising candidate for cost-effective CE materials for DSSC [10]. Moreover, α-Fe2O3 is an n-type indirect semiconductor that is able to utilize about 40% of sunlight. Miao et al. reported that flower-shaped α-Fe2O3 could be used as the photoanode of DSSC and achieved 1.24% power conversion efficiency (PCE) [11]. However, the poor conductivity of the α-Fe2O3 restricts its further development and application [7].
Graphene is also a suitable material for the CE of DSSCs due to its long-term stability, with atoms organized in close-packed conjugated hexagonal lattices similar to graphite, but a one-atom-thick sheet [12,13]. Graphene has attracted considerable interest for energy convention owing to its superior electroconductibility, chemical stability, large specific area, and broad electrochemical window [14,15]. Graphene CEs possess a large surface area, defects, and oxygen-containing groups (such as reduced graphene oxide, RGO), suggesting the possibility of achieving comparable electrochemical performances to Pt CEs. [16] However, the poor electrocatalytic activity of RGO limits its application in DSSCs as a CE material.
Above all, the synergistic effect arising from interactions between α-Fe2O3 and RGO is essential because the presence of RGO can increase the electrical conductivity of α-Fe2O3 as well as α-Fe2O3 improving the electrocatalytic activity and reducing the cost of the CE materials. Furthermore, previous work found that the open spaces caused by α-Fe2O3 between graphene nanosheets may mitigate the effect of RGO volume change [17], which is beneficial to the electron transfer process in catalysis. Therefore, the composites based on α-Fe2O3 and RGO (α-Fe2O3/RGO) are expected to take advantage of the structural characteristics to yield improved CE performance in DSSCs. Chen et al. reported that a DSSC with 3D α-Fe2O3/GFs as the CE material displayed a superior PCE to that of Pt due to the positive synergistic effects of α-Fe2O3 and the 3D graphene frameworks (GFs) [18]. However, the synthesis process of the reported α-Fe2O3/GF materials not only requires a long time hydrothermal treatment, but also requires a high-temperature annealing process in Ar2, which limits its large-scale application in DSSCs. Zhao et al. demonstrated 3D α-Fe2O3 hollow meso-microspheres on graphene sheets by applying a solvothermal strategy [19]. The synthesis process involves two hydrothermal treatments at 150 °C for hours. This structure delivers decent electrocatalytic performance in a dye-sensitized solar cell that is comparable to that of Pt. Employing α-Fe2O3, which has distinct morphology, as well as optimizing the energy and cost can further advance its potential in high-performance catalysts. In previous research, graphene oxide (GO) and Fe(OH)3 were used to synthesize the α-Fe2O3/RGO composites by a facile and cost-effective process [20]. The Fe(OH)3 sol contributed to the homogeneous size and excellent dispersity of α-Fe2O3 solid nanoparticles on RGO. The concentration of the composite was 73% α-Fe2O3/RGO, with the best performance exhibited when the volume ratio between the GO solution and the Fe(OH)3 sol was 2:1.
Here, in this work, α-Fe2O3/RGO composites are developed to be used as a cost-effective CE in a DSSC system and the performances of DSSCs with α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt CEs are compared under the same conditions. A facile synthesis process is used to develop α-Fe2O3 solid nanoparticles on RGO with outstanding homogeneity and dispersion by employing Fe(OH)3 and a GO sol. As a CE material for DSSCs, the α-Fe2O3/RGO composites have a larger specific surface area than pure RGO, which induces the composites to have better electrocatalytic activity. The PCE of the DSSC using α-Fe2O3/RGO CE is increased by 30.5% compared to the pure α-Fe2O3 CE (4.69%), which itself is much higher than pure RGO (98.7% increase). This work presents a promising route for a cost-effective production way for CE materials for DSSCs.

2. Results and Discussion

By comparing the X-ray diffraction (XRD) patterns of α-Fe2O3/RGO composites, α-Fe2O3, RGO, and GO, it can seen that GO has been transformed to RGO (Figure 1a). Pure and composite α-Fe2O3 particles are all highly crystalline, which agrees with the reference (PDF#89-0597). The diffraction peaks at 2θ of 24.1°, 33.1°, 35.6°, 40.8°, 49.4°, 54.0°, 62.4°, and 64.0° correspond to (012), (104), (110), (113), (024), (116), (214), and (300). There are no significant variations in the XRD data of pure α-Fe2O3 and α-Fe2O3/RGO, except that the α-Fe2O3/RGO has a diffraction peak at 2θ = 25.6°, which corresponds to the graphene crystal faces (d-spacing of 3.35 Å) [21].
Figure 1b shows the Raman spectra of the α-Fe2O3/RGO, α-Fe2O3, RGO, and GO samples. The D and G bands reveal typical peaks at 1350 cm−1 and 1589 cm−1 in all three graphene-related compounds. α-Fe2O3/RGO and RGO have higher ID/IG values than GO, indicating that GO has been reduced to RGO [22]. GO has an ID/IG ratio of 0.95, while α-Fe2O3/RGO has the most significant ratio at 1.57. Two peaks are presented for α-Fe2O3/RGO and pure α-Fe2O3 at 221 cm−1 and 285 cm−1, which correspond to hematite’s standard A1g and Eg Raman modes, respectively, while the peak for pure α-Fe2O3 at 1299 cm−1 is due to two magnetic oscillator scattering of hematite [23]. From the XRD and Raman result, the RGO sheets are probably attached by α-Fe2O3 nanoparticles.
FE-SEM was utilized to examine the morphologies of the α-Fe2O3/RGO, α-Fe2O3, and RGO materials fabricated as films that were used as CEs in DSSCs. Figure 2 presents a top view of films of different materials on FTO at the same magnification. As shown in Figure 2a, the pure α-Fe2O3 nanoparticles exhibit uniformly regular solid shape with a size of about 20–50 nm. The pure RGO has a typical wrinkled and folded structure as observed in Figure 2b. Figure 2c illustrates the SEM image of the α-Fe2O3/RGO composites film. The graphene sheets are evenly coated with uniform α-Fe2O3 particles on both sides and the α-Fe2O3/RGO film has a rough surface and many more pores than pure RGO. This variation in shape indicates that the strong force between GO sheets and Fe3+ has a significant impact on the crystalline growth of α-Fe2O3 nanoparticles [24].
TEM studies of the α-Fe2O3/RGO composites were performed to define their microstructure further (Figure 3). Figure 3a shows that α-Fe2O3 solid nanoparticles ranging in size from 20 to 50 nm are evenly dispersed over RGO. This result also reveals efficient assembly of the α-Fe2O3 particles and graphene sheets during the hydrothermal treatment. The 0.22 nm lattice spacing in the (113) plane identifies the α-Fe2O3 particles (Figure 3b) [25], matching with the XRD data. The three images (Figure 3d–f) of the targeted region (Figure 3c) illustrate the TEM elemental mapping findings of the α-Fe2O3/RGO composites. Figure 3d demonstrates that carbon (C) atoms are numerous in the composites, but iron (Fe) and oxygen (O) atoms are scarce in Figure 3e,f. However, all three images demonstrate the distribution of C, Fe, and O elements in the α-Fe2O3/RGO composites are highly homogeneous.
Figure 4a depicts the N2 adsorption–desorption isotherms of the α-Fe2O3/RGO composites. The sample showed a curve pattern between Type IV (BDDT classification) that displays hysteresis loops predominantly of type H3 in the adsorption isotherms in Figure 4a [26]. The BET specific surface area of the α-Fe2O3/RGO samples are measured to be 136.16 m2 g−1. According to the previous study [4], the addition of α-Fe2O3 provides a larger specific surface area for the composites (the specific area of RGO prepared by the same synthesis method is 32.2 m2 g−1). In addition, obvious pressure hysteresis can be observed in the N2 adsorption–desorption isotherms, revealing the porous nature of the α-Fe2O3/RGO nanostructure. This can be further identified by the BJH pore size distribution analysis which is shown in Figure 4b. The total pore volume of α-Fe2O3/RGO composites are 0.244 cm3 g−1 according to the BJH method. The α-Fe2O3/RGO composites have a narrow pore size distribution centered at 3.7 nm, while the RGO sample was suggested in the before work to be a structure without large number of holes. The incorporation of α-Fe2O3 particles enhanced the specific surface area of α-Fe2O3/RGO composites with a porous architecture. The abovementioned results are also consistent with the result in Figure 2. Liquid electrolyte may more readily permeate into this porous nanostructure, which significantly improves electrocatalytic performance.
To evaluate the electrocatalytic activities of as-prepared α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt, three-electrode cyclic voltammetry (CV) was performed (Figure 5). As previously reported, a representative curve with two couples of redox peaks was found for Pt. The I and I’ peaks correspond to the oxidation and reduction peaks of I3/I2, whereas the II and II’ peaks correspond to those of I/I3 [27]. During the electrochemical process in a DSSC, it is crucial that electrons from CE reduce I3 to I. Therefore, the II and II’ peaks represent the electro-catalytic capabilities of the CEs. By analyzing the reduction peaks of all the CEs, we can see that the α-Fe2O3/RGO and Pt have sharper II’ peaks, demonstrating that their catalytic activities are considerably higher than others. It is well known that peak-to-peak separation and peak current density are two essential metrics for assessing the electrocatalytic activities of CE [28]. The magnitude of peak current density is correlated to the capacity of the CE to decrease the I3 species. Compared with pure RGO and α-Fe2O3 CEs, the α-Fe2O3/RGO CE exhibits a greater peak current density (Figure 5), indicating that it has stronger electrocatalytic activity and a faster response rate.
To further examine the liquid electrolyte diffusion rate into the material and interfacial charge transfer properties of the triiodide/iodide pair on the electrode surface, Tafel polarization experiments were performed in a virtual device made with two duplicate electrodes. Figure 6 depicts the logarithmic current density (log J0) versus voltage (U) during the oxidation/reduction of triiodide to iodide. While the I3 is converted to I in the electrochemical cell, the impedance to charge transfer is inversely related to the exchange current density (J0). Using the Equation (1), this may be computed from the intersection of the tangent line of the polarization curve and the prolongation of the linear section to zero bias.
J0 = RT/nFRct
where R and F are constant, T is the room temperature, n is the amount of electrons participating in the reaction. The limiting diffusion exchange current density (Jlim) can also be computed from the Tafel curve using the Equation (2).
D = l Jlim/2nFC
where D is the diffusion coefficient of the triiodide, l is the thickness of the spacer, C is the triiodide concentration and n and F retain their defined meanings. Jlim depends on the diffusion rate of the I/I3 redox couple. An optimal auxiliary counter electrode should have high J0, Jlim, and lower Rct values.
Theoretically, the curve with a relatively low potential but higher than 0.1 V corresponds to the Tafel zone, in which the voltage is a linear function of the log of the current density (log J0) (Equation (1)). It is possible to determine the exchange current density (J0) in this region using Equation (1). In addition, the steeper the Tafel zone of the curve, the bigger the J0 and the greater the material’s catalytic activity. Pt has the greatest J0 in the Tafel zone, followed by the α-Fe2O3/RGO, α-Fe2O3, and RGO. This reveals the α-Fe2O3/RGO composite CE has more electrocatalytic activity when compared to the pure α-Fe2O3 and RGO CEs. The Tafel curves of the α-Fe2O3/RGO, α-Fe2O3, and RGO Ces are asymmetric may due to the different diffusion time of the electrolyte on two electrodes of symmetric dummy cells. To test the Tafel curve, the electrolyte was firstly dripped on one electrode (always the cathode), and then covered with another electrode on the electrolyte. This process causes the electrolyte to diffuse first on the cathode, resulting in asymmetry between the two electrodes in the Tafel curve. For this reason, the Pt electrode is also slightly asymmetrical. However, the asymmetric degree of the Pt CE is lighter because of the better catalytic ability.
To investigate the effects of the α-Fe2O3/RGO composites as a CE material for DSSC, the DSSCs were assembled using α-Fe2O3/RGO as the CE. For comparison, the properties of DSSCs fabricated with α-Fe2O3, RGO, and Pt Ces were also investigated. Five parallel devices for each sample were tested. The photovoltaic properties results of the four different DSSCs are summarized in Table 1. The short-circuit photocurrent (jsc), open-circuit voltage (Voc), fill factor (FF), and the power conversion efficiency (PCE, η) calculated photovoltaic parameters are provided. Due to its weak electrocatalytic activity, the DSSC with RGO CE has a poor transfer efficiency, of 3.08%, whereas the DSSC with α-Fe2O3/RGO CE has an open-circuit voltage of 645 mV, a short-circuit current of 15.43 mA cm−2, a fill factor of 0.61, and a cell efficiency of 6.12%, which is 88% of that of the DSSC with Pt CE (6.93%). In comparison to the α-Fe2O3 and RGO Ces, the α-Fe2O3/RGO CE has a substantially higher FF value. The photocurrent–voltage (I–V) curves of the DSSCs are shown in Figure 7a.
Figure 7b illustrates the resulting Nyquist plots from the EIS measurements performed on the DSSCs using the CEs above to obtain insight into the variation in FF values. The spectra were simulated using the matching circuit shown in the inset of Figure 7b, which contains high-frequency series resistance (Rs). The interfacial charge transfer resistance Rct and the capacitance of the electrical double layer (CPE1) are related to the RC processes of the CE/electrolyte interface in the intermediate frequency area. In addition, the interfacial charge transfer resistance (RR) and capacitance of the depletion layer of the TiO2 electrode (CPE2) are related to the RC processes of the TiO2 electrode/electrolyte interface in the low frequency range. The electrocatalytic property of the CE for triiodide reduction may be assessed, which is defined from the first hemicycle of the EIS spectrum [29]. According to Table 1, the α-Fe2O3 CE exhibits a greater Rct, suggesting weak electrocatalytic performance (as also proven by CV outcomes). When the RGO and the α-Fe2O3 are combined, the Rct decreases from 5.42 Ω to 3.81 Ω. The Rct value of the α-Fe2O3/RGO composites are much smaller than that of the RGO CE (11.02 Ω). This may be attributable to the integration of α-Fe2O3’s intrinsic high electrocatalytic activity on the RGO’s highly active electric transport channel [30]. Nonetheless, the α-Fe2O3/RGO CE has a substantially bigger sum of Rs and Rct than the Pt CE, leading to a lower FF and η values for the photovoltaic properties of the DSSCs [31].

3. Materials and Methods

3.1. Synthesis of Materials

First, a solution was prepared by adding 1.3 g FeCl3 (Aldrich, 98%, Shanghai, China) in 4 mL distilled water. Second, 50 mL distilled water was boiled. Third, the as-prepared solution (0.5 mL) was added into the boiling distilled water dropwise. Then, the mixture was kept boiling for several minutes to prepare the Fe(OH)3 sol. GO was manufactured from modified graphite oxide (GO), produced from natural scale graphite using a revised Hummers method technique [32,33]. In a typical synthesis, Fe(OH)3 sol was added dropwise to the GO solution (1 mg mL−1) at a volume ratio of 1:2 and then stirred for 30 min. Next, the mixture was heated to 85 °C in a water bath and hydrazine hydrate (Aldrich, 85%, Shanghai, China) was added into the mixture. Then, an ultrasonication process was applied for 30 min. After that, the solution was transferred into a Teflon-lined autoclave and heated at 150 °C for 6 h. The comparison mixture was prepared by centrifugation; α-Fe2O3 and RGO were prepared using the same technique, but without GO and FeCl3, respectively.

3.2. Fabrication of CEs and DSSCs

The CEs were fabricated by a doctor blade technique [34]. In one batch, films of a certain thickness were made. Comparatively, platinized CEs were made by a thermal breakdown. The chloro-platinic acid hexahydrate (H2PtCl6·6H2O, Pt ≥ 37.5%, AR, Aladdin, Shanghai, China) was mixed with isopropanol (C3H8O, ≥99.5%, AR, Aladdin, Shanghai, China) and the mixture was dropped on the cleaned fluorine-doped tin oxide (FTO) glass sheets (Nippon Sheet Glass Co., Osaka, Japan, surface resistance = 15 Ω/cm2, transmittance = 90%) and then heated for 15 min at 390 °C.
TiO2 was synthesized by using a sol–gel method [35]. Then, the TiO2 films (ca. 11 μm) were coated on the FTO glass using the doctor blade technique. The FTO glass coated with TiO2 was heated at 450 °C for 30 min. When the temperature dropped to around 90 °C, the electrodes were submerged for 24 h in a dye solution containing 0.5 mM cis-Ru (H2dcbpy)2 (NCS)2 (H2dcbpy = 4,4′-dicarboxy-2,2′-bipyridyl) (N3) dissolved in ethanol. After this, the TiO2 photoanodes are complete. The DSSCs were constructed using the aforementioned TiO2 photoanode, a CE, and a redox electrolyte comprising 0.5 M LiI (AR, Alfa, Zhengzhou, China), 0.05 M I2 (AR, Alfa), 0.6 M 4-tert-butylpyridine (TBP, >96%, Tokyo Chemical Industry Co., Ltd. Tokyo, Japan), and 0.6 M 3-hexyl-1-methylimidazolium iodide (HMII, AR, Alfa, Zhengzhou, China) in 3-methoxypropionitrile (MPN, 99%, GC, Alfa, Zhengzhou, China). The tested DSSCs were masked to a working area of 0.2 cm2.

3.3. Characterization

XRD (Cu Kαirradiation, D8-ADVANCE, Billerica, MA, USA) was used to analyze the crystalline structure and state of the samples. At room temperature, Raman spectra were acquired using a Raman spectroscope meter (Horiba LabRam HR Evolution, Tokyo, Japan) equipped with a 633 nm laser. Field emission scanning electron microscopy (FE-SEM, Gemini 300, Oberkochen, Germany) was used to examine the sample morphology (FE-SEM, Gemini 300, Oberkochen, Germany). The crystal structure was measured by high-resolution transmission electron microscopy (JEM-F200, Akishima, Japan). The specific sur face area and total pore volume of samples were determined by the Brunauer–Emmett–Teller (BET) method, in which the N2 adsorption at −195.8 °C was measured using an adsorption instrument (ASAP 2460, Norcross, GA, USA). The pore size distribution was estimated based on the desorption isotherm using the Barrett–Joyner–Halender (BJH) method [26].
Utilizing an electrochemical analyzer, cyclic voltammetry (CV) was performed in a three-electrode setup in an acetonitrile solution containing 0.1 M LiClO4, 10 mM LiI, and 1 mM I2 at a scan rate of 100 mV s−1 (Solartron SI 1287, Illinois, IL, USA). The Ag/Ag+ combination acted as the comparison electrode, while platinum served as the counter electrode. The spectra were fitted by the Zview software. Using an electrochemical workstation system (Solartron SI 1287, Illinois, IL, USA) in symmetric dummy cells built with two identical CEs and a scan speed of 50 mV S−1, Tafel polarization studies were performed. The photocurrent density–voltage (jsc-V) performance of the DSSCs was measured with a Keithley digital source meter (Keithley 2410, Cleveland, OH, USA) and simulated under AM 1.5 illumination (100 mW cm−2, Newport 69907). The incident light was calibrated with a power meter (model 350) and a detector (model 262). Electrochemical impedance spectroscopy (EIS) was performed on the Solartron SI 1260 frequency response analyzer and Solartron SI 1287 electrochemical interface system. The frequency range was 0.1–100 kHz and the AC voltage was 10 mV.

4. Conclusions

In this work, we demonstrated a novel Pt-free CE made of α-Fe2O3/RGO composites for DSSCs. The α-Fe2O3/RGO composites were proven to be synthesized successfully via a facile method as the α-Fe2O3 nanoparticles were found to be dispersed over the RGO surface. Compared with pure α-Fe2O3 and RGO, the α-Fe2O3/RGO composites have a larger specific surface area and a more porous microstructure, which provides an advantage to catalytic reactions. The α-Fe2O3/RGO CE was prepared on an FTO glass-substrate via the doctor blade process. Comprehensive electrochemical investigations revealed that the α-Fe2O3/RGO CE exhibited Pt-like electrocatalytic activity for I3 reduction owing to the synergistic effect between the inherently high catalytic performance of α-Fe2O3 and the easy charge transfer properties of RGO. The PCE of the DSSCs constructed with the α-Fe2O3/RGO CE was 6.12%, up to 88% of that achieved with the Pt CE. Therefore, the α-Fe2O3/RGO CE is a promising candidate for application as a cost-effective CE material in Pt-free transparent DSSCs. This work presents a solution that can easily prepare large-scale DSSCs with low cost, which is expected to improve the possibility of commercialization of DSSCs in the future.

Author Contributions

Data curation, X.L.; Formal analysis, S.L.; Investigation, S.L.; Methodology, L.S., Q.Z. and Q.L.; Project administration, J.S.; Resources, Q.Z. and Q.L.; Software, X.L.; Writing—original draft, L.S. and Q.Z.; Writing—review & editing, L.S., Q.Z., Q.L., W.L. and J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the open research fund of Songshan Lake Materials Laboratory (2021SLABFN21), the Hundreds of Talents Program of Sun Yat-sen University.

Acknowledgments

The authors would like to acknowledge financial support from the open research fund of Songshan Lake Materials Laboratory (2021SLABFN21), the Hundreds of Talents Program of Sun Yat-sen University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. O’regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  2. Saranya, K.; Rameez, M.; Subramania, A. Developments in conducting polymer based counter electrodes for dye-sensitized solar cells–An overview. Eur. Polym. J. 2015, 66, 207–227. [Google Scholar] [CrossRef]
  3. Kang, J.S.; Kim, J.; Kim, J.Y.; Lee, M.J.; Kang, J.; Son, Y.J.; Jeong, J.; Park, S.H.; Ko, M.J.; Sung, Y.E. Highly efficient bifacial dye-sensitized solar cells employing polymeric counter electrodes. ACS Appl. Mater. Interfaces 2018, 10, 8611–8620. [Google Scholar] [CrossRef]
  4. Zhang, Q.; Liu, Y.; Duan, Y.; Fu, N.; Liu, Q.; Fang, Y.; Sun, Q.; Lin, Y. Mn3O4/graphene composite as counter electrode in dye-sensitized solar cells. RSC Adv. 2014, 4, 15091–15097. [Google Scholar] [CrossRef]
  5. Singh, P.; Sharma, K.; Hasija, V.; Sharma, V.; Sharma, S.; Raizada, P.; Singh, M.; Saini, A.K.; Hosseini-Bandegharaei, A.; Thakur, V.K. Systematic review on applicability of magnetic iron oxides-integrated photocatalysts for degradation of organic pollutants in water. Mater. Today Chem. 2019, 14, 100186. [Google Scholar] [CrossRef]
  6. Ye, H.; Wang, Y.; Liu, X.J.; Xu, D.D.; Yuan, H.; Sun, H.Q.; Wang, S.B.; Ma, X. Magnetically steerable iron oxides-manganese dioxide core-shell micromotors for organic and microplastic removals. J. Colloid Interface Sci. 2021, 588, 510–521. [Google Scholar] [CrossRef]
  7. Kumar, Y.; Kumar, R.; Raizada, P.; Khan, A.A.P.; Singh, A.; Le, Q.V.; Nguyen, V.H.; Selvasembian, R.; Thakur, S.; Singh, P. Current status of hematite (α-Fe2O3) based Z-scheme photocatalytic systems for environmental and energy applications. J. Environ. Chem. Eng. 2022, 10, 107427. [Google Scholar] [CrossRef]
  8. Gao, R.J.; Wang, J.; Huang, Z.F.; Zhang, R.R.; Wang, W.; Pan, L.; Zhang, J.F.; Zhu, W.K.; Zhang, X.W.; Shi, C.X.; et al. Pt/Fe2O3 with Pt-Fe pair sites as a catalyst for oxygen reduction with ultralow Pt loading. Nat. Energy 2021, 6, 614–623. [Google Scholar] [CrossRef]
  9. Zhang, Q.; Liang, Q.; Liao, Q.; Ma, M.; Gao, F.; Zhao, X.; Song, Y.; Song, L.; Xun, X.; Zhang, Y. Amphiphobic hydraulic triboelectric nanogenerator for self-cleaning/charging power system. Adv. Funct. Mater. 2018, 28, 1803117. [Google Scholar] [CrossRef]
  10. Hou, Y.; Wang, D.; Yang, X.H.; Fang, W.Q.; Zhang, B.; Wang, H.F.; Lu, G.Z.; Hu, P.; Zhao, H.; Yang, H. Rational screening low-cost counter electrodes for dye-sensitized solar cells. Nat. Commun. 2013, 4, 1583. [Google Scholar] [CrossRef] [Green Version]
  11. Niu, H.; Zhang, S.; Ma, Q.; Qin, S.; Wan, L.; Xu, J.; Miao, S. Dye-sensitized solar cells based on flower-shaped α-Fe2O3 as a photoanode and reduced graphene oxide–polyaniline composite as a counter electrode. RSC Adv. 2013, 3, 17228–17235. [Google Scholar] [CrossRef]
  12. Wang, M.; Huang, M.; Luo, D.; Li, Y.; Choe, M.; Seong, W.K.; Kim, M.; Jin, S.; Wang, M.; Chatterjee, S.; et al. Single-crystal, large-area, fold-free monolayer graphene. Nature 2021, 596, 519–524. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, L.; Qing, M.; Wang, Y.; Chen, S. Defects in graphene: Generation, healing, and their effects on the properties of graphene: A review. J. Mater. Sci. Technol. 2015, 31, 599–606. [Google Scholar] [CrossRef]
  14. Fang, B.; Chang, D.; Xu, Z.; Gao, C. A review on graphene fibers: Expectations, advances, and prospects. Adv. Mater. 2020, 32, 1902664. [Google Scholar] [CrossRef]
  15. Zhang, Q.; Liang, Q.; Liao, Q.; Yi, F.; Zheng, X.; Ma, M.; Gao, F.; Zhang, Y. Service behavior of multifunctional triboelectric nanogenerators. Adv. Mater. 2017, 29, 1606703. [Google Scholar] [CrossRef]
  16. Kweon, D.H.; Baek, J.B. Edge-functionalized graphene nanoplatelets as metal-Free electrocatalysts for dye-sensitized solar cells. Adv. Mater. 2019, 31, 1804440. [Google Scholar] [CrossRef]
  17. Zhu, X.; Zhu, Y.; Murali, S.; Stoller, M.D.; Ruoff, R.S. Nanostructured reduced graphene oxide/Fe2O3 composite as a high-performance anode material for lithium ion batteries. ACS Nano 2011, 5, 3333–3338. [Google Scholar] [CrossRef]
  18. Yang, W.; Xu, X.W.; Li, Z.; Yang, F.; Zhang, L.Q.; Li, Y.F.; Wang, A.J.; Chen, S.L. Construction of efficient counter electrodes for dye-sensitized solar cells: Fe2O3 nanoparticles anchored onto graphene frameworks. Carbon 2016, 96, 947–954. [Google Scholar] [CrossRef]
  19. Zhao, G.M.; Xu, G.J.; Jin, S. α-Fe2O3 hollow meso-microspheres grown on graphene sheets function as a promising counter electrode in dye-sensitized solar cells. Rsc Adv. 2019, 9, 24164–24170. [Google Scholar] [CrossRef] [Green Version]
  20. Du, M.; Xu, C.; Sun, J.; Gao, L. Synthesis of α-Fe2O3 nanoparticles from Fe(OH)3 sol and their composite with reduced graphene oxide for lithium ion batteries. J. Mater. Chem. A 2013, 1, 7154–7158. [Google Scholar] [CrossRef]
  21. Wang, G. Preparation of α-Fe2O3/graphene composite and its electrochemical performance as an anode material for lithium ion batteries. J. Alloys Compd. 2011, 509, 216–220. [Google Scholar] [CrossRef]
  22. Pradhan, G.K.; Padhi, D.K.; Parida, K.M. Fabrication of α-Fe2O3 nanorod/RGO composite: A novel hybrid photocatalyst for phenol degradation. ACS Appl. Mater. Interfaces 2013, 5, 9101–9110. [Google Scholar] [CrossRef]
  23. De Faria, D.L.A.; Venâncio Silva, S.; De Oliveira, M.T. Raman microspectroscopy of some iron oxides and oxyhydroxides. J. Raman. Spectrosc. 1997, 28, 873–878. [Google Scholar] [CrossRef]
  24. Wang, H.; Xu, Z.; Yi, H.; Wei, H.; Guo, Z.; Wang, X. One-step preparation of single-crystalline Fe2O3 particles/graphene composite hydrogels as high performance anode materials for supercapacitors. Nano Energy 2014, 7, 86–96. [Google Scholar] [CrossRef]
  25. Sun, B.; Horvat, J.; Kim, H.S.; Kim, W.-S.; Ahn, J.; Wang, G. Synthesis of mesoporous α-Fe2O3 nanostructures for highly sensitive gas sensors and high capacity anode materials in lithium ion batteries. J. Phys. Chem. C 2010, 114, 18753–18761. [Google Scholar] [CrossRef]
  26. Xu, F.; Zhang, J.; Zhu, B.; Yu, J.; Xu, J. CuInS2 sensitized TiO2 hybrid nanofibers for improved photocatalytic CO2 reduction. Appl. Catal. B Environ. 2018, 230, 194–202. [Google Scholar] [CrossRef] [Green Version]
  27. Zhang, T.-L.; Chen, H.-Y.; Su, C.-Y.; Kuang, D.-B. A novel TCO-and Pt-free counter electrode for high efficiency dye-sensitized solar cells. J. Mater. Chem. A 2013, 1, 1724–1730. [Google Scholar] [CrossRef]
  28. Gong, F.; Wang, H.; Xu, X.; Zhou, G.; Wang, Z.-S. In situ growth of Co0.85Se and Ni0.85Se on conductive substrates as high-performance counter electrodes for dye-sensitized solar cells. J. Am. Chem. Soc. 2012, 134, 10953–10958. [Google Scholar] [CrossRef]
  29. Fu, N.Q.; Xiao, X.; Zhou, X.W.; Zhang, J.; Lin, Y. Electrodeposition of platinum on plastic substrates as counter electrodes for flexible dye-sensitized solar cells. J. Phys. Chem. C 2012, 116, 2850–2857. [Google Scholar] [CrossRef]
  30. Tai, S.-Y.; Liu, C.-J.; Chou, S.-W.; Chien, F.S.-S.; Lin, J.-Y.; Lin, T.-W. Few-layer MoS2 nanosheets coated onto multi-walled carbon nanotubes as a low-cost and highly electrocatalytic counter electrode for dye-sensitized solar cells. J. Mater. Chem. 2012, 22, 24753–24759. [Google Scholar] [CrossRef]
  31. Velten, J.; Mozer, A.J.; Li, D.; Officer, D.; Wallace, G.; Baughman, R.; Zakhidov, A. Carbon nanotube/graphene nanocomposite as efficient counter electrodes in dye-sensitized solar cells. Nanotechnology 2012, 23, 085201. [Google Scholar] [CrossRef] [Green Version]
  32. Alkhouzaam, A.; Abdelrazeq, H.; Khraisheh, M.; AlMomani, F.; Hameed, B.H.; Hassan, M.K.; Al-Ghouti, M.A.; Selvaraj, R. Spectral and structural properties of high-quality reduced graphene oxide produced via a simple approach using tetraethylenepentamine. Nanomaterials 2022, 12, 1240. [Google Scholar] [CrossRef]
  33. Hummers, W.S., Jr.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  34. Duan, Y.; Fu, N.; Zhang, Q.; Fang, Y.; Zhou, X.; Lin, Y. Influence of Sn source on the performance of dye-sensitized solar cells based on Sn-doped TiO2 photoanodes: A strategy for choosing an appropriate doping source. Electrochim. Acta 2013, 107, 473–480. [Google Scholar] [CrossRef]
  35. Duan, Y.; Fu, N.; Liu, Q.; Fang, Y.; Zhou, X.; Zhang, J.; Lin, Y. Sn-doped TiO2 photoanode for dye-sensitized solar cells. J. Phys. Chem. C 2012, 116, 8888–8893. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) Raman spectra of α-Fe2O3/RGO, α-Fe2O3, RGO, and GO.
Figure 1. (a) XRD patterns and (b) Raman spectra of α-Fe2O3/RGO, α-Fe2O3, RGO, and GO.
Catalysts 12 00645 g001
Figure 2. FE-SEM images of (a) α-Fe2O3, (b) RGO, (c) α-Fe2O3/RGO.
Figure 2. FE-SEM images of (a) α-Fe2O3, (b) RGO, (c) α-Fe2O3/RGO.
Catalysts 12 00645 g002
Figure 3. (a) TEM images of large-area and (b) high-resolution TEM image of α-Fe2O3/RGO composites; (cf) EDS elemental mapping images of α-Fe2O3/RGO composites.
Figure 3. (a) TEM images of large-area and (b) high-resolution TEM image of α-Fe2O3/RGO composites; (cf) EDS elemental mapping images of α-Fe2O3/RGO composites.
Catalysts 12 00645 g003
Figure 4. (a) N2 adsorption and desorption isotherms for the α-Fe2O3/RGO samples; (b) pore size distribution of α-Fe2O3/RGO.
Figure 4. (a) N2 adsorption and desorption isotherms for the α-Fe2O3/RGO samples; (b) pore size distribution of α-Fe2O3/RGO.
Catalysts 12 00645 g004
Figure 5. Cyclic voltammograms of α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt CEs at a scan rate of 100 mV s−1.
Figure 5. Cyclic voltammograms of α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt CEs at a scan rate of 100 mV s−1.
Catalysts 12 00645 g005
Figure 6. Tafel curves of the symmetrical cells fabricated with two identical α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt electrodes.
Figure 6. Tafel curves of the symmetrical cells fabricated with two identical α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt electrodes.
Catalysts 12 00645 g006
Figure 7. (a) Photocurrent density–voltage characteristics and (b) Nyquist plots of DSSCs with α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt CEs measured at AM 1.5 G illumination (100 mW cm−2). The inset in (b) is the equivalent circuit of the DSSCs.
Figure 7. (a) Photocurrent density–voltage characteristics and (b) Nyquist plots of DSSCs with α-Fe2O3/RGO, α-Fe2O3, RGO, and Pt CEs measured at AM 1.5 G illumination (100 mW cm−2). The inset in (b) is the equivalent circuit of the DSSCs.
Catalysts 12 00645 g007
Table 1. PCE performances and EIS parameters of the DSSCs with different CEs.
Table 1. PCE performances and EIS parameters of the DSSCs with different CEs.
CEsRctjscVocFFη
(Ω)(mA cm−2)(mV)(%)
Pt1.7816.336350.676.93
±0.06±0±0.00±0.05
α-Fe2O35.4213.36350.554.69
±0.05±5±0.01±0.03
RGO11.0212.475550.453.08
±0.05±5±0.01±0.05
α-Fe2O3/RGO3.8115.436450.616.12
±0.00±0±0.00±0.004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sun, L.; Zhang, Q.; Liang, Q.; Li, W.; Li, X.; Liu, S.; Shuai, J. α-Fe2O3/Reduced Graphene Oxide Composites as Cost-Effective Counter Electrode for Dye-Sensitized Solar Cells. Catalysts 2022, 12, 645. https://doi.org/10.3390/catal12060645

AMA Style

Sun L, Zhang Q, Liang Q, Li W, Li X, Liu S, Shuai J. α-Fe2O3/Reduced Graphene Oxide Composites as Cost-Effective Counter Electrode for Dye-Sensitized Solar Cells. Catalysts. 2022; 12(6):645. https://doi.org/10.3390/catal12060645

Chicago/Turabian Style

Sun, Lian, Qian Zhang, Qijie Liang, Wenbo Li, Xiangguo Li, Shenghua Liu, and Jing Shuai. 2022. "α-Fe2O3/Reduced Graphene Oxide Composites as Cost-Effective Counter Electrode for Dye-Sensitized Solar Cells" Catalysts 12, no. 6: 645. https://doi.org/10.3390/catal12060645

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