Mingyu
Li
a,
Chang
Yu
*a,
Chao
Hu
b,
Changtai
Zhao
a,
Mengdi
Zhang
a,
Yiwang
Ding
a,
Xiuna
Wang
a and
Jieshan
Qiu
*a
aState Key Lab of Fine Chemicals, School of Chemical Engineering, Liaoning Key Lab for Energy Materials and Chemical Engineering, Dalian University of Technology, Dalian 116024, China. E-mail: chang.yu@dlut.edu.cn; jqiu@dlut.edu.cn; Tel: +86-411-84986080
bSchool of Chemical Engineering and Technology, Xi'an Jiaotong University, Xi'an, 710049, China
First published on 21st November 2017
Counter electrodes (CEs) play critical roles in the reduction and regeneration of triiodide/iodide redox couple in dye-sensitized solar cells (DSSCs). Compared to commercial Pt, cost-efficient CEs with excellent electrocatalytic activity and superior electrochemical stability are highly desired. Herein, we report a facile, template- and active agents-free fabrication strategy for the synthesis of carbon nanosheets (CNSs) via annealing of small molecular precursors. This process was achieved by a combined strategy, including a low-temperature solid-phase cross-linking reaction and a subsequent high temperature annealing. When employed as metal-free CEs for DSSCs, the as-obtained CNSs demonstrated an annealing temperature-dependent electrochemical behavior. Owing to the superior electrical conductivity and electrocatalytic activity, the CNSs obtained by annealing at 1200 °C exhibit the best electrochemical performance with a power conversion efficiency of 8.71%, which is superior to that of Pt CE (7.24%), thus being attractive alternatives to precious metal Pt CEs. This study presents a simple and effective strategy to configure nanostructured carbonaceous materials for high-performance energy storage and conversion.
Suitable alternatives such as carbonaceous materials, conductive polymers and transition metal compounds, have been used as CEs in DSSCs.11 Among these candidates, the conventional carbonaceous materials as metal-free CEs have shown a unique superiority owing to their high surface area, high electrical conductivity and low cost in terms of green chemistry and sustainability.12–14 The template-assisted synthetic strategy is an efficient method to fabricate such carbonaceous nanomaterials with a controlled morphology and nonaggregated structure. Moreover, KOH activation is a general and acceptable method to increase the defects in carbonaceous nanomaterials. However, the addition and removal of the template and the used alkali are time-consuming, environmentally unfriendly and expensive.15,16 As such, it is challenging and imperative to explore a novel, template- and active agents-free approach for the synthesis of highly electrochemically active carbonaceous material as CEs in DSSCs.
Herein, for the first time, we report a facile, green and template-free strategy for the synthesis of interconnected carbon nanosheets (CNSs) via low-temperature cross-linking coupled with high temperature annealing of urea and sodium citrate. In particular, the abundant defects and active sites are also generated within the carbon matrix by in situ activation of the gases produced by the intermediate decomposition. When applied as CEs for DSSCs, the as-prepared porous CNSs exhibit an annealing temperature-dependent electrochemical behavior. The sample obtained by annealing at 1200 °C delivers a high PCE of 8.71% as CEs, which is attributed to the superior electrical conductivity and electrocatalytic activity of the nanosheets, outperforming the Pt CE (7.24%).
Moreover, the weight loss is accompanied by the removal of the functional groups, which was confirmed by FTIR spectra (Fig. 1b). It can be noted that after cross-linking, the peak at ca. 3340 cm−1, corresponding to the stretching vibration of N–H, partially disappears compared with the raw mixture. The disappearance of the peak is related to the partial decomposition of urea, which could further promote the interaction of urea and the sodium citrate in the cross-linking reaction. Moreover, the two new peaks at ca. 2300 cm−1 and 2100 cm−1 are assigned to the vibration of CN and CC, respectively. These combined results present the interaction between urea and sodium citrate derived from the cross-linking. After annealing, the peaks assigned to N–H, CN, CC and other functional groups below ca. 1500 cm−1 almost disappear. These results are in agreement with those of elemental analysis (Table S1†). Specifically, the carbon content increases from 83.75 wt% to 94.55 wt% when the annealing temperature varies from 1000 °C to 1400 °C. In addition, the XRD patterns shown in Fig. S1† further demonstrate the formation of inorganic intermediate products, such as Na2CO3 and NaHCO3 during the annealing process. When the annealing temperature reaches 1200 °C, characteristic peaks of the intermediate products disappear, indicating the complete decomposition of intermediates. The as-obtained CNSs also exhibit an annealing temperature-dependent crystallinity behavior with the enhancement of the graphite (002) peak (Fig. 1c).
The surface defect states of the CNSs were further characterized by Raman spectroscopy, of which the measured results are shown in Fig. 1d. After the annealing process, two peaks at ca. 1360 cm−1 (D band) and 1580 cm−1 (G band) are produced. When the annealing temperature reaches 1200 °C, a peak corresponding to 2D band further appears, indicative of the graphene-like structure. The ID/IG values of CNS-1000 and CNS-1400 are clearly lower than that of CNS-1200 (1.26), indicative of the increased defects in CNS-1200.
The typical FESEM images of the as-prepared CNS-1200 are shown in Fig. 2a–c, indicative of an interconnected and hierarchical microstructure configured by numerous large-area carbon nanosheets without aggregation. The unique porous sheet-shaped structure was further confirmed by the TEM image shown in Fig. 2d. The corresponding energy dispersive X-ray spectra (Fig. S2†) demonstrate the uniform distribution of C, N and O elements. The microstructure and morphology changes of the CNSs were also explored with annealing temperature ranging from 1000 °C to 1400 °C (Fig. 2 and Fig. S3†). With an increase of the annealing temperature, the smooth surface gradually becomes rough, which could be attributed to the partial decomposition of sodium citrate and the formation of inorganic intermediates. In particular, the slightly small-sized nanosheets are present due to the activation and etching function of the gases produced by the complete decomposition of intermediates at 1200 °C. When the annealing temperature reaches 1400 °C, the sheet structure is collapsed and destroyed to some degree due to the enhanced harsh conditions, indicating the morphological destruction derived from high temperature annealing. The morphology differences are also in accordance with the results of XRD patterns and Raman mentioned above.
Fig. 2 (a–c) FESEM images of the as-prepared CNS-1200 with different magnifications. (d) TEM image of the as-prepared CNS-1200. |
The photovoltaic performance of as-prepared samples was measured as CEs for DSSCs. As illustrated in Table 1 and Fig. 3a, the short-circuit photocurrent density (Jsc) and open-circuit voltage (Voc) of CNS-1200 are 15.94 mA cm−2 and 0.77 V, respectively, which are clearly superior to those of the Pt reference as well as the CNS-1000 and CNS-1400 samples. This indicates that the as-prepared CNS-1200 has the highest electrocatalytic performance towards the I3− reduction. As a result, the CNS-1200 exhibits the highest PCE of 8.71% than that of CNS-1000 (8.38%) and CNS-1400 (6.98%), which is comparable to that of other reported CEs in the literature (Table S2†). The superior electrochemical activity is also verified by CV curves (Fig. 3b and Fig. S4†). It is noted that a couple of oxidation and reduction peaks (Aox/Ared and Box/Bred) could be clearly observed in all the samples, corresponding to oxidation and reduction of I3−/I− and I2/I3−, respectively.19,20 The CNS-1200 has a smaller Epp value (0.22 V) and higher Ared peak current density (−3.37 mA cm−2) than those of commercial Pt (0.25 V, −1.34 mA cm−2), CNS-1000 (0.23 V, −2.99 mA cm−2) and CNS-1400 (0.23 V, −2.58 mA cm−2), suggesting that the CNS-1200 features more electrocatalytic active sites towards I3− reduction.
Samples | V oc (V) | J sc (mA cm−2) | FF (%) | PCE (%) |
---|---|---|---|---|
CNS-1000 | 0.76 ± 0.01 | 15.40 ± 0.26 | 71.27 ± 0.39 | 8.38 ± 0.02 |
CNS-1200 | 0.77 ± 0.01 | 15.94 ± 0.18 | 71.03 ± 0.19 | 8.71 ± 0.11 |
CNS-1400 | 0.72 ± 0.01 | 14.43 ± 0.16 | 67.74 ± 0.14 | 6.98 ± 0.06 |
Pt | 0.75 ± 0.01 | 15.51 ± 0.03 | 70.98 ± 5.71 | 7.24 ± 0.06 |
The EIS measurements were performed to reflect the electrochemical characteristics and the results are summarized in Fig. 3c and Table S3.† Three semicircles are present in the Nyquist plot of CNS-1200, which match with the Nernst diffusion resistance in pores (Zpore), charge-transfer resistance (Rct) at the electrode/electrolyte interface and Nernst diffusion impedance (ZN) derived from mass transport limitation in the electrolyte.14 Moreover, the high-frequency intercept on real axis determines the serial resistance (Rs). In contrast, Zpore is absent in commercial Pt.21,22 It should be noted that the Rct (0.35 Ω cm2) and Rs (3.48 Ω cm2) of CNS-1200 are both smaller than those of Pt CEs (Rct = 1.52 Ω cm2, Rs = 6.00 Ω cm2), revealing the superior electrocatalytic activity of CNS-1200.23–25 Furthermore, a large exchange current density (J0) and slope for the cathodic or anodic branch in polarization zone of the Tafel polarization curves are delivered for the as-prepared CNS-1200 (Fig. 3d). High J0 corresponds to low Rct, which is consistent with the EIS results mentioned above.26–29 In conclusion, the as-prepared CNS-1200 achieves a small Epp value, high Ared peak current density, low resistance value and better electrochemical activity, indicating that the CNS-1200 possesses great potential for Pt replacement.
The electrochemical stability of the CEs is another important issue for practical implementation of DSSCs, and it is primarily dependent on Rct in principle. In this case, freshly assembled identical dummy cells were subjected to repeated EIS measurements, of which the results are shown in Fig. 4. It could be observed in Fig. 4a and c that Rs, Rct and ZN values for CNS-1200 dummy cells undergo negligible changes after 10 cycles, indicative of a relatively high electrochemical stability. In contrast, the Rct of Pt CEs increases dramatically with the increase in cycling number, indicative of their low electrochemical stability (Fig. 4b and c).30 As a result, the CNS-1200 CEs with a prominently enhanced I3− reduction activity and the outstanding electrochemical stability show great potential for replacing noble metal Pt as a CE material for commercial application of DSSCs.
Fig. 4 Electrochemical stability of identical dummy cells with (a) CNS-1200 and (b) Pt CEs in acetonitrile solution of I3−/I−. (c) Rct changes versus the EIS scan numbers for CNS-1200 and Pt CEs. |
Footnote |
† Electronic supplementary information (ESI) available: FESEM images, EDS mapping, XRD patterns, elemental analysis and electrochemical performance. See DOI: 10.1039/c7gc02701j |
This journal is © The Royal Society of Chemistry 2018 |