Next Article in Journal
Optimization of the Split-Spinach Aptamer for Monitoring Nanoparticle Assembly Involving Multiple Contiguous RNAs
Previous Article in Journal
A Facile Method for Batch Preparation of Electrochemically Reduced Graphene Oxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phosphate Ion-Modified RuO2/Ti3C2 Composite as a High-Performance Supercapacitor Material

1
Key Laboratory of Marine Environmental Corrosion and Bio-Fouling, Institute of Oceanology, Chinese Academy of Sciences, 7 Nanhai Road, Qingdao 266071, China
2
University of Chinese Academy of Science, Beijing 100049, China
3
Qingdao National Laboratory for Marine Science and Technology, Wenhai Road, Aoshanwei, Jimo, Qingdao 266237, China
4
School of Chemical Engineering and Technology, Sun Yat-sen University, Zhuhai 519082, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(3), 377; https://doi.org/10.3390/nano9030377
Submission received: 30 January 2019 / Revised: 27 February 2019 / Accepted: 27 February 2019 / Published: 5 March 2019

Abstract

:
Pseudocapitor materials, usually metal oxides, are used as active materials in an electrode to achieve high energy density. However, these kinds of materials often suffer from poor conductivity and high cost. Herein, a phosphate ion-modified RuO2/Ti3C2 composite is prepared via a chemical solution synthesis followed by an annealing process. In this composite material, Ti3C2 layers are introduced to improve the conductivity and the binary material is doped with phosphate ions into to increase the number of active reaction sites. As a result, the phosphate ion-modified RuO2/Ti3C2 delivers a high specific capacitance of 612.72 F g−1 at a current density of 2 A g−1 in H2SO4 electrolyte. What is more, the capacitance of the phosphate ion-modified RuO2/Ti3C2 can retain 97.95% (600.14 F g−1) of the original value even after 10,000 cycles at a current density of 2 A g−1.

1. Introduction

With the rapid development of electronic products, supercapacitors have attracted much effort in recent years owing to their merits of a quick charge–discharge process, high power density and long cycle lifetime [1,2,3,4,5]. Supercapacitors can be divided into two categories based on their energy storage mechanisms: electrical double-layer capacitors (EDLCs) based on the electrostatic charge accumulated at the electrode/electrolyte interface and pseudo-capacitors based on the reversible faradaic processes [4,6,7,8,9]. The performance of supercapacitors is largely dependent on the properties of electrode materials [10,11]. Porous carbon materials are the most widely used EDLC-type materials in recent commercial applications because they are cost-efficient and easy to synthesize [12,13,14,15]. However, the limited double-layer capacitance of carbon-based materials cannot meet the growing demands of power systems.
An effective option to increase the capacitance is to introduce a pseudo-capacitive material into the electrodes. Traditional pseudo-capacitance materials include metal oxide and conductive polymer materials [16,17,18,19]. Among the various metal oxide reported so far, ruthenium oxide (RuO2) has attracted great attention as an ideal candidate in view of its high specific capacitance, chemical stability and thermal stability [20,21,22,23,24,25,26,27]. However, the main issue of RuO2 against its application as a pseudo-capacitive material lies in its particle aggregation, which may lead to compromised electrochemical performance [25,28,29]. The combination of RuO2 with carbon materials has been demonstrated to be an effective strategy to address this issue. For example, Shen et al. prepared a RuO2 nanodots/reduced graphene oxide composite as a pseudocapacitive material which demonstrated improved cycling stability due to the layered structure of the ultrathin carbon sheets [30]. Zhu et al. synthesized a carbon quantum dot-decorated RuO2 network. The composites exhibited a specific capacitance of 460 F g−1 and excellent rate capability [25].
Ti3C2, as a novel kind of 2D material, was prepared by selectively etching the Al element from Ti3AlC2 in hydrofluoric acid. Ti3C2 showed a promising performance as an energy storage material due to its unique accordion-like layered structure, electrolyte wettability and good metallic conductivity [1,2,3,31,32,33]. For example, Kurra et al. fabricated a coplanar micro-supercapacitor by using Ti3C2 as the electrode material [34]. Peng et al. also fabricated an all Ti3C2-based solid-state micro-supercapacitor [35]. These devices exhibited excellent capacitances and great cyclic performances. Moreover, Ti3C2 is regarded as an excellent host for metal oxide to construct high-performance active materials [36]. For instance, Xiong et al. designed a sandwich-like SnO2/Ti3C2/SnO2 architecture through a wet chemistry approach, and the composite delivered high reversible capacity and long-term cycling (up to 810 mAh g−1 after 200 cycles) when used as a lithium battery electrode material [36]. Rakhi et al. deposited MnO2 over Ti3C2 nanosheets, and the fabricated composite exhibited excellent cycling stability in the application of a supercapacitor [37]. Therefore, combining the advantages of metal oxide and Ti3C2 to prepare Ti3C2/metal oxide composites is an effective method to construct high-performance energy storage materials.
Recently, Zhai and co-workers found that phosphate doping into Co3O4 could improve the fast electrode kinetics and stimulate high chemical reactivity [38]. Following a similar strategy, a phosphate ion-modified RuO2/ Ti3C2 composite (denoted as PRT) was fabricated in this work. As far as we know, phosphate ion-modified Ti3C2/RuO2 composites have not been reported as a supercapacitor electrode material. In general, the combination of phosphate ion, Ti3C2 and RuO2 can bring the following merits: (1) Ti3C2 serves as a support to prevent RuO2 particles from aggregating for longer cycling life; (2) the layered structure and high conductivity of Ti3C2 enables fast ion diffusion and fast electrons transfer, leading to a superior rate capability; (3) phosphate ion doping into RuO2 could highly improve the chemical reactivity of RuO2, leading to an enhancement in specific capacitance.

2. Materials and Methods

2.1. Material Preparation

All chemicals were used as received without further purification. Ti3C2 (11 technology Co., Ltd., Changchun, China). RuCl3·3H2O, NaOH, polytetrafluoroethylene (PTFE, 60%), H2SO4, NaH2PO2•H2O and other chemical reagents used in experiments were purchased from Adamas.

2.2. Synthesis of Phosphate Ion-Modified RuO2/Ti3C2

In a typical preparation process, 50 mg of Ti3C2 was added into an ethanol/water mixture with a ratio of 1:1 (v:v), followed by a sonication process for 1 h. Then, RuCl3·xH2O solution was dropped into the Ti3C2 suspension and stirred for 1 h. After that, the pH of the suspension was adjusted to 7 by 0.5 M NaOH solution. After being sonicated for 6 h, the RuO2/Ti3C2 composite was collected by centrifugation at 8000 rpm, rinsed with deionized waterand ethanol, then dried at 60 °C overnight. To obtain PRT, RuO2/Ti3C2 composite and NaH2PO2•H2O powder were placed in a combustion boat and annealed in a tube furnace filled with an Ar atmosphere at 250 °C for 1 h.
In order to obtain optimal samples, three kinds of PRT were prepared by using 30 mg, 60 mg and 90 mg RuCl3·xH2O during the synthesis process (denoted as PRT-30, PRT-60 and PRT-90, respectively). The PRT sample in Figure 5 refers to PRT-60.

2.3. Electrochemical Characterization

The working electrodes were prepared by pressing the mixture of 80 wt % as-prepared PRT materials, 10 wt % carbon black and 10 wt % PTFE binder onto stainless steel mesh. The mixture was then dried in a vacuum oven at 80 °C over 12 h. The pressure applied on the electrodes was 9 MPa. The mass loading of active material in every electrode was ~6 mg/cm2.Electrochemical measurements were carried out with a GAMRY Reference 3000 electrochemistry workstation in 1 M H2SO4 aqueous solution using a three-electrode mode. A platinum electrode was used as the counter electrode and an Ag/AgCl electrode was used as the reference electrode.

2.4. Material Characterization

The crystal structures of the composites were characterized by an X-ray diffractometer (XRD; D/Max 2500 V PC, Cu-Ka radiation, Rigaku, Osaka, Japan) and X-ray photoelectron spectroscopy (XPS; Escalab250, Thermo Fisher Scientific Inc., MA, USA). The Brunauer–Emmett–Teller (BET) surface area of the samples was characterized by using the Micromeritics surface area and porosimetry system (ASAP 2420, Atlanta, GA, USA ). The morphology of the products was characterized by scanning electron microscopy (SEM, HITACHI S-3400N, 5 Kv, Hitachi Ltd., Tokyo, Japan) and transmission electron microscopy (TEM, JEM-2100, 200 kV, JEOL Ltd., Osaka, Japan).

3. Results and Discussion

The synthesis process of PRT is shown in Figure 1. The phosphate ion functionalization mechanism can be described as follows: Na2PO2•H2O was decomposed into PH3 gas, H2O gas and Na2HPO4 at high temperatures. After that, RuO2 was reduced to RuO2−x by PH3 gas and H2PO4 was introduced onto the surface of RuO2−x at the same time. Finally, the phosphate ion-modified RuO2/Ti3C2 was collected.
The structure and morphology of the products were investigated by XRD. The XRD curves (Figure 2a) of RuO2 showed no discernible peaks, indicating the amorphous state of RuO2. For Ti3C2, a prominent peak appeared at 2ϴ = 6.90°, characteristic of the (002) plane of Ti3C2, which fitted well with other related results [39,40,41]. For PRT, the peaks of Ti3C2 became weaker compared with the pure Ti3C2, which was mainly influenced by the introduction of RuO2 on the surface of the Ti3C2. Compared with RuO2/Ti3C2, PRT showed no discernible change, indicating that no phase transformation occurred during the annealing process. The nitrogen adsorption and desorption isotherms of Ti3C2 and PRT are shown in Figure 2b. The BET surface area value of the PRT sample was 120.33 m2 g−1, while Ti3C2 only possessed a surface area of 23.64 m2 g−1. The increased surface area of PRT suggests that the introduction of RuO2 on the surface of the Ti3C2 can effectively inhibit the restacking effect of Ti3C2 and prevent RuO2 particles from aggregating.
XPS investigations were employed to survey the surface chemical composition and chemical bonding states of the products. Figure 3a shows the XPS survey of RuO2/Ti3C2 and PRT. Compared with RuO2/Ti3C2, the appearance of the P peak for PRT samples confirmed the phosphate doping. The O1s spectra at 531 eV shown in Figure 3b,c demonstrated the existence of RuO2 (the value from the literature is also 531 eV) [20]. The new characteristic peaks at 531.7 and 532.4 eV for PRT were related to the oxygen species of H2PO4 and PO3, respectively (values from the literature are 531.6 and 532.6 eV) [38,42]. The above results verify that the phosphate ions were successfully introduced on the surface of RuO2.
An SEM analysis was conducted to investigate the morphology of Ti3C2, RuO2/Ti3C2 and PRT. An SEM image of Ti3C2 is shown in Figure 4a. The impressive accordion-like layered structure of Ti3C2 can be observed. The interspaces between layers provide channels for material transfer and ion diffusion [36,37,43]. Figure 4b,c show the SEM images of RuO2/Ti3C2 and PRT, respectively. Obviously, the unique layered architecture was still maintained even after being subjected to a high-temperature process. At the same time, RuO2 particles were deposited on the surface of Ti3C2. Furthermore, PRT (Figure 4c) showed no morphology change compared with RuO2/Ti3C2 (Figure 4b), which is consistent with the XRD results. The layered architecture of PRT is expected to promote the reversible redox reactions and ion adsorption–desorption process.
The structures of Ti3C2, RuO2/Ti3C2 and PRT were also revealed by the TEM images. A typical TEM image of Ti3C2 sheets can be observed in Figure 4d. The transparent nanosheets of Ti3C2 mainly originated from the ultrasonic processing before the TEM was conducted. The TEM images of RuO2/Ti3C2 and PRT are shown in Figure 4e,f. Compared with Ti3C2 sheets (Figure 4d), nanoparticles were observed on the surface of Ti3C2. The Energy Dispersive Spectrometer (EDS) results (shown at the bottom of Figure 4g) further demonstrated the uniform distribution of RuO2 particles and phosphate doping. The dimensions of the RuO2 particles were calculated to be ~3.8 nm according to the TEM results (shown in Figure 4h). The appearance of the sodium element mainly comes from the residue of NaH2PO2•H2O. It has been reported that the phosphate ion functionalization strategy is an effective way to promote the performance of metal oxides [38,44]. After the phosphate ions were introduced onto the surface of metal oxides, longer bonds and smaller electronegativity can promote the surface reactivity and electrode kinetics of the metal oxides [45]. According to this theory, the phosphate ion-modified RuO2/Ti3C2 composite is expected to exhibit a better electrochemical performance than the pure RuO2 and RuO2/Ti3C2 samples when they are applied as supercapacitor electrode materials.
In order to verify the above inference, the electrochemical performances of Ti3C2, RuO2, RuO2/Ti3C2 and PRT-60 electrodes were first evaluated using a three-electrode system at the scan rate of 100 mV s−1 with 1 M H2SO4 as the solution. Figure 5a shows the cyclic voltammetry (CV) curves of these electrodes in a voltage window of −0.2 to 0.4 V (vs. Ag/AgCl). At any given scanning rate, the different integral areas of the CV curve indicated the difference in capacity [46,47]. Compared with Ti3C2, RuO2 and RuO2/Ti3C2 samples, the CV curve of PRT-60 showed the largest areas, indicating that it had the highest specific capacitance. Furthermore, even at the high scan rate of 100 mV s−1, the CV curve of PRT-60 still retained a relatively rectangular shape, implying the excellent rate capability and low contact resistance of the PRT-60 electrode. To investigate the electrochemical performances of the PRT-60 electrode, galvanostatic charge–discharge (GCD) cycling experiments were conducted at a current density of 2 A g−1. As shown in Figure 5b, the symmetrical triangle shape of the GCD curves with a small internal resistance (IR)-drop demonstrated the excellent reversibility and charge–discharge properties. The electrode’s specific capacity was calculated according to the equation: C S = I Δ t / m Δ V , where C S (F g−1) is the mass specific capacitance, I (A) is the discharge current, Δ t (s) is the discharge time, m (g) is mass loading of active component and Δ V (V) is potential window [48]. The specific capacitances of Ti3C2, RuO2, RuO2/Ti3C2 and PRT-60 electrodes were calculated to be 24.97 F g−1, 125.7 F g−1, 306.6 F g−1 and 612.7 F g−1, respectively. The PRT-60 electrode exhibited the maximum specific capacitance, which is in accordance with the CV results mentioned above. This improvement in electrochemical properties for the PRT-60 electrode can be attributed to the synergistic effects of phosphate ion modification and the introduction of Ti3C2.
To further investigate the advantages of these composite materials, an electrical impedance spectroscopy (EIS) measurement was conducted. The Nyquist impedance plots of the EIS results are shown in Figure 5c. As we know, the diameter of the semicircular curve at a high frequency represents the charge transfer resistance (Rct) at the electrode–electrolyte interface [49]. In particular, the RuO2/Ti3C2 electrode exhibited a lower Rct than that of pure RuO2, indicating that the layered Ti3C2 is beneficial for fast electrode kinetics. What is more, PRT-60 exhibited an even lower Rct than RuO2/Ti3C2, illustrating that phosphate ion modification helps to accelerate the charge transfer process.
The long-term cycle performance of electrodes is a crucial parameter in practical application. Therefore, the cycle stability of PRT-60 was evaluated by repeating the GCD test for 10,000 cycles at a current density of 2 A g−1. As shown in Figure 5d, the increase of the capacitance at the beginning cycles may be caused by the presence of the activation process in the electrode, which is commonly observed for metal oxides [49,50]. After 10,000 cycles, the specific capacitance was still maintained at 600.14 F g−1, and only decreased by 2.05% of the initial capacitance (612.72 F g−1).
The EIS analysis of PRT-60 was measured at the cycles 1 and 10,000. As shown in the insert of Figure 4d, the EIS values were almost similar, further demonstrating the electrochemical stability. In view of the above discussions, we conclude that the excellent performance of PRT-60 can be ascribed to the following reasons: First, the existence of Ti3C2 not only improves the electrical conductivity of the composite but also provides channel that facilitates the ion diffusion of the electrolyte. Furthermore, the phosphate ion-functionalized RuO2 provides a higher number of active reaction sites and accelerates the charge transfer process, resulting in an excellent electrochemical performance.
In order to obtain the optimized samples, three kinds of PRT with various RuO2 loadings were prepared by using 30 mg, 60 mg and 90 mg RuCl3·xH2O during the synthesis process (denoted as PRT-30, PRT-60 and PRT-90, respectively). CV curves with different scan rates are shown in Figure 6a–c. Obviously, the CV curves of all samples retained a typical rectangular shape at a low sweep rate, indicating a good charge transfer process [51]. However, at a high scanning rate of 200 mV s−1, the curves of PRT-30 and PRT-90 electrodes deviated from the rectangular shape, while PRT-60 still retained the ideal shape. This suggests that reasonable mass loading is beneficial for rate capability. This advantage also applies to the electrochemical capacity. As shown in Figure 6e and Table 1, the excellent rate capability of PRT-60 is also reflected by the histogram of the specific capacity at different current densities. When the current density was less than 2 A g−1, PRT-90 exhibited the highest capacity compared with the other two samples. However, with the increase of the current density, the capacity of all three samples showed a declining trend and the capacity of PRT-90 (328.67 F g−1) was smaller than that of PRT-60 (466.87 F g−1) when the current density reached 10 A g−1. Even at a high current density of 100 A g−1, the PRT-60 electrode still delivered a high capacitance of 320.83 F g−1, which was about 2.5 times larger than that of PRT-90 (128.65 F g−1) electrode and seven times larger than that of PRT-30 (45.85 F g−1). This result is in accordance with the CV results shown in Figure 6a–c. The obtained PRT-60 also showed better performance compared with other reported RuO2-based materials (as shown in Table 2). Figure 6f compares the Nyquist plots of the three samples. The electronic resistance (Rs) of these samples increased with the increase of ruthenium oxide content, further confirming the effect of Ti3C2 on enhancing electronic conductivity. When the proportion of Ti3C2 nanosheets decreased, the conductivity of the composites began to decrease as well.
According to the above analysis of electrochemical properties, it is evident that PRT-60 delivered better electrochemical performance than the other two samples. (I) For PRT-30, the low content of RuO2 led to a lower capacity compared with the other two samples. (II) For PRT-90, while it showed the highest discharge capacity at a low current density, its capacity attenuation was the most serious at a high current density (only 128.65 F g−1 at 100 A g−1). The poor rate capacity of PRT-90 is mainly due to the accumulation of excessive RuO2 nanoparticles, which occupy the spaces between the Ti3C2 layers and prevent the electrolyte from entering.

4. Conclusions

In summary, a phosphate ion-modified RuO2/Ti3C2 composite was successfully synthesized and applied in a supercapacitor electrode. The resulting PRT composite combines the advantages of Ti3C2, RuO2 and phosphate ion modification. Morphological and chemical characterization indicated that the Ti3C2 not only acts as a conductive substrate to prevent RuO2 nanoparticles from aggregating, but also provides electrolyte-filled channels for ion diffusion. Electrochemical studies showed that the phosphate ion modification can greatly improve the electrochemical properties of the composites. The PRT composite exhibited a specific capacitance of 612.7 F g−1 at 2.0 A g−1 and a satisfactory rate capacity (320.83 F g−1 at 100 A g−1). In addition, the capacitance of PRT could retain 97.95% (600.14 F g−1) of the original value after 10,000 cycles at a current density of 2 A g−1. More importantly, the method delivered in our experiment can be also applied to other energy storage materials.

Author Contributions

J.Z. and F.L. designed the experiments and analyzed the data; J.Z. performed the experiments and wrote the first draft of the manuscript; W.L. supervised the work and finalized the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Nos. 51879292, 51773106, 51708541), the National Science Fund for Distinguished Young Scholars (No. 51525903) and the Aoshan Talents Outstanding Scientist Program Supported by Qingdao National Laboratory for Marine Science and Technology (No. 2017ASTCP-OS09).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, Q.; Sun, D.; Wu, Q.; Wang, H.; Wang, L.; Liu, B.; Zhou, A.; He, J. MXene: A new family of promising hydrogen storage medium. J. Phys. Chem. A 2013, 117, 14253–14260. [Google Scholar] [CrossRef] [PubMed]
  2. Shen, C.; Wang, L.; Zhou, A.; Wang, B.; Wang, X.; Lian, W.; Hu, Q.; Qin, G.; Liu, X. Synthesis and Electrochemical Properties of Two-Dimensional RGO/Ti(3)C(2)Tx Nanocomposites. Nanomaterials (Basel) 2018, 8, 80. [Google Scholar] [CrossRef] [PubMed]
  3. Fan, G.; Li, X.; Xu, C.; Jiang, W.; Zhang, Y.; Gao, D.; Bi, J.; Wang, Y. Palladium Supported on Titanium Carbide: A Highly Efficient, Durable, and Recyclable Bifunctional Catalyst for the Transformation of 4-Chlorophenol and 4-Nitrophenol. Nanomaterials (Basel) 2018, 8, 141. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, Y.; Song, Y.; Xia, Y. Electrochemical capacitors: mechanism, materials, systems, characterization and applications. Chem. Soc. Rev. 2016, 45, 5925–5950. [Google Scholar] [CrossRef] [PubMed]
  5. Myung, Y.; Jung, S.; Tung, T.T.; Tripathi, K.M.; Kim, T. Graphene-Based Aerogels Derived from Biomass for Energy Storage and Environmental Remediation. ACS Sustain. Chem. Eng. 2019, 7, 3772–3782. [Google Scholar] [CrossRef]
  6. Chiu, H.Y.; Cho, C.P. Mixed-Phase MnO(2)/N-Containing Graphene Composites Applied as Electrode Active Materials for Flexible Asymmetric Solid-State Supercapacitors. Nanomaterials (Basel) 2018, 8, 924. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, L.; Zhao, J.; Li, W.-H.; Zhang, X.-Y.; Xu, H.-Z.; Li, K.; Kai, Y.; Sun, L.-S.; Li, C.-Q.; Liu, F.-Q. Synthesis of Ag nanoparticles decorated MnO2/sulfonated graphene composites with 3D macroporous structure for high performance capacitors electrode materials. RSC Adv. 2016, 6, 94682–94686. [Google Scholar] [CrossRef]
  8. Tyagi, A.; Tripathi, K.M.; Singh, N.; Choudhary, S.; Gupta, R.K. Green synthesis of carbon quantum dots from lemon peel waste: applications in sensing and photocatalysis. RSC Adv. 2016, 6, 72423–72432. [Google Scholar] [CrossRef]
  9. Zhang, J.; Jiang, J.; Li, H.; Zhao, X.S. A high-performance asymmetric supercapacitor fabricated with graphene-based electrodes. Energy Environ. Sci. 2011, 4, 4009–4015. [Google Scholar] [CrossRef]
  10. Jiang, H.; Dai, Y.; Hu, Y.; Chen, W.; Li, C. Nanostructured Ternary Nanocomposite of rGO/CNTs/MnO2 for High-Rate Supercapacitors. ACS Sustain. Chem. Eng. 2014, 2, 70–74. [Google Scholar] [CrossRef]
  11. Wang, Y.; Dong, L.; Lai, G.; Wei, M.; Jiang, X.; Bai, L. Nitrogen-Doped Hierarchically Porous Carbons Derived from Polybenzoxazine for Enhanced Supercapacitor Performance. Nanomaterials (Basel) 2019, 9, 131. [Google Scholar] [CrossRef] [PubMed]
  12. Xu, C.; Xu, B.; Gu, Y.; Xiong, Z.; Sun, J.; Zhao, X.S. Graphene-based electrodes for electrochemical energy storage. Energy Environ. Sci. 2013, 6, 1388–1414. [Google Scholar] [CrossRef]
  13. Bonaccorso, F.; Colombo, L.; Yu, G.; Stoller, M.; Tozzini, V.; Ferrari, A.C.; Ruoff, R.S.; Pellegrini, V. 2D materials. Graphene, related two-dimensional crystals, and hybrid systems for energy conversion and storage. Science 2015, 347, 1246501. [Google Scholar] [CrossRef] [PubMed]
  14. Nardecchia, S.; Carriazo, D.; Ferrer, M.L.; Gutierrez, M.C.; del Monte, F. Three dimensional macroporous architectures and aerogels built of carbon nanotubes and/or graphene: Synthesis and applications. Chem. Soc. Rev. 2013, 42, 794–830. [Google Scholar] [CrossRef] [PubMed]
  15. Yan, J.; Wang, Q.; Wei, T.; Fan, Z. Recent Advances in Design and Fabrication of Electrochemical Supercapacitors with High Energy Densities. Adv. Energy Mater. 2014, 4, 1300816. [Google Scholar] [CrossRef]
  16. Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, Y.; Guo, J.; Wang, T.; Shao, J.; Wang, D.; Yang, Y.W. Mesoporous Transition Metal Oxides for Supercapacitors. Nanomaterials (Basel) 2015, 5, 1667–1689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Tran, T.S.; Tripathi, K.M.; Kim, B.N.; You, I.-K.; Park, B.J.; Han, Y.H.; Kim, T. Three-dimensionally assembled Graphene/α-MnO2 nanowire hybrid hydrogels for high performance supercapacitors. Mater. Res. Bull. 2017, 96, 395–404. [Google Scholar] [CrossRef]
  19. Kim, J.Y.; Kim, K.H.; Yoon, S.B.; Kim, H.K.; Park, S.H.; Kim, K.B. In situ chemical synthesis of ruthenium oxide/reduced graphene oxide nanocomposites for electrochemical capacitor applications. Nanoscale 2013, 5, 6804–6811. [Google Scholar] [CrossRef] [PubMed]
  20. Hwang, J.Y.; El-Kady, M.F.; Wang, Y.; Wang, L.; Shao, Y.; Marsh, K.; Ko, J.M.; Kaner, R.B. Direct preparation and processing of graphene/RuO2 nanocomposite electrodes for high-performance capacitive energy storage. Nano Energy 2015, 18, 57–70. [Google Scholar] [CrossRef]
  21. Gui, Z.; Gillette, E.; Duay, J.; Hu, J.; Kim, N.; Lee, S.B. Co-electrodeposition of RuO2-MnO2 nanowires and the contribution of RuO2 to the capacitance increase. Phys. Chem. Chem. Phys. PCCP 2015, 17, 15173–15180. [Google Scholar] [CrossRef] [PubMed]
  22. Deshmukh, P.R.; Bulakhe, R.N.; Pusawale, S.N.; Sartale, S.D.; Lokhande, C.D. Polyaniline–RuO2 composite for high performance supercapacitors: Chemical synthesis and properties. RSC Adv. 2015, 5, 28687–28695. [Google Scholar] [CrossRef]
  23. Meng, Y.; Wang, L.; Xiao, H.; Ma, Y.; Chao, L.; Xie, Q. Facile electrochemical preparation of a composite film of ruthenium dioxide and carboxylated graphene for a high performance supercapacitor. RSC Adv. 2016, 6, 33666–33675. [Google Scholar] [CrossRef]
  24. Gao, Z.D.; Zhu, X.; Li, Y.H.; Zhou, X.; Song, Y.Y.; Schmuki, P. Carbon cladded TiO2 nanotubes: Fabrication and use in 3D-RuO2 based supercapacitors. Chem. Commun. 2015, 51, 7614–7617. [Google Scholar] [CrossRef] [PubMed]
  25. Zhu, Y.; Ji, X.; Pan, C.; Sun, Q.; Song, W.; Fang, L.; Chen, Q.; Banks, C.E. A carbon quantum dot decorated RuO2 network: Outstanding supercapacitances under ultrafast charge and discharge. Energy Environ. Sci. 2013, 6, 3665–3675. [Google Scholar] [CrossRef]
  26. Ma, H.; Kong, D.; Xu, Y.; Xie, X.; Tao, Y.; Xiao, Z.; Lv, W.; Jang, H.D.; Huang, J.; Yang, Q.H. Disassembly-Reassembly Approach to RuO2/Graphene Composites for Ultrahigh Volumetric Capacitance Supercapacitor. Small 2017, 13, 1701026. [Google Scholar] [CrossRef] [PubMed]
  27. Lin, N.; Tian, J.; Shan, Z.; Chen, K.; Liao, W. Hydrothermal synthesis of hydrous ruthenium oxide/graphene sheets for high-performance supercapacitors. Electrochim. Acta 2013, 99, 219–224. [Google Scholar] [CrossRef]
  28. Zhang, C.; Zhou, H.; Yu, X.; Shan, D.; Ye, T.; Huang, Z.; Kuang, Y. Synthesis of RuO2 decorated quasi graphene nanosheets and their application in supercapacitors. RSC Adv. 2014, 4, 11197–11205. [Google Scholar] [CrossRef]
  29. Amir, F.Z.; Pham, V.H.; Dickerson, J.H. Facile synthesis of ultra-small ruthenium oxide nanoparticles anchored on reduced graphene oxide nanosheets for high-performance supercapacitors. RSC Adv. 2015, 5, 67638–67645. [Google Scholar] [CrossRef]
  30. Shen, B.; Zhang, X.; Guo, R.; Lang, J.; Chen, J.; Yan, X. Carbon encapsulated RuO2 nano-dots anchoring on graphene as an electrode for asymmetric supercapacitors with ultralong cycle life in an ionic liquid electrolyte. J. Mater. Chem. A 2016, 4, 8180–8189. [Google Scholar] [CrossRef]
  31. Sun, D.; Wang, M.; Li, Z.; Fan, G.; Fan, L.-Z.; Zhou, A. Two-dimensional Ti3C2 as anode material for Li-ion batteries. Electrochem. Commun. 2014, 47, 80–83. [Google Scholar] [CrossRef]
  32. Halim, J.; Lukatskaya, M.R.; Cook, K.M.; Lu, J.; Smith, C.R.; Naslund, L.A.; May, S.J.; Hultman, L.; Gogotsi, Y.; Eklund, P.; et al. Transparent Conductive Two-Dimensional Titanium Carbide Epitaxial Thin Films. Chem. Mater. Publ. Am. Chem. Soc. 2014, 26, 2374–2381. [Google Scholar] [CrossRef] [PubMed]
  33. Naguib, M.; Halim, J.; Lu, J.; Cook, K.M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. New two-dimensional niobium and vanadium carbides as promising materials for Li-ion batteries. J. Am. Chem. Soc. 2013, 135, 15966–15969. [Google Scholar] [CrossRef] [PubMed]
  34. Kurra, N.; Ahmed, B.; Gogotsi, Y.; Alshareef, H.N. MXene-on-Paper Coplanar Microsupercapacitors. Adv. Energy Mater. 2016, 6, 1601372. [Google Scholar] [CrossRef]
  35. Peng, Y.-Y.; Akuzum, B.; Kurra, N.; Zhao, M.-Q.; Alhabeb, M.; Anasori, B.; Kumbur, E.C.; Alshareef, H.N.; Ger, M.-D.; Gogotsi, Y. All-MXene (2D titanium carbide) solid-state microsupercapacitors for on-chip energy storage. Energy Environ. Sci. 2016, 9, 2847–2854. [Google Scholar] [CrossRef] [Green Version]
  36. Xiong, J.; Pan, L.; Wang, H.; Du, F.; Chen, Y.; Yang, J.; Zhang, C. Synergistically enhanced lithium storage performance based on titanium carbide nanosheets (MXene) backbone and SnO2 quantum dots. Electrochim. Acta 2018, 268, 503–511. [Google Scholar] [CrossRef]
  37. Rakhi, R.B.; Ahmed, B.; Anjum, D.; Alshareef, H.N. Direct Chemical Synthesis of MnO2 Nanowhiskers on Transition-Metal Carbide Surfaces for Supercapacitor Applications. ACS Appl. Mater. interfaces 2016, 8, 18806–18814. [Google Scholar] [CrossRef] [PubMed]
  38. Zhai, T.; Wan, L.; Sun, S.; Chen, Q.; Sun, J.; Xia, Q.; Xia, H. Phosphate Ion Functionalized Co3O4 Ultrathin Nanosheets with Greatly Improved Surface Reactivity for High Performance Pseudocapacitors. Adv. Mater. 2017, 29, 1604167. [Google Scholar] [CrossRef] [PubMed]
  39. Lian, P.; Dong, Y.; Wu, Z.-S.; Zheng, S.; Wang, X.; Sen, W.; Sun, C.; Qin, J.; Shi, X.; Bao, X. Alkalized Ti 3C2 MXene nanoribbons with expanded interlayer spacing for high-capacity sodium and potassium ion batteries. Nano Energy 2017, 40, 1–8. [Google Scholar] [CrossRef]
  40. Ahmed, B.; Anjum, D.H.; Gogotsi, Y.; Alshareef, H.N. Atomic layer deposition of SnO2 on MXene for Li-ion battery anodes. Nano Energy 2017, 34, 249–256. [Google Scholar] [CrossRef]
  41. Xu, S.; Wei, G.; Li, J.; Ji, Y.; Klyui, N.; Izotov, V.; Han, W. Binder-free Ti 3C2Tx MXene electrode film for supercapacitor produced by electrophoretic deposition method. Chem. Eng. J. 2017, 317, 1026–1036. [Google Scholar] [CrossRef]
  42. Ghosh, K.; Balog, E.R.M.; Sista, P.; Williams, D.J.; Kelly, D.; Martinez, J.S.; Rocha, R.C. Temperature-dependent morphology of hybrid nanoflowers from elastin-like polypeptides. APL Mater. 2014, 2, 021101. [Google Scholar] [CrossRef]
  43. Xie, X.; Zhao, M.-Q.; Anasori, B.; Maleski, K.; Ren, C.E.; Li, J.; Byles, B.W.; Pomerantseva, E.; Wang, G.; Gogotsi, Y. Porous heterostructured MXene/carbon nanotube composite paper with high volumetric capacity for sodium-based energy storage devices. Nano Energy 2016, 26, 513–523. [Google Scholar] [CrossRef]
  44. Jo, W.J.; Jang, J.W.; Kong, K.J.; Kang, H.J.; Kim, J.Y.; Jun, H.; Parmar, K.P.; Lee, J.S. Phosphate doping into monoclinic BiVO4 for enhanced photoelectrochemical water oxidation activity. Angew. Chem. 2012, 51, 3147–3151. [Google Scholar] [CrossRef] [PubMed]
  45. Huang, Y.; Zeng, Y.; Yu, M.; Liu, P.; Tong, Y.; Cheng, F.; Lu, X. Recent Smart Methods for Achieving High-Energy Asymmetric Supercapacitors. Small Method. 2018, 2, 1700230. [Google Scholar] [CrossRef]
  46. Chen, H.; Hu, L.; Chen, M.; Yan, Y.; Wu, L. Nickel-Cobalt Layered Double Hydroxide Nanosheets for High-performance Supercapacitor Electrode Materials. Adv. Funct. Mater. 2014, 24, 934–942. [Google Scholar] [CrossRef]
  47. Pang, M.; Long, G.; Jiang, S.; Ji, Y.; Han, W.; Wang, B.; Liu, X.; Xi, Y. Rapid synthesis of graphene/amorphous α-MnO2 composite with enhanced electrochemical performance for electrochemical capacitor. Mater. Sci. Eng. B 2015, 194, 41–47. [Google Scholar] [CrossRef]
  48. Zhang, J.; Feng, H.; Yang, J.; Qin, Q.; Fan, H.; Wei, C.; Zheng, W. Solvothermal Synthesis of Three-Dimensional Hierarchical CuS Microspheres from a Cu-Based Ionic Liquid Precursor for High-Performance Asymmetric Supercapacitors. ACS Appl. Mater. Interfaces 2015, 7, 21735–21744. [Google Scholar] [CrossRef] [PubMed]
  49. Yan, J.; Fan, Z.; Wei, T.; Qian, W.; Zhang, M.; Wei, F. Fast and reversible surface redox reaction of graphene–MnO2 composites as supercapacitor electrodes. Carbon 2010, 48, 3825–3833. [Google Scholar] [CrossRef]
  50. Xu, X.; Pei, L.; Yang, Y.; Shen, J.; Ye, M. Facile synthesis of NiWO4/reduced graphene oxide nanocomposite with excellent capacitive performance for supercapacitors. J. Alloys Compd. 2016, 654, 23–31. [Google Scholar] [CrossRef]
  51. Zhang, Z.; Xiao, F.; Qian, L.; Xiao, J.; Wang, S.; Liu, Y. Facile Synthesis of 3D MnO2–Graphene and Carbon Nanotube–Graphene Composite Networks for High-Performance, Flexible, All-Solid-State Asymmetric Supercapacitors. Adv. Energy Mater. 2014, 4, 1400064. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the procedure for preparing the phosphate ion-modified RuO2/ Ti3C2 composite (PRT).
Figure 1. Schematic illustration of the procedure for preparing the phosphate ion-modified RuO2/ Ti3C2 composite (PRT).
Nanomaterials 09 00377 g001
Figure 2. (a) X-ray diffractometry (XRD) patterns of RuO2, Ti3C2, RuO2/Ti3C2 and PRT; (b) BET nitrogen adsorption–desorption isotherms.
Figure 2. (a) X-ray diffractometry (XRD) patterns of RuO2, Ti3C2, RuO2/Ti3C2 and PRT; (b) BET nitrogen adsorption–desorption isotherms.
Nanomaterials 09 00377 g002
Figure 3. (a) X-ray photoelectron spectroscopy (XPS) survey of RuO2/Ti3C2 and PRT. High-resolution O1s XPS spectrum of (b) RuO2/Ti3C2 and (c) PRT.
Figure 3. (a) X-ray photoelectron spectroscopy (XPS) survey of RuO2/Ti3C2 and PRT. High-resolution O1s XPS spectrum of (b) RuO2/Ti3C2 and (c) PRT.
Nanomaterials 09 00377 g003
Figure 4. (a) SEM image of Ti3C2; (b) SEM image of RuO2/Ti3C2; (c) SEM image of PRT; (d) TEM image of Ti3C2; (e) TEM image of RuO2/Ti3C2; (f) TEM image of PRT; (g) element mapping of PRT; (h) size histogram of RuO2.
Figure 4. (a) SEM image of Ti3C2; (b) SEM image of RuO2/Ti3C2; (c) SEM image of PRT; (d) TEM image of Ti3C2; (e) TEM image of RuO2/Ti3C2; (f) TEM image of PRT; (g) element mapping of PRT; (h) size histogram of RuO2.
Nanomaterials 09 00377 g004
Figure 5. (a) Cyclic voltammetry (CV) curves, (b) galvanostatic charge–discharge (GCD) curves and (c) Nyquist plots of the Ti3C2, RuO2, RuO2/Ti3C2 and PRT-60 electrodes; (d) long-term cycling test by measuring 10,000 cycles at a constant current density of 2 A g−1 (the insert shows the EIS values cycles 1 and 10,000).
Figure 5. (a) Cyclic voltammetry (CV) curves, (b) galvanostatic charge–discharge (GCD) curves and (c) Nyquist plots of the Ti3C2, RuO2, RuO2/Ti3C2 and PRT-60 electrodes; (d) long-term cycling test by measuring 10,000 cycles at a constant current density of 2 A g−1 (the insert shows the EIS values cycles 1 and 10,000).
Nanomaterials 09 00377 g005
Figure 6. CV curves of (a) PRT-30, (b) PRT-60 and (c) PRT-90 electrodes at scan rates varying from 10 to 200 mV s−1; (d) GCD curves of the three kinds of samples; (e) histogram of the specific capacity for the three kinds of samples at different current densities; (f) Nyquist plots of the samples (the inset is the expanded view of the high-frequency range).
Figure 6. CV curves of (a) PRT-30, (b) PRT-60 and (c) PRT-90 electrodes at scan rates varying from 10 to 200 mV s−1; (d) GCD curves of the three kinds of samples; (e) histogram of the specific capacity for the three kinds of samples at different current densities; (f) Nyquist plots of the samples (the inset is the expanded view of the high-frequency range).
Nanomaterials 09 00377 g006
Table 1. Specific capacitance of PRT-30, PRT-60 and PRT-90 at different current densities.
Table 1. Specific capacitance of PRT-30, PRT-60 and PRT-90 at different current densities.
SamplesSpecific Capacitance (F g−1)
0.5 A g−11 A g−12 A g−15 A g−110 A g−150 A g−1100 A g−1
PRT-30585.04484.63384.62199.91100.2566.9245.85
PRT-60768.03693.02612.72578.02466.87380.64320.83
PRT-901004.3888.54737.59474.42328.67220.52128.65
Table 2. Comparison of PRT with other reported RuO2 capacitors.
Table 2. Comparison of PRT with other reported RuO2 capacitors.
MaterialCapacitance (F g−1)Cycle Life (Cycles)Reference
Reduced graphene oxide sheets modified with RuO24002500[9]
RuO2/reduced graphene oxide nanocomposites4891000[19]
RuO2 deposited on the surface of graphene sheets5512000[27]
Reduced graphene oxide–RuO2 hybrid materials5092000[29]
Phosphate ion-modified RuO2/Ti3C2 composite69310,000This work

Share and Cite

MDPI and ACS Style

Zhao, J.; Liu, F.; Li, W. Phosphate Ion-Modified RuO2/Ti3C2 Composite as a High-Performance Supercapacitor Material. Nanomaterials 2019, 9, 377. https://doi.org/10.3390/nano9030377

AMA Style

Zhao J, Liu F, Li W. Phosphate Ion-Modified RuO2/Ti3C2 Composite as a High-Performance Supercapacitor Material. Nanomaterials. 2019; 9(3):377. https://doi.org/10.3390/nano9030377

Chicago/Turabian Style

Zhao, Jie, Faqian Liu, and Weihua Li. 2019. "Phosphate Ion-Modified RuO2/Ti3C2 Composite as a High-Performance Supercapacitor Material" Nanomaterials 9, no. 3: 377. https://doi.org/10.3390/nano9030377

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