Skip to main content
  • Research article
  • Open access
  • Published:

Electrochemical study of TiO2 in aqueous AlCl3 electrolyte via vacuum impregnation for superior high-rate electrode performance

Abstract

This communication elucidates the charge storage mechanism of a TiO2 electrode in 1 mol dm− 3 AlCl3 for use in aqueous-ion batteries. Cyclic voltammetry studies suggest a surface contribution to charge storage and that cycle life can be improved by limiting the potential ≥ − 1.0 V vs SCE. In order to enhance this surface contribution, a simple vacuum impregnation technique was employed to improve electrode-electrolyte contact. This resulted in a significant improvement in the high rate performance of TiO2, where a capacity of 15 mA h g− 1 was maintained at the very high specific current of 40 A g− 1, a decrease of only 25% from when the electrode was cycled at 1 A g− 1. The vacuum impregnation process was also applied to copper-hexacyanoferrate, envisaged as a possible positive electrode, again resulting in significant improvements to high-rate performance. This demonstrates the potential for using this simple technique for improving electrode performance in other aqueous electrolyte battery systems.

Introduction

Asymmetric and hybrid devices based on a combination of capacitive, psuedocapacitive or battery intercalation electrodes have gained interest lately due to performance characteristics that could bridge the gap between the high energy density of Li-ion chemistries and high power of supercapacitors. Furthermore, there is interest in the use of aqueous electrolytes, which can provide advantages in terms of cost, ease of handling, toxicity, and environmental benignity. The use of electrolytes based on Na, K or Al salts also benefit from the higher natural abundance and relative ease of processing of these materials compared to Li salts and organic solvents.

Excluding Li-ion, negative electrodes for aqueous metal-ion systems are relatively limited with NaTi2(PO4)3 and metallic Zn commonly used in aqueous Na-ion and Zn-ion cells [1,2,3,4]. Similarly, while a variety of positive electrodes have been explored for aqueous super/psuedocapacitors, such as MnO2 [5,6,7], RuO2 [8,9,10], Ni (OH)2 [11], Co (OH)2 [12], Co3O4 [13] or Prussian blue analogues [14, 15], the number of negative electrodes is far lower. As such, a non-exhaustive review of aqueous super/pseudo-capacitive devices, reported since 2014, shows that the majority of negative electrodes consist of activated carbon (AC) or other carbon-based materials, as shown in Additional file 1: Table S1. A recent review on asymmetric supercapacitors by Shao et al., further illustrates this, where the majority of studies and devices make use of AC, graphene or graphene oxide as the negative electrode [16]. The use of AC is likely to limit cell voltage and capacity while graphene materials can result in high cost and low scalability. TiO2 provides a possible option for a negative electrode, having been studied in aqueous aluminium salt electrolytes and shown to have working potentials lower than ca. < − 0.5 V vs SCE [15,16,17,18,19,20], presenting the opportunity for dual-ion devices working at higher voltages [21,22,23,24]. TiO2 nanotube arrays, synthesised by Liu et al., allowed a capacity of ca. 75 mA h g− 1 when cycled at 4 mA cm− 2 in 1 mol dm− 3 AlCl3 [17]. He et al., have provided evidence for Al3+ intercalation, where XRD analysis showed anatase-TiO2 lattice parameters changing with state-of-charge [19]. However, capacitive or surface contributions to charge storage cannot be entirely ruled out, especially at high rates. For example, cathodic redox peaks from TiO2, associated with Al3+ insertion, are at more negative potentials than the plateaus observed during constant current cycling [17, 19, 20]. Furthermore, a graphene incorporated TiO2 electrode studied by Lahan et al., provided a capacity of approximately 20 mA h g− 1 at 6.25 A g− 1, though the electrode showed very limited redox peaks during CV scans, suggesting the possibility of a capacitive or psuedocapacitive mechanism [1]. Previous work has also shown high rate capability, up to 360 C (7.2 A g− 1), from commercial TiO2 nanopowders, though relatively low capacities were measured [25].

Building upon previous work, this communication elucidates the charge storage mechanism of commercial TiO2 powder electrodes in 1 mol dm− 3 AlCl3 and demonstrates a TiO2 electrode capable of stable cycling at 40.0 A g− 1 with close to 100% charge efficiency.

Experimental procedures

TiO2 electrodes were manufactured using a 5 nm TiO2 powder purchased from US-nano. Electrodes consisted of 85 wt% TiO2, 5 wt% carbon black (Vulcan 72-CR), 6 wt% Nafion and 4 wt% PTFE. TiO2, carbon black, Nafion and propanol (approximately 3 g for 1 g dry powder) were mixed using a Silverson high speed shear mixer at 5000 rpm for 30 min. PTFE was then added with the ink sonicated for a further 15 min. The ink is coated onto carbon polymer (Sigracell PV15) current collectors with excess allowed to drip off before being laid flat to dry in ambient conditions. Once dry, additional layers were added to manufacture electrodes with mass loadings as high as 6.5 mg cm− 2 covering an area of 7 cm2.

Brunauer-Emmett-Teller (BET) surface area and pore size distribution were calculated from N2 adsorption-desorption isotherms, measured at − 196 °C, using a Gemini 2375 analyser.

Cyclic voltammetry and constant current cycling were performed in standard glass 3-electrode cells using a saturated calomel electrode (SCE) as reference and an oversized CuHCF, typically 7–8 times by mass, as a reversible counter electrode. Cyclic voltammetry was performed in 1 mol dm− 3 AlCl3 while constant current cycling was performed in an electrolyte comprising 1 mol dm− 3 AlCl3 and 1 mol dm− 3 KCl. A solartron 1470E battery analyser was used for constant current cycling while cyclic voltammetry was performed on an Ivium-n-Stat potentiostat.

Vacuum impregnation, previously used by Yong et al. for impregnation of textile supercapacitor electrodes [26], was performed using the experimental set-up shown in Fig. 1a with a proposed schematic of how electrolyte is forced into electrode pores being presented in Fig. 1b. This method was performed by placing electrodes, submerged in the electrolyte of 1 mol dm− 3 AlCl3/1 mol dm− 3 KCl, in a Buchi tube. A filtration vacuum pump was used to create a vacuum of 20 mbar within the tube for approximately 15 min. The air vent was then opened, letting air in, before the process was repeated a further three times by which point air bubbles were no longer visible at the electrode surface.

Fig. 1
figure 1

(a) Buchi-tube/pump set-up employed for vacuum impregnation of composite electrodes. b proposed process of forced electrode wetting

Results and discussion

Figure 2a shows the N2 adsorption-desorption BET isotherm from the TiO2 nanopowder. The BET surface area was calculated to be 269 m2 g− 1. The profile resembles a type IV isotherm according to the IUPAC classification. Fig. 2b shows a pore size distribution between 25 Å to 100 Å with a dominant peak at around 55 Å. This corresponds well with the isotherm in Fig. 2a, which demonstrates the mesoporous nature (2 nm – 50 nm) of the powder.

Fig. 2
figure 2

(a) N2 adsorption and desorption isotherm and (b) pores size distribution from the used TiO2 nanopowder

Cyclic voltammetry of two TiO2 electrodes was performed at various scan rates, ν, in 1 mol dm− 3 AlCl3 aqueous solution. One electrode was scanned between 0 V to − 1.3 V vs SCE and the other between the potential range of 0 V to − 1.0 V vs SCE. By limiting the potential window, charge storage may be limited to a capacitive or surface controlled mechanism. Figure 3a presents the profiles measured from TiO2 at the 5th, 18th and 25th cycles when swept between the extended potential range of 0 V to − 1.3 V vs SCE. Between the 5th and 18th cycle there is small reduction in the cathodic peak, from − 13.0 A g− 1 to − 11.8 A g− 1, while the anodic peak potential shifts from − 1.03 V, during the 10th scan, to − 0.97 V vs SCE during the 18th. Figure 3b shows the profile from TiO2 during the 5th, 25th and 80th scan at 10 mV s− 1 between 0 V to − 1.0 V vs SCE. The profiles can be seen to be nearly identical irrespective of scan number, suggesting improved stability from TiO2 when cycled at a more positive minimum potential.

Fig. 3
figure 3

CV responses from TiO2 at 10 mV s− 1. a shows the 5th, 18th and 25th cycles when scanned between potential limits of 0 V to − 1.3 V vs SCE. b shows the 5th, 25th and 80th scan of a separate electrode with potential limits of 0 V to − 1.0 V vs SCE

Figure 4a shows the CV scans at 3, 9 and 16 mV s− 1 between 0 V to − 1.3 V vs SCE. The profile shapes at these three scan rates closely resemble each other. During the cathodic sweep, current curves down between − 0.55 V to − 0.95 V where there is a brief plateau till ca. -1.1 V. The current curves down to a prominent peak between − 1.15 V and − 1.20 V vs SCE. At 3 mV s− 1, the reverse sweep gives rise to a prominent peak at − 1.05 V. The position of this peak becomes more positive with increasing scan rate with peak position being approximately − 0.95 V at 16 mV s− 1. As with the cathodic sweep, the anodic sweep gives rise to a slight shoulder and plateau – between ca. -0.9 V and − 0.75 V, when current drops steadily to zero at approximately − 0.5 V. Fig. 4b gives the peak currents against the square root of the scan rates. A linear fit, with an x-y intercept set to zero, shows there is an approximately linear relationship between the measured current and square root of the scan rate for both cathodic and anodic sweeps. A linear relationship suggests a diffusion-limited process, as described by the power law given by equation… (1), where a and b are adjustable values, i is the measured current and ν the scan rate [27, 28].

$$ i=a{\nu}^b $$
(1)
Fig. 4
figure 4

(a) and (b) give the CV scans and peak currents from TiO2, in 1 mol dm− 3 AlCl3, at scan rates between 3 mV s− 1 to 16 mV s− 1 between potentials limits of 0 V to − 1.3 V vs SCE. c and d give the normalised CV scans and peak currents from TiO2 at scan rates between 2 mV s− 1 to 100 mV s− 1 with potential limits set between 0 V to − 1.0 V vs SCE

A b-value of 0.5 is often measured from intercalation electrodes, with the measured current limited by the solid-state diffusion (intercalation) of the cation through the electrode. This may be true for the case of TiO2 and Al3+, given the use of a relatively high concentration electrolyte, which should negate the possibility of a reaction being limited by the diffusion of Al3+ through the electrolyte to the electrode surface.

However, the greater stability of TiO2 when scanned with the more positive minimum potential of − 1.0 V vs SCE, compared to − 1.3 V, suggests the possibility of a separate charge storage mechanism compared to when the electrode is scanned to − 1.3 V. That is, the redox reaction of Ti4+ to Ti3+ may only take place once more negative potentials are reached. As such, further CV scans were performed between 0 V to − 1.0 V vs SCE. Fig. 4c shows these CV profiles at scan rates between 2 mV s− 1 to 100 mV s− 1, normalised by scan rate. That the profiles do not fall onto a single profile means that charge storage in this potential range is not purely capacitive. Further analysis of the CV profiles can be performed by calculating the capacity of the electrodes at different scan rates. This technique has previously been used in the literature with materials, such as Nb2O5, NiCo2O4, LaB6, conductive polymers and for Li+ insertion into mesoporous titania [29,30,31,32]. The analysis can provide an indication of charge storage arising from bulk or surface mechanisms at given scan rates. Fig. 4d shows the cathodic and anodic voltammetric capacities against ν-1/2. For the cathodic charge input, the volumetric capacity is linearly proportional to ν-1/2 at scan rates up to 30 mV s− 1, (0.182 mV s− 1)-1/2. Extrapolation of the linear fit to 0 (mV s− 1)-1/2 suggests a surface charge storage contribution of approximately 12 mA h g− 1. Therefore, at a scan rate of 10 mV s− 1, for example, the surface contribution to capacity would be approximately 50%. The remaining charge could then be the result of a bulk process such as intercalation. Alternatively, it could suggest there are areas of the electrode, such as narrow pores, that are difficult to access. At scan rates above 30 mV s− 1, the charge vs ν-1/2 plot deviates from linearity, suggesting a change in the rate-limiting charge storage process or that charge storage is almost entirely dominated by a semi-infinite diffusion. At lower scan rates, between 2 to 30 mV s− 1, the extrapolation of the linear dependence of cathodic capacity vs ν-1/2, to approximately 12 mA h g− 1, suggests that charge storage is diffusion controlled. Given the low capacities, it is still unlikely that this diffusion limitation is a result of Al3+ intercalation through the crystal structure of anatase-TiO2 but may instead be due to the limited diffusion of electrolyte and Al3+, due to the short time constants at these high scan rates, through the electrodes pores. While there may be a capacitive contribution, as deduced from the extrapolation of the infinite scan rate capacity, the non-conformity of the normalised scan rates suggests there is also a diffusion controlled charge storage mechanism.

The existence of a surface controlled storage mechanism, along with the mesoporous structure of the 5 nm TiO2 powder (Fig. 2), suggests performance can be improved through greater electrolyte-electrode contact. In order to achieve this, a simple vacuum impregnation technique was employed to ensure proper electrode wetting. The experimental set-up and proposed schematic of forced electrode wetting were presented in Fig. 1. It is proposed that electrode pores previously inaccessible to electrolyte, due to surface tension and the hydrophobicity of the nanopowder electrode, are filled with electrolyte due to the removal of air and creation of low pressure voids within the electrode. Constant current cycling was then performed on a vacuum impregnated electrode in a 3-electrode cell between 0.4 V to − 1.0 V vs SCE. The coulombic efficiency and discharge capacity of the vacuum impregnated electrode when cycled at specific currents between 0.2 to 40.0 A g− 1 is shown in Fig. 5a. The figure shows the 10th cycle at a given specific current between cycles 70–120 for as-manufactured TiO2 and cycles 70–200 for impregnated TiO2. For comparison, the performance of an as-manufactured electrode, when cycled up to 6.0 A g− 1, is also shown in Fig. 5a. Additional file 1: Figure S1 shows the discharge capacity and coulombic efficiency of the two electrodes vs cycle number. Between 0.2 A g− 1 and 1.0 A g− 1, discharge capacity from the vacuum impregnated electrode decreases from 21.8 mA h g− 1 to 19.8 mA h g− 1, with coulombic efficiency increasing from 89.8 to 96.9%. At 2.0 A g− 1, coulombic efficiency was 99.4%, though discharge capacity was also measured at 19.8 mA h g− 1. Between 1.0 A g− 1 to 25 A g− 1, discharge capacity decreased by only 12.2% to 17.4 mA h g− 1. At 40.0 A g− 1, discharge capacity was measured at 15.3 mA h g− 1. Above 2.0 A g− 1, coulombic efficiency remained around 99.9%, though some error will be present due to the rapid charge discharge times, i.e. at 40.0 A g− 1 discharge occurs in 1.43 s, even at the used measurement rate of 80 data points per second. Coulombic efficiency of an as-manufactured electrode is lower throughout and while discharge capacity is comparable up to 2.0 A g− 1, once cycled at 6.0 A g− 1, discharge capacity was measured at 15.7 mA h g− 1 compared to 19.33 mA h g− 1 for the vacuum impregnated electrode.

Fig. 5
figure 5

(a) Discharge capacity and coulombic efficiency of a vacuum impregnated and a non-impregnated TiO2 electrode, as a function of specific current, when cycled in 1 mol dm− 3 AlCl3/1 mol dm− 3 KCl between 0.4 to − 1.0 V vs SCE. Corresponding charge-discharge curves of the impregnated TiO2 electrode cycled at 1.0, 10, 20, 30 and 40 A g− 1

The voltage profiles from the vacuum impregnated electrode between 1.0 A g− 1 to 40 A g− 1 are given by Fig. 5b. Voltage profiles can be seen to be similar, irrespective of the specific current used. The initial IR-drop at 1 A g− 1 is minimal, being less than 10 mV and only becoming noticeable at higher specific currents. At 10.0 A g− 1, the IR-drop is measured as 44 mV, increasing to 162 mV at 40.0 A g− 1, with the average charge and discharge potentials at 40.0 A g− 1 being − 0.826 V and − 0.627 V, respectively. For comparison, the IR drop from the as-manufactured electrode at 6 A g− 1 was already 124 mV. The results presented in Fig. 5 show a clear improvement in rate capability of electrodes subjected to the vacuum impregnation technique. This specific currents reached are considerably higher than have been previously reported for TiO2 in aqueous Al3+-containing electrolytes. It should also be noted that the experiment was performed on an electrode with a relatively high mass loading of 6.5 mg cm− 2, so that the corresponding current density at 40 A g− 1 is a very high value of 260 mA cm− 2. In comparison, capacities of 50 mA h g− 1 and ca. 62 mA h g− 1 were measured from MnHCF (positive) and graphene (negative) electrodes were achieved at the current density of 5 mA cm− 2 in LiNO3 [33, 34]. These capacities and current densities are toward the maximum reported for aqueous capacitive devices. Furthermore, the relative stability of the voltage profiles and capacity, where discharge capacity drops by < 25% over an order of magnitude increase in specific current, provides evidence that charge storage from these TiO2 electrodes in aqueous Al3+ electrolyte are predominantly capacitive or controlled by surface reactions at high currents, similar to psuedocapacitive materials. However, care should be taken in describing TiO2 as psuedocapacitive given the relatively clear voltage plateaus observed during constant current cycling, in aqueous Al3+-containing electrolytes, which is in contrast to the electrochemical characteristics of a capacitor.

The vacuum impregnation process was also repeated on a CuHCF electrode, envisaged as a potential positive electrode, with the effect on voltage profiles and capacities at various rates shown in Additional file 1: Figure S2. Capacity from the vacuum impregnated electrode, which had a mass loading of 8.8 mg cm− 2, was measured at 47.08 mA h g− 1 at 0.5 A g− 1 and maintained a capacity of 28.2 mA h g− 1 at 8 A g− 1. The capacity of the as-manufactured CuHCF electrode, with a mass loading of 8 mg cm− 2, was 44.42 mA h g− 1 at 0.5 A g− 1 and decreased to 14.1 mA h g− 1 at 6 A g− 1. The results demonstrate the applicability of the vacuum impregnation process for improving the performance of alternative electrodes.

Conclusions

Analysis of the CV response from TiO2 at different scan rates suggested the contribution of a surface controlled charge storage mechanism. Enhancing this surface contribution was achieved through the application of a vacuum impregnation technique to achieve good electrode wetting and improve electrode-electrolyte contact. This vacuum impregnation step allowed a 1.5 cm × 2 cm, 6.5 mg cm− 2 TiO2 electrode to maintain a capacity of 15 mA h g− 1 at the very high specific current of 40 A g− 1 with potential hysteresis between charge and discharge being only 200 mV. A 25% drop in capacity over an order of magnitude increase in specific current adds further evidence to the presence of a surface controlled or capacitive charge storage mechanism from the TiO2 electrode. The results demonstrate the considerable performance improvements possible from this simple vacuum impregnation technique.

Availability of data and materials

Supporting data and background material can be found in the accompanying supplementary material document.

References

  1. Kim H, et al. Aqueous rechargeable li and Na ion batteries. Chem Rev. 2014;114(23):11788–827.

    Article  Google Scholar 

  2. Kim H, et al. Oxygen vacancies enhance pseudocapacitive charge storage properties of MoO3−x. Nat Mater. 2017;16(4):454.

    Article  Google Scholar 

  3. Xu J, et al. ACS applied materials & interfaces, 2019; boosting the stable Na storage performance in 1D Oxysulfide. Adv Energy Mater. 2019;9(20):1900170.

    Article  Google Scholar 

  4. Song W, et al. Pseudo-capacitive Na+ Insertion in Ti-OC Channels of TiO2-C Nanofibers with High-rate and Ultra-stable Performance. ACS Appl Mater Interfaces. 2019;11(19):17416–24.

    Article  Google Scholar 

  5. Zhang BH, et al. Nanowire Na0.35MnO2 from a hydrothermal method as a cathode material for aqueous asymmetric supercapacitors. J Power Sources. 2014;253:98–103.

    Article  Google Scholar 

  6. Hou Z, et al. An aqueous rechargeable sodium ion battery based on a NaMnO2–NaTi2(PO4)3 hybrid system for stationary energy storage. J Mater Chem A. 2015;3(4):1400–4.

    Article  Google Scholar 

  7. Kumar A, et al. An efficient α-MnO2 nanorods forests electrode for electrochemical capacitors with neutral aqueous electrolytes. Electrochim Acta. 2016;220:712–20.

    Article  Google Scholar 

  8. Long JW, et al. Voltammetric characterization of ruthenium oxide-based aerogels and other RuO2 solids: the nature of capacitance in nanostructured materials. Langmuir. 1999;15(3):780–5.

    Article  Google Scholar 

  9. Wang W, et al. Hydrous ruthenium oxide nanoparticles anchored to Graphene and carbon nanotube hybrid foam for Supercapacitors. Sci Rep. 2014;4:4452.

    Article  Google Scholar 

  10. Mohajernia S, et al. Semimetallic core-shell TiO2 nanotubes as a high conductivity scaffold and use in efficient 3D-RuO2 supercapacitors. Mater Today Energy. 2017;6:46–52.

    Article  Google Scholar 

  11. Tang Z, Tang C-h, Gong H. A high energy density asymmetric Supercapacitor from Nano-architectured Ni (OH)2/carbon nanotube electrodes. Adv Funct Mater. 2012;22(6):1272–8.

    Article  Google Scholar 

  12. Tang Y, et al. Morphology controlled synthesis of monodisperse cobalt hydroxide for supercapacitor with high performance and long cycle life. J Power Sources. 2014;256:160–9.

    Article  Google Scholar 

  13. Meher SK, Rao GR. Ultralayered Co3O4 for high-performance Supercapacitor applications. J Phys Chem C. 2011;115(31):15646–54.

    Article  Google Scholar 

  14. Wessells CD, Huggins RA, Cui Y. Copper hexacyanoferrate battery electrodes with long cycle life and high power. Nat Commun. 2011;2:550.

    Article  Google Scholar 

  15. Wessells CD, et al. Nickel Hexacyanoferrate nanoparticle electrodes for aqueous sodium and potassium ion batteries. Nano Lett. 2011;11(12):5421–5.

    Article  Google Scholar 

  16. Shao Y, et al. Design and mechanisms of asymmetric Supercapacitors. Chem Rev. 2018;118(18):9233–80.

    Article  Google Scholar 

  17. Liu S, et al. Aluminum storage behavior of anatase TiO2 nanotube arrays in aqueous solution for aluminum ion batteries. Energy Environ Sci. 2012;5(12):9743–6.

    Article  Google Scholar 

  18. Liu Y, et al. The electrochemical behavior of cl− assisted Al3+ insertion into titanium dioxide nanotube arrays in aqueous solution for aluminum ion batteries. Electrochim Acta. 2014;143:340–6.

    Article  Google Scholar 

  19. He YJ, et al. Black mesoporous anatase TiO2 nanoleaves: a high capacity and high rate anode for aqueous Al-ion batteries. J Mater Chem A. 2014;2(6):1721–31.

    Article  Google Scholar 

  20. Kazazi M, Abdollahi P, Mirzaei-Moghadam M. High surface area TiO2 nanospheres as a high-rate anode material for aqueous aluminium-ion batteries. Solid State Ionics. 2017;300:32–7.

    Article  Google Scholar 

  21. Holland A, et al. An aluminium battery operating with an aqueous electrolyte. J Appl Electrochem. 2018;48(3):243–50.

    Article  Google Scholar 

  22. Liang C, et al. Aqueous batteries based on mixed Monovalence metal ions: a new battery family. ChemSusChem. 2014;7(8):2295–302.

    Article  Google Scholar 

  23. Yao H-R, et al. Rechargeable dual-metal-ion batteries for advanced energy storage. Phys Chem Chem Phys. 2016;18(14):9326–33.

    Article  Google Scholar 

  24. Lu K, et al. A rechargeable Na-Zn hybrid aqueous battery fabricated with nickel hexacyanoferrate and nanostructured zinc. J Power Sources. 2016;321:257–63.

    Article  Google Scholar 

  25. Holland AW, et al. TiO2 nanopowder as a high rate, long cycle life electrode in aqueous aluminium electrolyte. Mater Today Energy. 2018;10:208–13.

    Article  Google Scholar 

  26. Yong S, Owen J, Beeby S. Solid-state Supercapacitor fabricated in a single woven textile layer for E-textiles applications. Adv Eng Mater. 2018;20(5):1700860.

    Article  Google Scholar 

  27. Augustyn V, et al. High-rate electrochemical energy storage through li+ intercalation pseudocapacitance. Nat Mater. 2013;12:518.

    Article  Google Scholar 

  28. Lindström H, et al. Li+ ion insertion in TiO2 (Anatase). 2. Voltammetry on Nanoporous films. J Phys Chem B. 1997;101(39):7717–22.

    Article  Google Scholar 

  29. Augustyn V, Simon P, Dunn B. Pseudocapacitive oxide materials for high-rate electrochemical energy storage. Energy Environ Sci. 2014;7(5):1597–614.

    Article  Google Scholar 

  30. Yoon S-B, Kim K-B. Effect of poly (3,4-ethylenedioxythiophene) (PEDOT) on the pseudocapacitive properties of manganese oxide (MnO2) in the PEDOT/MnO2/multiwall carbon nanotube (MWNT) composite. Electrochim Acta. 2013;106:135–42.

    Article  Google Scholar 

  31. Xue Q, et al. LaB6 nanowires for supercapacitors. Mater Today Energy. 2018;10:28–33.

    Article  Google Scholar 

  32. Xu J, et al. Capacitive lithium storage of lithiated mesoporous titania. Mater Today Energy. 2018;9:240–6.

    Article  Google Scholar 

  33. Pazhamalai P, et al. Fabrication of high energy li-ion hybrid capacitor using manganese hexacyanoferrate nanocubes and graphene electrodes. J Ind Eng Chem. 2018;64:134–42.

    Article  Google Scholar 

  34. Lahan H, et al. Anatase TiO2 as an anode material for rechargeable aqueous aluminum-ion batteries: remarkable Graphene induced aluminum ion storage phenomenon. J Phys Chem C. 2017;121(47):26241–9.

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to acknowledge the financial support received through the European Union’s Horizon 2020 research and innovation programme under grant agreement No 646286.

Funding

This project has received funding from the European Union’s Horizon 2020 research and innovation programme under grant agreement No 646286. This contribution was financial only and did not include design of the study or collection, analysis, and interpretation of data or writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

AWH (first author, electrochemical analysis), AC (co-supervisor, editing), AZ (BET analysis), AH (experimental design) and RGAW (corresponding author, supervisor, grant holder, electrochemical analysis). All authors have read and approved the final manuscript.

Authors’ information

Not applicable.

Corresponding author

Correspondence to R. G. A. Wills.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Additional file 1.

Supplementary materials.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Holland, A.W., Cruden, A., Zerey, A. et al. Electrochemical study of TiO2 in aqueous AlCl3 electrolyte via vacuum impregnation for superior high-rate electrode performance. BMC Energy 1, 10 (2019). https://doi.org/10.1186/s42500-019-0010-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s42500-019-0010-9

Keywords