Next Article in Journal
Molten Bismuth–Bismuth/Zinc Oxide Composites for High-Temperature Thermal Energy Storage
Previous Article in Journal
An Old Material for a New World: Prussian Blue and Its Analogues as Catalysts for Modern Needs
Previous Article in Special Issue
An Electrochemically Prepared Mixed Phase of Cobalt Hydroxide/Oxyhydroxide as a Cathode for Aqueous Zinc Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Investigation of Lithium Perchlorate-Doped Polypyrrole Growing on Titanium Substrate

1
School of Chemistry and Chemical Engineering, Southeast University, Nanjing 211189, China
2
Suzhou Research Institute, Southeast University, Suzhou 215123, China
3
College of Ecology and Resource Engineering, Wuyi University, Wuyishan 354300, China
4
School of Intelligent and Automotive Engineering, Wuxi Vocational Institute of Commerce, Wuxi 214153, China
5
Institute of Shandong Non-Metallic Materials, Jinan 250031, China
*
Author to whom correspondence should be addressed.
Inorganics 2024, 12(4), 125; https://doi.org/10.3390/inorganics12040125
Submission received: 5 March 2024 / Revised: 17 April 2024 / Accepted: 19 April 2024 / Published: 22 April 2024

Abstract

:
Lithium perchlorate-doped polypyrrole growing on titanium substrate (LiClO4-PPy/Ti) has been fabricated to act as electroactive electrode material for feasible electrochemical energy storage. A theoretical and experimental investigation is adopted to disclose the conductivity, electroactivity properties and interfacial interaction-dependent capacitance of LiClO4-PPy/Ti electrode. The experimental measurement results disclose that LiClO4-PPy/Ti reveals lower ohmic resistance (0.2226 Ω cm−2) and charge transfer resistance (2116 Ω cm−2) to exhibit higher electrochemical conductivity, a more reactive surface, and feasible ion diffusion to present higher double-layer capacitance (0.1930 mF cm−2) rather than LiClO4/Ti (0.3660 Ω cm−2, 65,250 Ω cm−2, 0.0334 mF cm−2). LiClO4-PPy/Ti reveals higher Faradaic capacitance caused by the reversible doping and dedoping process of perchlorate ion on PPy than the electrical double-layer capacitance of LiClO4/Ti caused by the reversible adsorption and desorption process of the LiClO4 electrolyte on Ti. Theoretical simulation calculation results prove that a more intensive electrostatic interaction of pyrrole N···Ti (2.450 Å) in LiClO4-PPy/Ti rather than perchlorate O···Ti (3.537 Å) in LiClO4/Ti. LiClO4-PPy/Ti exhibits higher density of states (57.321 electrons/eV) at Fermi energy and lower HOMO-LUMO molecule orbital energy gap (0.032 eV) than LiClO4/Ti (9.652 electrons/eV, 0.340 eV) to present the enhanced electronic conductivity. LiClO4-PPy/Ti also exhibits a more declined interface energy (−1.461 × 104) than LiClO4/Ti (−5.202 × 103 eV) to present the intensified interfacial interaction. LiClO4-PPy/Ti accordingly exhibits much higher specific capacitances of 0.123~0.0122 mF cm−2 at current densities of 0.01~0.10 mA cm−2 rather than LiClO4/Ti (0.010~0.0095 mF cm−2, presenting superior electroactivity and electrochemical capacitance properties. LiClO4-PPy/Ti could well act as the electroactive supercapacitor electrode for feasible energy storage.

1. Introduction

The electrochemical behaviors and performances of energy storage devices are related to the electrode and electrolyte materials as well as their interfacial interaction [1,2,3]. In addition, the electrode substrate materials are also related to the properties of functional electrodes. The electrochemical capacity and mechanical stability are most concerned for the electrodes of energy storage devices. These electrode materials mainly involve the various graphite carbon, conductive polymers, transition metal sulfates or oxides, and their composites [4,5,6,7]. In particular, the conductive polymers become suitable supercapacitor electrode materials for flexible energy storage [8,9]. The electrical conductive materials, such as carbon fiber, graphene, graphite carbon, titanium, and so on, have been used to prepare the energy storage electrodes [10,11,12]. They could act as the multifunctional roles, such as the electrode substrate, current collector, and electrode. However, the contradictory behaviors of these substrate materials also emerged in respect to the electrochemical activity and electrochemical stability. For example, titanium metal could resist the chemical and electrochemical corrosion in neutral, alkaline, and even acidic conditions [13]. Titanium keeps the chemical and electrochemical stability in a certain degree, and is applied to electrode material as well as electrode-supporting material [14,15,16,17]. The structure-adjusted titanium, such as titanium sheet, titanium mesh, and titanium foam with high electrochemical stability, can be used to support various electroactive materials [18,19,20,21]. On the other hand, titanium metal also behaves the relatively low electrochemical activity due to the passivation layer of titanium oxides on the surface [22]. Surface modification using conductive polymers is a useful method to Improve the electrochemical activity of titanium. The Ti metal sheet can act as the electrode substrate. The conductive polymers usually includes polyaniline, polypyrrole, and polythiophene, which have been used to act as electroactive materials for the supercapacitor electrode application [23]. The different electroactive materials can be coated on Ti substrate to form functional electrodes. The polypyrrole (pPy) has been electrodeposited on Ti substrate to form pPy/Ti electrode through cyclic voltammetry polymerization process [24]. The polyaniline (PANI) has been electrodeposited and coated on Ti substrate to form PANI/Ti electrode through a pulse potentiostatic polymerization process [25]. These conductive polymers could conduct the reversible ion doping–dedoping process to provide superior capacitance [26,27,28]. The interfacial bonding strength between the doped conductive polymers and metal substrates is related to the electrochemical energy storage performance [29]. In addition, the reactive electrolytes also influence the electrical double-layer capacitance due to the interfacial adsorption on metal substrates and the Faradaic capacitance due to the ion doping state of conductive polymers. In previous studies, the system of a lithium perchlorate electrolyte and polypyrrole has been attempted for various applications in different areas [30,31]. However, the interfacial interaction between lithium perchlorate doping polypyrrole and metal titanium substrate is still sparsely investigated [32].
This study focuses on disclosing the promotion effect of electroactive polypyrrole on the electrochemical properties of lithium perchlorate interacting with titanium substrate. The polypyrrole is used as the typical conductive polymer material, and a metal titanium sheet is used as the stable substrate material. The titanium sheet has chemical and electrochemical stability to resist the electrochemical corrosion to act as a suitable current collector. The polypyrrole growing on titanium substrate (PPy/Ti) is designed as the supercapacitor electrode in the application of electrochemical energy storage, which is prepared through the electrochemical polymerization process. The lithium perchlorate (LiClO4) is used as a neural electrolyte to study the electrical double-layer capacitance of titanium metal substrate and the Faradaic capacitance of polypyrrole. The interfacial interaction is intensively considered between the LiClO4 electrolyte and the PPy/Ti electrode when the electrochemical behavior is investigated for LiClO4-Ppy/Ti. The electrochemical performances are fully evaluated through cyclic voltammetry, electrochemical impedance spectrum, and galvanostatic charge/discharge measurements. The theoretical simulation calculation has been applied to disclose the atomic charge distribution, electrostatic potential, and interfacial affinity of LiClO4-PPy/Ti, which could lead to a better understanding of capacitance performance.

2. Results and Discussion

2.1. Microstructure and Morphology Characterization

Figure 1A–D shows SEM images of Ti and PPy/Ti. The Ti sheet substrate reveals the uniquely complete surface structure without any crack. The chemical polishing treatment could lead to forming the uniform metal surface. Comparatively, PPy/Ti revels the compact covering layer of PPy film supported on the Ti sheet. The PPy also is composed of the aggregating particles, forming uniquely complete film structure without any crack. Concerning the surface microstructure, the Ti sheet shows smooth surface and PPy/Ti shows rough surface. The insert in Figure 1D shows the cross-section SEM image of PPy/Ti. The PPy covered on the surface of the Ti substrate exhibits a close-packing film. The morphology, microstructure, and thickness of such PPy film could be adjusted by modulating the electropolymerization reaction parameters in the CV electrodeposition process. This study focuses on the interfacial interaction between PPy film and Ti substrate, which is related to the microstructure of PPy. It is believed that PPy/Ti could conduct more feasible ion diffusion and doping reaction in the presence of lithium perchlorate in comparison the ion adsorption of PPy/Ti.
Figure 2A shows the XRD pattern of the titanium sheet substrate. Concerning the titanium sheet, the distinct characteristic diffraction peaks at 2θ of 35.1°, 38.5°, 40.3°, 53.1°, 63.1°, 70.8°, 76.3°, 77.5°, and 82.3° correspond to the respective crystal planes of (100), (002), (101), (102), (110), (103), (112), (201), and (202) [33]. The titanium sheet involves the close-packed hexagonal crystal phase structure, which is acted as stable metal substrate. Figure 2B shows the Raman spectrum of the Ti substrate. The Ti substrate does not reveal any characteristic Raman peaks in full wavenumber range due to the absence of electronic transition-relaxation of crystal lattice Ti atoms at 785 nm excitation photon energy. Figure 2C shows the Raman spectrum of PPy/Ti synthesized in this study. The characteristic Raman peaks of PPy/Ti are observed at 983 and 1050 cm−1, with a broad band of 1340–1410 cm−1 and 1580 cm−1, which are assigned to C–H ring deformation vibration, C–N stretching vibration, and C=C aromatic stretching vibration, respectively. The characteristic Raman peak position of the literature-reported PPy is located at 980, 1040, 1314–1410 cm−1, and 1597 cm−1 [34]. The characteristic Raman peaks of PPy are well-matched between this study-formed PPy/Ti and the literature-reported PPy. Therefore, the Raman spectrum analysis result could well confirm the formation of PPy in the as-prepared PPy/Ti electrode in this study. Figure 2D shows the FTIR spectrum of PPy/Ti. The characteristic peaks at 1540, 1480, 1190, and 1040 cm−1 were ascribed to C=C double bond stretching vibration, C-C single bond stretching vibration, C-N single bond stretching vibration, and N-H in-plane bending vibration. The characteristic peaks at 912, 768, 742, and 675 cm−1 are ascribed to the C-H out-of-plane bending vibration. The observation of the C=C double bond, C-C single bond, and C-N single bond could be ascribed to the formation of PPy. Accordingly, PPy fully grew on the Ti substate to form PPy/Ti through the electrochemical polymerization and deposition process.

2.2. Electrochemical Properties

The CV curves can be used to calculate the CV-based capacitance (CCV, mF cm−2) using Equation (1). The GCD curves can be used to calculate the GCD-based capacitance (CGCD, mF cm−2) using Equation (2).
C CV = i / v = [ ( V c V a i ( ν ) d ν ) / ( V a V c ) ] / ν = [ 1 2 ( i ( ν ) d ν ) / Δ V ] / ν
C G C D = I t V
where: Va, upper potential; Vc, lower potential; i(v), response current density (mA cm−2); v, potential sweep rate (V s−1); I, discharge current density (mA cm−2); t, discharge time (s); and ΔV, potential window (V).
Figure 3A,B shows the CV curves of LiClO4/Ti and LiClO4-PPy/Ti at scan rates of 5~200 mV s−1 and the potential range of 0~0.8 V versus saturated calomel electrode (SCE) in 1.0 M LiClO4 electrolyte solution. The CV curves of LiClO4/Ti demonstrate the rectangle-like shape to present electrical double-layer capacitance due to the reversible adsorption–desorption of LiClO4 electrolyte ion on the Ti substrate. The oxidation potential of Ti2+/Ti is −1.39 V versus SCE. This means that the electrochemical oxidation of Ti is unlikely to occur in the potential range of 0~0.8 V versus SCE, presenting the high electrochemical stability of the Ti substrate in LiClO4 electrolyte. Comparatively, the CV curves of LiClO4-PPy/Ti exhibit an obviously deviated rectangle shape in the potential range of 0~0.8 V, presenting the polarization effect. The redox potential of PPy is around 0.5 V versus SCE [35]. LiClO4-PPy conducts reversible ClO4 doping-dedoping reaction which contributes to the Faradaic capacitance performance.
Figure 3C,D shows CV-based capacitance curves of LiClO4/Ti and LiClO4-PPy/Ti. When the CV cures of LiClO4/Ti and LiClO4-PPy/Ti conduct the scan rate increasing from 5 to 200 mV s−1, their mean response current increases from 0.13 to 2.47 μA cm−2 and from 2.28 to 9.011 μA cm−2, respectively. The CV-based capacitance decreases from 0.0256 to 0.0124 mF cm−2 and from 0.456–0.045 mF cm−2, respectively. In comparison with LiClO4/Ti, LiClO4-PPy/Ti exhibits a much higher mean response current and specific capacitance performance at the same scan rate, presenting its superior electrochemical activity in the neutral LiClO4 electrolyte. The oxidation potential of Ti2+/Ti is 1.39 V versus SCE (Ti − 2e → Ti2+). The oxidation potential of PPy at the doping process is approximate 0.5 V versus SCE (PPy − e + ClO4 → PPy+ClO4). PPy/Ti could conduct reversible doping–dedoping process of PPy rather than the electrochemical oxidation of the Ti substrate in the potential range of 0~0.8 V. Additionally, the PPy covering layer could also resist such electrochemical corrosion reaction of Ti substrate. Meanwhile, PPy could conduct the redox reaction through the reversible doping–dedoping ClO4 anion, contributing to improving the electrochemical activity and, accordingly, the Faradaic capacitance performance.
Figure 4A,B shows the GCD curves of LiClO4/Ti and LiClO4-PPy/Ti in the potential range of 0~0.8 V (vs. SCE) and 1.0 M LiClO4 electrolyte solution. The GCD curves of LiClO4/Ti show complete symmetric characteristics. This shows the good reversible properties of LiClO4 ion adsorption–desorption during the charge–discharge process. Such LiClO4/Ti exhibits ideal electrical double-layer capacitance behavior. Comparatively, the GCD curves of LiClO4-PPy/Ti shows the sloping plateau. The less-symmetric characteristic is ascribed to the redox reaction in the charge–discharge process. Such LiClO4-PPy/Ti exhibits Faradaic capacitance behavior. Therefore, LiClO4-PPy/Ti demonstrates much higher capacitance performance than LiClO4/Ti in the neutral LiClO4 electrolyte. Furthermore, LiClO4-PPy/Ti conducts the doping–dedoping reaction of PPy, and LiClO4/Ti conducts the ion adsorption–desorption process of Ti. The GCD curve shape of LiClO4-PPy/Ti shows fewer identical characteristics than that of LiClO4/Ti. Figure 4C,D shows GCD-based capacitance curves of LiClO4/Ti and LiClO4-PPy/Ti. According to the GCD curves at current densities of 0.01~0.10 mA cm−2, the corresponding GCD-based specific capacitances achieve 0.010~0.0095 mF cm−2 and 0.123~0.0122 mF cm−2. Therefore, LiClO4-PPy/Ti exhibits much higher capacitance than LiClO4/Ti, since electroactive LiClO4-PPy mostly contributes to Faradaic capacitance rather than electrical double layer capacitance of Ti substrate in LiClO4 electrolyte. These GCD measurement results agree with the CV measurement results. Furthermore, this LiClO4-PPy/Ti electrode keeps the similar capacitance performance in comparison with the ion doped PPy reported in the literature [36,37]. Figure 4E shows the cycling capacity retention curve of LiClO4-PPy/Ti. The capacity retention achieves 76.1% after 1000 cycles during the cycling charge–discharge process. Therefore, the PPy/Ti electrode exhibits the reasonable cycling stability in the electrochemical Faradaic reaction process.
An electrochemical impedance spectroscopy (EIS) is adopted to investigate the electrolyte ion diffusion and interfacial charge transfer properties of electrode materials. Figure 5A,B shows the Nyquist plots of the EIS spectra of LiClO4/Ti and LiClO4-PPy/Ti. Figure 5C shows equivalent circuit model for fitting curves of EIS spectra. The equivalent circuit includes charge-transfer resistance (Rct), ohmic resistance (Ro), double layer capacitance (Cdl), constant phase element (CPE), and CPE exponent (n). n = 1 indicates ideal capacitor behavior, and n = 0 indicates pure resistor behavior. Table 1 lists the fitting parameter values of these elements used in an equivalent circuit model. In comparison with LiClO4/Ti, the slightly lowered Ro value of LiClO4-PPy/Ti indicates higher electronic conductivity. The ohm resistance of LiClO4-PPy/Ti is mostly dependent on the Ti substrate. The conducting LiClO4-PPy layer prevents the formation of semi-conducting titanium oxide. The quite lowered Rct value indicates the more electroactive interface of LiClO4-PPy for the charge transfer. Its higher Cdl value indicates higher double-layer capacitance due to the rougher surface of PPy/Ti rather than Ti substate. However, its higher CPE value indicates the higher Warburg impedance in view of electrolyte diffusion. The Warburg impedance of Ti substrate keeps relatively low value. The lithium ion has relatively high diffusion coefficient and plays a more predominant role than the electrode materials in the LiClO4 electrolyte. LiClO4/Ti carries out a shorter ion diffusion length than LiClO4-PPy/Ti. Therefore, LiClO4/Ti reveals a relatively lower CPE value than LiClO4-PPy/Ti. Moreover, LiClO4-PPy/Ti exhibits a larger CPE exponent (n value) than LiClO4/Ti. This indicates that LiClO4-PPy/Ti more likely approaches capacitor behavior rather than resistance behavior. The LiClO4-PPy layer provides more active interface for electrolyte ion adsorption and charge transfer. Therefore, LiClO4-PPy/Ti exhibits higher electroactivity than LiClO4/Ti, contributing to a higher specific capacitance in neutral LiClO4 electrolyte solution.

2.3. DFT Simulation Calculation of LiClO4/Ti and LiClO4-PPy/Ti

The LiClO4 molecule conducts the adsorption–desorption to establish interfacial interaction between LiClO4 and Ti substrate through Van der Waals forces in the electrochemical process. The LiClO4 molecule also conducts the doping–dedoping process to establish electrostatic interaction between LiClO4-PPy and Ti substrate. The atomic charge distribution could affect the interaction of LiClO4/Ti and LiClO4-PPy/Ti. The DFT simulation calculation is used to investigate electronic properties of LiClO4/Ti and LiClO4-PPy/Ti [38].
Figure 6A,B shows molecule the structure models of LiClO4/Ti and LiClO4-PPy/Ti. LiClO4/Ti involves an interfacial electrostatic interaction between the LiClO4 and Ti substrate. LiClO4-PPy/Ti involves an interfacial electrostatic interaction between the LiClO4-PPy and Ti substrate. LiClO4-PPy presents the perchlorate ion doping state of PPy.
Figure 6C,D shows the molecular electrostatic potentials of LiClO4/Ti and LiClO4-PPy/Ti. Concerning LiClO4/Ti, the negative charge of ClO4 ion could induce the charge dissociation of the Ti substrate, which leads to forming positive and negative crystal lattice Ti atoms. Concerning LiClO4-PPy/Ti, the doping state of LiClO4-PPy reveals a positive charge of hydrogen atoms and negative charge of conjugated pyrrole rings, which leads to the forming of a positive and negative crystal lattice of Ti atoms. The intermolecular interaction distance is determined to be 3.537 Å for perchlorate O…Ti of LiClO4/Ti and 2.450 Å for pyrrole N…Ti of LiClO4-PPy/Ti, respectively. Therefore, LiClO4-PPy/Ti presents stronger electrostatic interaction than LiClO4/Ti. The electrostatic interaction is allowed to occur for LiClO4/Ti and LiClO4-PPy/Ti. The intensified interaction has been established between the LiClO4-PPy molecule chain and the Ti substrate, contributing to the activation of the LiClO4-PPy/Ti electrode.
Figure 6E,F shows the density of states curves of LiClO4/Ti and LiClO4-PPy/Ti. Both LiClO4/Ti and LiClO4-PPy/Ti demonstrate metal conductivity characteristics, since the characteristic density of states at Fermi energy 0 eV cross from the valance band to the conduction band. Comparatively, LiClO4-PPy/Ti demonstrates more orbit energy levels than the LiClO4/Ti at valence band and conduction band regions, which are ascribed to the interfacial interaction between the LiClO4-PPy and the Ti substrate. Furthermore, LiClO4-PPy/Ti shows a higher DOS level (57.321 electrons/eV) than LiClO4/Ti (9.652 electrons/eV) at Fermi energy 0 eV. LiClO4-PPy/Ti exhibits higher electron states and higher electronic conductivity.
Figure 6G,H shows the band gap structure curves of LiClO4/Ti and LiClO4-PPy/Ti. The band gaps of LiClO4/Ti and LiClO4-PPy/Ti approach 0.03 and 0 eV, presenting similar superior conductivity of metal substrates. Comparatively, LiClO4-PPy/Ti demonstrates much greater energy levels at the valance band and conduction band regions than LiClO4/Ti, which presents superior electrical conductivity of LiClO4-PPy/Ti. Therefore, the band gap analysis results agree with the DOS analysis results, proving the superior electronic conductivity of LiClO4-PPy/Ti rather than LiClO4/Ti. The declined interfacial energies are −5.202 × 103 eV for LiClO4/Ti and −1.461 × 104 for LiClO4-PPy/Ti, which presents the intensified interfacial interaction of LiClO4-PPy/Ti. Therefore, LiClO4-PPy/Ti shows higher electroactivity than LiClO4/Ti. LiClO4-PPy has more electronic distribution than LiClO4-PPy/Ti, which leads to stronger interaction, lower interfacial energy, and, accordingly, higher capacitance performance.
Figure 6I shows the HOMO and LUMO HOMO and LUMO frontier molecular orbital charge distribution of LiClO4/Ti and LiClO4-PPy/Ti. Concerning HOMO and LUMO electron distribution in LiClO4/Ti and LiClO4-PPy/Ti, their molecular orbitals are mostly dependent on Ti rather than LiClO4 and LiClO4-PPy. In Ti atoms, there occurs an obvious electronic transition, which causes heterogeneous electron distribution. Comparatively, different O and Cl atoms in LiClO4 occur insignificant electronic transition due to weak electron transfer. Similarly, the electron transition of different atoms in LiClO4-PPy also becomes relatively neglective because of weak electronic transfer. The HOMO-LUMO molecule orbital energy gaps of LiClO4/Ti and LiClO4-PPy/Ti are determined to be 0.340 eV and 0.032 eV, respectively. Therefore, LiClO4-PPy/Ti with a low HOMO-LUMO gap presents its high conductivity and high electroactivity, also. The HOMO-LUMO gap analysis results agree with the DOS and band gap analysis results. The DFT simulation calculation results support the CV and GCD measurement results, proving the higher capacitance of LiClO4-PPy/Ti.
Figure 7 shows the Mulliken charge population of C, O, N, Cl, and Ti atoms of LiClO4/Ti and LiClO4-PPy/Ti. The interfacial interaction in LiClO4/Ti and LiClO4-PPy/Ti leads to the center charge offset of the Ti lattice atoms and the negative charge density distribution of oxygen atoms as well as nitrogen atoms. Accordingly, Ti atoms involve the dissociation distribution of negative and positive charge in LiClO4/Ti and LiClO4-PPy/Ti. Table 2 lists the Mulliken charges of C, O, N, Cl, and Ti atoms in LiClO4/Ti and LiClO4-PPy/Ti. The negative charge densities of oxygen atoms are changed from LiClO4/Ti (−0.78, −0.80, −0.82, −0.95) to LiClO4-PPy/Ti (−0.76, −0.77, −0.78, −0.90), indicating a slightly decreased negative charge. The dissociated positive and negative charge densities of Ti substrate are (+0.13~+0.05)/(−0.23~−0.04) for LiClO4/Ti and (+0.62~+0.08)/(−0.05~−0.01) for LiClO4-PPy/Ti, respectively. Therefore, the electrostatic interaction is formed between negative-charged perchlorate O and positive-charged Ti. Additionally, the negative charge densities of nitrogen atom are (−0.58~−0.47). Obviously, the strong electrostatic interaction is further formed between the negative-charged pyrrole N and the positive-charged Ti in LiClO4-PPy/Ti. Herein, LiClO4 exhibits the surface adsorption state of Ti in LiClO4/Ti and the doping state of PPy in LiClO4-PPy/Ti. The Mulliken charge population analysis proves the shorter intermolecular interaction distance in LiClO4-PPy/Ti in comparison with LiClO4/Ti. Accordingly, LiClO4-PPy/Ti achieves higher electroactivity than LiClO4/Ti. The DFT simulation calculation results agree with the experimental measurement results in confirming the superior capacitance of LiClO4-PPy/Ti.

3. Experimental Section

3.1. Preparation of LiClO4-PPy/Ti

The titanium sheet (Ti, purity 99.5%, size 10 × 50 mm, thickness 0.2 mm.) was washed using acetone, ethanol, and then deionized water to remove surface stains. The Ti sheet then conducted the chemical polishing treatment in HF-HNO3 solution to dissolve the surface-covered oxide layer. The chemical cleaned Ti sheet was obtained. The Ti sheet conducted the electrostatic adsorption in lithium perchlorate solution form LiClO4/Ti. A normal pulse voltammetry process was employed to synthetize LiClO4-PPy/Ti using a three-electrode reaction system, which included the fresh Ti sheet working electrode, Pt foil counter electrode, the Hg/Hg2Cl2 reference electrode, and acetonitrile solvent solution of 0.1 M LiClO4 electrolyte and 0.15 M pyrrole. The reaction parameters included a pulse potential range of 0.7~1.1 V and pulse potential increment of 0.001 V.

3.2. Characterization and Measurement

The surface microstructure characterization of LiClO4-PPy/Ti and LiClO4/Ti was conducted using a scanning electron microscope (SEM, ZEISS Ultra Plus, Germany). A Raman spectrum measurement of LiClO4-PPy/Ti was conducted using a Raman spectroscope (Thermo Fisher DXR3, He-Ne laser, Emitting wavelength of 785 nm). A crystal phase structure characterization of titanium substrate was conducted using an X-ray diffraction (XRD, Bruker D8-Discover, Wavelength of X-ray source, 0.154 nm) apparatus. Electrochemical capacitive measurements of LiClO4-PPy/Ti and LiClO4/Ti were performed using an electrochemical workstation (CHI 760D, CH Instruments Co., Ltd., China), which includes galvanostatic charge/discharge (GCD) and cyclic voltammetry (CV). The electrochemical impedance spectrum (EIS) measurement was carried out at the frequency range of 10−2~105 Hz. The 1.0 M LiClO4 was used as the doping ion source and the electrolyte solution, contributing to maintaining the electrochemical stability of the titanium substrate in LiClO4-PPy/Ti. electrode. The total surface area was used to evaluate electrochemical capacitance.

3.3. Theoretical Simulation Calculation

Density functional theory (DFT) simulation calculation was carried out to investigate electronic properties and atomic charge distribution of LiClO4 and LiClO4-PPy interaction at the interface of Ti substrate [38]. The total density of states (DOS), Mulliken charge population, interfacial binding energy, highest occupied molecular orbital (HOMO), lowest unoccupied molecular orbital (LUMO), and electrostatic potential were determined and discussed for better understanding of the electrostatic interfacial interaction of LiClO4-PPy/Ti and LiClO4/Ti. The DMol3 package was applied to calculate HOMO-LUMO energy gaps, and the CASTEP package was applied to calculate electronic energy gap and charge density difference and DOS value.

4. Conclusions

The PPy/Ti has been prepared through an electro-polymerization deposition process which is applied as supercapacitor electrode for electrochemical energy storage. Theoretical and experimental investigations are adopted to disclose the conductivity and interfacial interaction-dependent capacitance of LiClO4-PPy/Ti. LiClO4/Ti reveals ideal electrical double-layer capacitance due to the reversible adsorption/desorption of LiClO4 electrolyte on Ti substrate in the potential range of 0~0.8 V. LiClO4-PPy/Ti reveals Faradaic capacitance due to reversible doping/dedoping perchlorate ion in PPy. LiClO4-PPy/Ti reveals lower charge transfer resistance, lower ohmic resistance, and lower Warburg resistance to present more active interface, higher electronic conductivity, and more feasible ion diffusion properties than LiClO4/Ti. LiClO4/Ti and LiClO4-PPy/Ti achieve the mean response current of 0.13~2.47 μA cm−2 and 2.28~9.011 μA cm−2 at 5~200 mV s−1. The corresponding specific capacitances are 0.010~0.0095 mF cm−2 and 0.123~0.0122 mF cm−2 at 0.01~0.10 mA cm−2. LiClO4-PPy/Ti keeps much higher electrochemical conductivity and higher specific capacitance than LiClO4/Ti. Theoretical calculation results prove that interfacial electrostatic adsorption of LiClO4/Ti and ClO4 anion doping of LiClO4-PPy/Ti could change atomic charge density distribution. LiClO4-PPy/Ti exhibits higher electronic conductivity and lower interface energy than LiClO4/Ti. LiClO4-PPy/Ti can act as a superior supercapacitor electrode to exhibit electroactivity and capacitance for the promising energy storage application.

Author Contributions

Formal analysis, Y.X. and J.X.; data curation, Y.X., L.L. and C.X.; writing—original draft preparation, Y.X.; writing—review and editing, Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

The Science and Technology Program of Suzhou City, China (SYG202342).

Data Availability Statement

Data are contained within the article.

Acknowledgments

This research work is supported by Science and Technology Program of Suzhou City, China (SYG202342), and Big Data Computing Center of Southeast University, China.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chen, G.Z. Understanding supercapacitors based on nano-hybrid materials with interfacial conjugation. Prog. Natl. Sci. 2013, 23, 245–255. [Google Scholar] [CrossRef]
  2. Adalati, R.; Sharma, M.; Sharma, S.; Kumar, A.; Malik, G.; Boukherroub, R.; Chandra, R. Metal nitrides as efficient electrode material for supercapacitors: A review. J. Energy Stor. 2022, 56, 105912. [Google Scholar] [CrossRef]
  3. Yadav, M.S. Metal oxides nanostructure-based electrode materials for supercapacitor application. J. Nanopart. Res. 2020, 22, 367. [Google Scholar] [CrossRef]
  4. Parveen, N.; Ansari, S.A.; Ansari, M.Z.; Ansari, M.O. Manganese oxide as an effective electrode material for energy storage: A review. Environ. Chem. Lett. 2022, 20, 283–309. [Google Scholar] [CrossRef]
  5. Ansari, M.Z.; Nandi, D.K.; Janicek, P.; Ansari, S.A.; Ramesh, R.; Cheon, T.; Shong, B.; Kim, S.-H. Low-Temperature Atomic Layer Deposition of Highly Conformal Tin Nitride Thin Films for Energy Storage Devices. Acs Appl. Mater. Interface 2019, 11, 43608–43621. [Google Scholar] [CrossRef]
  6. Parveen, N.; Hilal, M.; Han, J.I. Newly Design Porous/Sponge Red Phosphorus@Graphene and Highly Conductive Ni2P Electrode for Asymmetric Solid State Supercapacitive Device with Excellent Performance. Nano-Micro Lett. 2020, 12, 25. [Google Scholar] [CrossRef] [PubMed]
  7. Karnan, M.; Hari Prakash, K.; Badhulika, S. Revealing the super capacitive performance of N-doped hierarchical porous activated carbon in aqueous, ionic liquid, and redox additive electrolytes. J. Energy Stor. 2022, 53, 105189. [Google Scholar] [CrossRef]
  8. Xie, Y. Electrochemical Performance of Polyaniline Support on Electrochemical Activated Carbon Fiber. J. Mater. Eng. Perform. 2022, 31, 1949–1955. [Google Scholar] [CrossRef]
  9. Wang, H.; Xie, Y. Hydrogen bond enforced polyaniline grown on activated carbon fibers substrate for wearable bracelet supercapacitor. J. Energy Stor. 2022, 52, 105042. [Google Scholar] [CrossRef]
  10. Xie, Y. Electrochemical and hydrothermal activation of carbon fiber supercapacitor electrode. Fiber Polym. 2022, 23, 10–17. [Google Scholar] [CrossRef]
  11. Xie, Y.; Wang, Y.; Wang, L.; Liang, J. Theoretical and Experimental Investigations of Oxygen Activation Effect of Carbon Nanofibers Interacting with Polypyrrole. Fibers 2024, 12, 4. [Google Scholar] [CrossRef]
  12. Zhai, Z.; Zhang, L.; Du, T.; Ren, B.; Xu, Y.; Wang, S.; Miao, J.; Liu, Z. A review of carbon materials for supercapacitors. Mater. Des. 2022, 221, 111017. [Google Scholar] [CrossRef]
  13. Jáquez-Muñoz, J.M.; Gaona-Tiburcio, C.; Chacón-Nava, J.; Cabral-Miramontes, J.; Nieves-Mendoza, D.; Maldonado-Bandala, E.; Delgado, A.D.; Flores-De los Rios, J.P.; Bocchetta, P.; Almeraya-Calderón, F. Electrochemical Corrosion of Titanium and Titanium Alloys Anodized in H2SO4 and H3PO4 Solutions. Coatings 2022, 12, 325. [Google Scholar] [CrossRef]
  14. Ansari, S.A.; Khan, N.A.; Hasan, Z.; Shaikh, A.A.; Ferdousi, F.K.; Barai, H.R.; Lopa, N.S.; Rahman, M.M. Electrochemical synthesis of titanium nitride nanoparticles onto titanium foil for electrochemical supercapacitors with ultrafast charge/discharge. Sustain. Energy Fuels 2020, 4, 2480–2490. [Google Scholar] [CrossRef]
  15. Liu, X.Y.; Chen, H.; Li, G.; Peng, J.H.; Zhang, Y.X. One-pot synthesis of pearl-chain-like manganese dioxide-decorated titanium grids as advanced binder-free supercapacitors electrodes. Ceram. Int. 2016, 42, 9227–9233. [Google Scholar] [CrossRef]
  16. Wang, L.; Ma, Y.; Yang, M.; Qi, Y. Titanium plate supported MoS2 nanosheet arrays for supercapacitor application. Appl. Surf. Sci. 2017, 396, 1466–1471. [Google Scholar] [CrossRef]
  17. Wei, J.; Wei, S.; Wang, G.; He, X.; Gao, B.; Zhao, C. PPy modified titanium foam electrode with high performance for supercapacitor. Eur. Polym. J. 2013, 49, 3651–3656. [Google Scholar] [CrossRef]
  18. Zhang, J.; Yu, X.; Zhao, Z.-y.; Zhang, Z.; Li, J. Influence of pore size of Ti substrate on structural and capacitive properties of Ti/boron doped diamond electrode. J. Alloys Compd. 2019, 777, 84–93. [Google Scholar] [CrossRef]
  19. Chang, L.; Chen, B.; Qiao, H.; Huang, H.; Guo, Z.; He, Y.; Xu, R.; Xionghui, X. Study of the Effects of Pretreatment Processing on the Properties of Metal Oxide Coatings on Ti-Based Sheet. J. Electrochem. Soc. 2021, 168, 033501. [Google Scholar] [CrossRef]
  20. Kishimoto, A.; Yamada, Y.; Funatsu, K.; Uda, T. Suitable Electrode Materials for Titanium Sheet Deposition. Adv. Eng. Mater. 2020, 22, 1900747. [Google Scholar] [CrossRef]
  21. Lamberti, A. Flexible supercapacitor electrodes based on MoS2-intercalated rGO membranes on Ti mesh. Mater. Sci. Semicond. Process. 2018, 73, 106–110. [Google Scholar] [CrossRef]
  22. Seo, H.-S.; Bae, J.-U.; Kim, D.-H.; Park, Y.; Kim, C.-D.; Kang, I.B.; Chung, I.-J.; Choi, J.-H.; Myoung, J.-M. Reliable Bottom Gate Amorphous Indium-Gallium-Zinc Oxide Thin-Film Transistors with TiOx Passivation Layer. Electrochem. Solid State Lett. 2009, 12, H348–H351. [Google Scholar] [CrossRef]
  23. Reddy, P.C.H.; Amalraj, J.; Ranganatha, S.; Patil, S.S.; Chandrasekaran, S. A review on effect of conducting polymers on carbon-based electrode materials for electrochemical supercapacitors. Synth. Met. 2023, 298, 117447. [Google Scholar] [CrossRef]
  24. Tsekova, D.S.; Karastoyanov, V.; Peychev, D.; Valova, I. Crystallization of ferritin on biocompatible Surfaces—Bare Ti and Ti covered by polypyrrole (PPy). J. Cryst. Growth 2024, 631, 127616. [Google Scholar] [CrossRef]
  25. Chen, J.; He, Y.; Li, L. Real-time probing electrodeposition growth of polyaniline thin film via in-situ spectroscopic ellipsometry. Thin Solid Films 2022, 762, 139565. [Google Scholar] [CrossRef]
  26. Kondratiev, V.V.; Holze, R. Intrinsically conducting polymers and their combinations with redox-active molecules for rechargeable battery electrodes: An update. Chem. Pap. 2021, 75, 4981–5007. [Google Scholar] [CrossRef]
  27. Stejskal, J.; Sapurina, I.; Vilčáková, J.; Humpolíček, P.; Truong, T.H.; Shishov, M.A.; Trchová, M.; Kopecký, D.; Kolská, Z.; Prokeš, J.; et al. Conducting polypyrrole-coated macroporous melamine sponges: A simple toy or an advanced material? Chem. Pap. 2021, 75, 5035–5055. [Google Scholar] [CrossRef]
  28. Ji, S.; Yang, J.; Cao, J.; Zhao, X.; Mohammed, M.A.; He, P.; Dryfe, R.A.W.; Kinloch, I.A. A Universal Electrolyte Formulation for the Electrodeposition of Pristine Carbon and Polypyrrole Composites for Supercapacitors. Acs Appl. Mater. Interface 2020, 12, 13386–13399. [Google Scholar] [CrossRef] [PubMed]
  29. Zhu, H.; Xie, Y. Electrochemical performance of bridge molecule-reinforced activated carbon fiber-m-aminobenzenesulfonic acid-polyaniline for braidable-supercapacitor application. Chem. Eng. J. 2023, 478, 147416. [Google Scholar] [CrossRef]
  30. Jakhar, P.; Shukla, M.; Singh, V. Influence of LiClO4 Concentration on 1-D Polypyrrole Nanofibers for Enhanced Performance of Glucose Biosensor. J. Electrochem. Soc. 2018, 165, G80–G89. [Google Scholar] [CrossRef]
  31. Santino, L.M.; Acharya, S.; D’Arcy, J.M. Low-temperature vapour phase polymerized polypyrrole nanobrushes for supercapacitors. J. Mater. Chem. A 2017, 5, 11772–11780. [Google Scholar] [CrossRef]
  32. Sharifi-Viand, A.; Mahjani, M.G.; Moshrefi, R.; Jafarian, M. Diffusion through the self-affine surface of polypyrrole film. Vacuum 2015, 114, 17–20. [Google Scholar] [CrossRef]
  33. Wysocki, B.; Maj, P.; Sitek, R.; Buhagiar, J.; Kurzydłowski, K.J.; Święszkowski, W. Laser and Electron Beam Additive Manufacturing Methods of Fabricating Titanium Bone Implants. Appl. Sci. 2017, 7, 657. [Google Scholar] [CrossRef]
  34. Hasoon, S. Electrochemical polymerization and Raman study of polypyrrole and polyaniline thin films HS Abdullah. Int. J. Phys. Sci. 2012, 7, 5468–5476. [Google Scholar]
  35. Fernández Romero, A.J.; López Cascales, J.J.; Fernández Otero, T. Perchlorate Interchange during the Redox Process of PPy/PVS Films in an Acetonitrile Medium. A Voltammetric and EDX Study. J. Phys. Chem. B 2005, 109, 907–914. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, J.; Wu, C.; Wu, P.; Li, X.; Zhang, M.; Zhu, J. Polypyrrole capacitance characteristics with different doping ions and thicknesses. Phys. Chem. Chem. Phys. 2017, 19, 21165–21173. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, S.; Liu, H.; Wang, Y.; Xu, S.; Liu, W.; He, D.; Liu, X.; Liu, J.; Hu, C. Electrochemical Capacitance of Spherical Nanoparticles Formed by Electrodeposition of Intrinsic Polypyrrole onto Au Electrode. Electrochim. Acta 2017, 232, 72–79. [Google Scholar] [CrossRef]
  38. Xie, Y.; Mu, Y. Interface Mo-N coordination bonding MoSxNy@Polyaniline for stable structured supercapacitor electrode. Electrochim. Acta 2021, 391, 138953. [Google Scholar] [CrossRef]
Figure 1. (A) SEM image and (B) magnified SEM image of Ti; (C) SEM image of PPy/Ti; and (D) magnified SEM image and cross-section SEM image of PPy/Ti.
Figure 1. (A) SEM image and (B) magnified SEM image of Ti; (C) SEM image of PPy/Ti; and (D) magnified SEM image and cross-section SEM image of PPy/Ti.
Inorganics 12 00125 g001
Figure 2. (A) XRD pattern and (B) Raman spectrum of Ti substrate; (C) Raman spectrum of PPy/Ti; and (D) FTIR spectrum of PPy/Ti.
Figure 2. (A) XRD pattern and (B) Raman spectrum of Ti substrate; (C) Raman spectrum of PPy/Ti; and (D) FTIR spectrum of PPy/Ti.
Inorganics 12 00125 g002
Figure 3. (A,B) CV curves and (C,D) CV-based capacitance curves of LiClO4/Ti and LiClO4-PPy/Ti at scan rates of 5~200 mV s−1 in 1.0 M LiClO4 electrolyte solution.
Figure 3. (A,B) CV curves and (C,D) CV-based capacitance curves of LiClO4/Ti and LiClO4-PPy/Ti at scan rates of 5~200 mV s−1 in 1.0 M LiClO4 electrolyte solution.
Inorganics 12 00125 g003
Figure 4. (A,B) GCD curves and (C,D) specific capacitance curves of LiClO4/Ti and LiClO4-PPy/Ti at current densities of 0.01~0.1 mA cm−2 in 1.0 M LiClO4 electrolyte solution; and (E) the cycling capacity retention curve of LiClO4-PPy/Ti.
Figure 4. (A,B) GCD curves and (C,D) specific capacitance curves of LiClO4/Ti and LiClO4-PPy/Ti at current densities of 0.01~0.1 mA cm−2 in 1.0 M LiClO4 electrolyte solution; and (E) the cycling capacity retention curve of LiClO4-PPy/Ti.
Inorganics 12 00125 g004
Figure 5. (A,B) Nyquist plots of EIS spectra of LiClO4/Ti and LiClO4-PPy/Ti.; and (C) equivalent circuit model used for fitting curves of EIS spectra.
Figure 5. (A,B) Nyquist plots of EIS spectra of LiClO4/Ti and LiClO4-PPy/Ti.; and (C) equivalent circuit model used for fitting curves of EIS spectra.
Inorganics 12 00125 g005
Figure 6. (A,B) Molecule structure models, (C,D) molecular electrostatic potentials (difference charge density), (E,F) total density of states curves, (G,H) band gap structure curves, (I) HOMO and LUMO frontier molecular orbital charge distribution of LiClO4/Ti and LiClO4-PPy/Ti.
Figure 6. (A,B) Molecule structure models, (C,D) molecular electrostatic potentials (difference charge density), (E,F) total density of states curves, (G,H) band gap structure curves, (I) HOMO and LUMO frontier molecular orbital charge distribution of LiClO4/Ti and LiClO4-PPy/Ti.
Inorganics 12 00125 g006aInorganics 12 00125 g006b
Figure 7. Mulliken charge population of (A) LiClO4/Ti and (B) LiClO4-PPy/Ti.
Figure 7. Mulliken charge population of (A) LiClO4/Ti and (B) LiClO4-PPy/Ti.
Inorganics 12 00125 g007
Table 1. Fitting parameter values of elements used in equivalent circuit model of LiClO4/Ti and LiClO4-PPy/Ti.
Table 1. Fitting parameter values of elements used in equivalent circuit model of LiClO4/Ti and LiClO4-PPy/Ti.
SubstanceCdl (mF cm−2)Rct (Ω cm−2)Ro (Ω cm−2)CPE (S sn cm−2)n
LiClO4/Ti0.0334 65,2500.3660 0.002630.746
LiClO4-PPy/Ti0.193021160.2226 0.034270.912
Table 2. Mulliken charge population of C, O, N, Cl, and Ti atoms in LiClO4/Ti and LiClO4-PPy/Ti.
Table 2. Mulliken charge population of C, O, N, Cl, and Ti atoms in LiClO4/Ti and LiClO4-PPy/Ti.
ElementLiClO4-TiLiClO4-PPy-Ti
O1−0.78−0.76
O2−0.80−0.77
O3−0.82−0.78
O4−0.95−0.90
Cl+2.45+2.47
C1/+0.03
C2/+0.04
N/(−0.58~−0.47)
Ti(−0.23~−0.04)(−0.05~−0.01)
(+0.13~+0.05)(+0.62~+0.08)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xie, Y.; Xu, J.; Lu, L.; Xia, C. Electrochemical Investigation of Lithium Perchlorate-Doped Polypyrrole Growing on Titanium Substrate. Inorganics 2024, 12, 125. https://doi.org/10.3390/inorganics12040125

AMA Style

Xie Y, Xu J, Lu L, Xia C. Electrochemical Investigation of Lithium Perchlorate-Doped Polypyrrole Growing on Titanium Substrate. Inorganics. 2024; 12(4):125. https://doi.org/10.3390/inorganics12040125

Chicago/Turabian Style

Xie, Yibing, Jing Xu, Lu Lu, and Chi Xia. 2024. "Electrochemical Investigation of Lithium Perchlorate-Doped Polypyrrole Growing on Titanium Substrate" Inorganics 12, no. 4: 125. https://doi.org/10.3390/inorganics12040125

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