Next Article in Journal
Super-Tough and Biodegradable Poly(lactide-co-glycolide) (PLGA) Transparent Thin Films Toughened by Star-Shaped PCL-b-PDLA Plasticizers
Next Article in Special Issue
Cellulose Nanocrystal Embedded Composite Foam and Its Carbonization for Energy Application
Previous Article in Journal
Simulational Tests of the Rouse Model
Previous Article in Special Issue
Numerical Modeling of the Mixing of Highly Viscous Polymer Suspensions in Partially Filled Sigma Blade Mixers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Key Factors in Enhancing Pseudocapacitive Properties of PANI-InOx Hybrid Thin Films Prepared by Sequential Infiltration Synthesis

1
Department of Materials Science and Engineering, Chungnam National University, Daejeon 34134, Republic of Korea
2
Department of Plasma Engineering, Korea Institute of Machinery & Materials (KIMM), Daejeon 34103, Republic of Korea
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(12), 2616; https://doi.org/10.3390/polym15122616
Submission received: 11 May 2023 / Revised: 1 June 2023 / Accepted: 6 June 2023 / Published: 8 June 2023
(This article belongs to the Special Issue Polymers Physics: From Theory to Experimental Applications)

Abstract

:
Sequential infiltration synthesis (SIS) is an emerging vapor-phase synthetic route for the preparation of organic–inorganic composites. Previously, we investigated the potential of polyaniline (PANI)-InOx composite thin films prepared using SIS for application in electrochemical energy storage. In this study, we investigated the effects of the number of InOx SIS cycles on the chemical and electrochemical properties of PANI-InOx thin films via combined characterization using X-ray photoelectron spectroscopy, ultraviolet–visible spectroscopy, Raman spectroscopy, Fourier transform infrared spectroscopy, and cyclic voltammetry. The area-specific capacitance values of PANI-InOx samples prepared with 10, 20, 50, and 100 SIS cycles were 1.1, 0.8, 1.4, and 0.96 mF/cm², respectively. Our result shows that the formation of an enlarged PANI-InOx mixed region directly exposed to the electrolyte is key to enhancing the pseudocapacitive properties of the composite films.

Graphical Abstract

1. Introduction

Organic–inorganic hybrid materials have received continued attention owing to their novel functionalities, which are not demonstrated in single-phase materials, whether organic or inorganic. In particular, electrically conductive organic–inorganic composite films have demonstrated enhanced properties for application in electrochemistry-related areas. Some recently reported examples of the benefits of using PANI–metal oxide composite films include enhanced efficiency in photovoltaic cells [1], sensitivities and linearities in sensors [2], photocatalytic efficiencies in catalysts [3], and retention properties in supercapacitors [4]. The different properties of hybrid thin films and their performances in different areas depend primarily on the synthesis routes of the thin films. Therefore, the development of novel techniques for preparing organic–inorganic hybrid thin films has received attention in various areas.
Sequential infiltration synthesis (SIS) is a new vacuum-based technique for preparing organic–inorganic composites and is considered a variant of atomic layer deposition (ALD). Although ALD exploits self-limiting chemical reactions on the surfaces, SIS utilizes chemical reactions within the bulk polymer phase [5]. For SIS reactions to occur readily, the precursors designated to be used must first infiltrate the polymer phases. Typically, the entrapment of infiltrated SIS precursors by specific functional groups of the polymer is preferred. Poly(methylmethacrylate) (PMMA) [6] and poly(2-vinylpyridine) (P2VP) [7] are typical types of polymers that have been widely used for SIS because the carbonyl groups in PMMA and pyridine groups in P2VP undergo Lewis acid–base reactions with typical SIS precursors such as trimethylaluminum and titanium tetrachloride (TiCl4). However, the potential applications of hybrid thin films based on PMMA and P2VP are limited to areas where electrical conductivity is not required. Only several studies have been conducted on SIS with conducting polymers and their applications in electrochemistry.
Polyaniline (PANI) is a representative conducting polymer with controllable electrical conductivity owing to doping. PANI can exist in three different chemical states, namely, leucoemeraldine, pernigraniline, and emeraldine, depending on the degrees of oxidation and reduction [8,9]. PANI with an emeraldine base can be transformed into an emeraldine salt, which exhibits high electrical conductivity (~102 S/cm) when acid-doped [10]. Wang et al. reported the SIS process of doping PANI with SnCl4 and MoCl5 vapors [11], which exhibit the Lewis acidic nature; the doped PANI exhibited a moderate conductivity of ~9.8 × 10−5 S/cm. PANI-ZnO (~18.42 S/cm) [12] and PANI-InOx (4–9 S/cm) [4] composite thin films prepared via SIS showed electrically conductive properties. Previously, we demonstrated the significant potential of PANI-InOx composite films prepared using SIS for electrochemical energy storage, which warrants follow-up studies on the same system [4].
The aim of this study was to investigate the influence of metal oxides on the electrochemical properties of polyaniline–metal oxide composites as energy storage materials utilizing the SIS. Research related to SIS focusing on conducting polymers is limited to several papers, and studies specifically investigating their electrochemical properties are scarce. In this study, we investigated the variations in the chemical and electrochemical properties of PANI-InOx films prepared via SIS as a function of the SIS cycle number. PANI-InOx films exhibit a graded concentration of InOx along the direction of the film thickness, where their structure is determined by the number of cycles. A combination of ultraviolet–visible (UV-vis) spectroscopy, Raman spectroscopy, and attenuated total reflectance–Fourier transform infrared (ATR-FTIR) spectroscopy was performed to understand the variation in the chemical structure of PANI in response to alloying with InOx. The superior pseudocapacitive properties of the sample with the optimized cycle number (50 cy) are attributable to the increased volume of the PANI-InOx mixed region, which is exposed to the electrolyte.

2. Materials and Methods

2.1. Sample Preparation

PANI with an emeraldine base powder (Mw ~10,000, Sigma–Aldrich, Saint Louis, MI, USA) was dissolved in methyl-2-pyrrolidone (≥99%, Sigma–Aldrich, Saint Louis, MI, USA) with a concentration of 30 mg/mL. The solution was stirred for 24 h at 80 °C and 850 rpm. The solution was spun onto prepared substrates at 2000 rpm, and the as-spun substrates were baked at 70 °C in air. The thickness of the prepared PANI thin films was approximately 37 nm. Subsequently, SIS was performed using a thermal ALD reactor (Daeki HighTech, Daejeon, Republic of Korea) with a cross-flow design. The precursors used for the SIS were trimethylindium (TMIn, 99.999%, EasyChem) and H2O (99.999%, Sigma–Aldrich, Saint Louis, MI, USA). Ar carrier gas (99.999%) was continuously flowed at 5 sccm during the entire SIS. Both the TMIn and H2O half-cycles involved a 1 s dose, 120 s of exposure, and 120 s of purging. The reactor chamber was isolated from the pump during the exposure step to facilitate the infiltration of the precursors into the polymer matrix. The SIS-prepared substrates were annealed at 270 °C for 1 h in the forming gas of H2–N2 (~3.9% H2 in N2).
Different types of substrates were used for different characterization methods: an Si substrate (n type, 1–10 Ohm·cm, iTASCO) with a 500-nm-thick SiO2 layer was used for X-ray photoelectron spectroscopy (XPS) and Raman spectroscopy. Au-coated Si substrate (Au thickness: ~90 nm) was used for ATR-FTIR spectroscopy. A fused silica (iNexus, Inc., Seongnam, Republic of Korea), which had a transmittance of ~90% or higher in the wavelength range > 250 nm, was utilized for UV-vis spectroscopy. Electrochemical experiments were performed using glass substrates with a fluorine-doped tin oxide (FTO) layer measuring ~600 nm thick (NSG TEC 7, Pilkington, Lathom, UK).

2.2. Sample Characterization

HRXPS depth profiling was performed using an X-ray photoelectron spectrometer (K-alpha, Thermo Scientific, Waltham, MA, USA) with Ar+ ion beams at 1 kV and an etching time of 10 s. The X-ray source used was monochromatic Al Kα (1487 eV). The surfaces of the annealed samples were partly scratched using stainless-steel tweezers to create a surface step on the sample, and the thickness of the PANI-InOx thin film was measured using a stylus profiler (Alpha-Step® D-500, KLA, Milpitas, CA, USA). The cleanliness of the scratched surface was confirmed based on spectroscopic ellipsometry data (FS-1, Film Sense, Lincoln, NE, USA) obtained from the surface and a comparison with those of a bare Si substrate. Raman spectroscopy was performed using a Raman spectrometer (LabRAM HR-800, Horiba, Japan) under the following conditions: 514 nm laser source measuring 0.7 μm, 1800 gr/mm grating, 10 s acquisition time, and 10 specular accumulations. UV-vis spectroscopy was performed using a UV-vis spectrometer (UV-2600, Shimadzu, Japan), where an FTIR spectrometer (Vertex 80v, Bruker, Billerica, MA, USA) with a mercury–cadmium–telluride detector and a diamond attenuated total reflection (ATR) crystal were used to obtain the ATR-FTIR spectroscopy data at a spectral resolution of 4 cm−1. An electrochemical analyzer (CHI602E, CH Instruments, Bee Cave, TX, USA) was used to perform cyclic voltammetry (CV) experiments using a three-electrode setup comprising a Ag/AgCl reference electrode, a Pt wire counter electrode, and a PANI-InOx FTO/glass working electrode. CV data were obtained using a pH 7 buffer solution as the electrolyte.

3. Results and Discussion

We analyzed PANI-InOx composite thin films, which were prepared under different SIS cycles (10, 20, 50, and 100 cy) and annealed in a reducing atmosphere. The sample structure could be summarized as follows: (1) an InOx-rich region, (2) a PANI-InOx mixed region, and (3) a PANI-rich region, as shown in Figure 1a. The thicknesses and chemical compositions of the three regions differed depending on the number of SIS cycles (Figure 1b). The samples with 10 and 20 SIS cycles did not contain an InOx surface layer and only presented a PANI-InOx mixed region and a PANI bulk region. Owing to the repetition of the SIS cycle, the InOx-rich region near the surface of the PANI at times prevented the additional infiltration of the TMIn precursor, thereby resulting in the formation of a thicker InOx layer, which further developed via a mechanism similar to ALD. However, TMIn infiltration in later SIS cycles may not have been completely hindered because the concentration in the PANI-rich region of the 100 cy sample decreased gradually from ~40 to 0 at%.
Figure 1c shows the oxygen (O) and nitrogen (N) HRXPS data obtained at different locations on the PANI-InOx film, as shown in Figure 1b. The O 1s HRXPS data were deconvoluted into three or four peaks originating from the following components: lattice oxygen (In-O) from InOx at ~529.9 eV, oxygen vacancy ( V o ˙ ˙ ) at ~531.0 eV, indium hydroxide (In-OH) at ~532.1 eV, and lattice oxygen (Si-O) from SiO2 at ~533.1 eV [13,14]. In all four samples, the In-O component was the most dominant in the topmost region, which was either an InOx-PANI mixed region (samples with 10 and 20 SIS cy) or InOx-rich regions (Samples with 50 and 100 SIS cy). The components of higher BEs (i.e., −OH and V o ˙ ˙ ) became more dominant compared to the In-O component as the HRXPS analysis region shifted toward the substrate (i.e., PANI-rich region). This is consistent with the previous SIS results, which indicated that the oxidation of the SIS precursors within the polymer matrix was less complete than that on the polymer surface [6,15]. The average stoichiometries of the four samples were InO0.85, InO, InO1.38 and InO1.44 for 10, 20, 50, and 100 cycles, respectively. The stoichiometry trend was reasonable, considering that the proportion of surface-like InOx compared with that of bulk-like InOx (i.e., synthesized within the polymer matrix) enhanced as the SIS cycle number increased.
The N 1s HRXPS data of the four samples were deconvoluted into three components: quinonoid imine (−N=) at ~398.4 eV, benzenoid amine (−NH−) at ~399.5 eV, and protonated amine/imine state (−NH2+–, =NH+−) at ~400.4 eV [16]. In all the HRXPS spectra, the amine component was more dominant than those of the other components. Meanwhile, the PANI with an emeraldine base usually contained equal amounts of imine and amine components. The presence of protonated species along with a decrease in the number of imine units suggested that the protonated species may have originated from the imine units. PANI doped with HCl contains protonated species transformed from the imine components [17]. However, no clear correlation was indicated between the percentage of protonated species and the InOx content in any sample. Therefore, further studies are necessary to determine the potential chemical reactions contributing to the formation of protonated species during InOx alloying.
Figure 2a shows the UV-vis transmittance spectra of the four samples. The samples with 10, 20, and 50 cy showed a weak absorption band at ~610 nm, which was assigned to n–π* transition between the benzenoid and quinonoid rings. The increase in the absorption below ~400 nm from the three sample was likely related to the π–π* transition of the benzenoid ring, which is known to be located at ~320 nm [18]. The 100 cy sample showed stronger absorption in the UV region (<400 nm), whereas the n–π* transition was primarily suppressed. The absorption in the UV region was likely due to absorption by the thick InOx layer. The Tauc plot of the 100 cy SIS shows that the bandgap of InOx was ~3.5 eV, which was slightly lower level than that of the dense InOx thin films reported in the literature [19,20]. The smaller bandgap of InOx along with the presence of tail states was reasonable, considering that a significant portion of the InOx phase was present within the polymer matrix along with a high concentration of oxygen vacancies. The optical bandgap of PANI-InOx samples tends to decrease as the number of SIS cycles decreases (Figure S1). The ATR-FTIR spectra of the four samples (Figure 2b) showed IR bands related to the PANI phase: (1) stretching of the quinonoid ring at ~1600 cm−1, (2) stretching of the benzenoid ring at ~1512 cm−1, (3) stretching of C–N of the secondary aromatic amine at 1300 cm−1, and (4) out-of-plane C–H deformation of the 1,4-distributed aromatic ring at 823 cm−1 [21,22,23]. The 100 cy samples showed significant IR bands, with lower intensities compared with the other samples owing to the presence of the InOx surface layer. Similarly, the Raman spectra of the four samples (Figure 2c) exhibited significant Raman bands associated with the PANI phase, as summarized in Table 1. The presence of the C–N+• stretching (radical cations) band at ~1350 cm−1 in all samples is consistent with the presence of protonated amine/imine species indicated in the HRXPS analysis. Radical cation bands are typically observed in acid-doped PANI, which suggests that alloying with InOx may offer similar effects on the acid doping of PANI.
The CV results for the samples, measured at a scan rate of 10 mV/s, are shown in Figure 3. All the samples showed a pair of redox peaks at similar potentials (i.e., ~0.2 V vs. Ag/AgCl and ~−0.05 V vs. Ag/AgCl). The 50 cy sample indicated better-defined redox peaks with a higher current compared with the other samples. The area-specific capacitance values of PANI-InOx samples prepared with 10, 20, 50, and 100 SIS cycles were 1.1, 0.8, 1.4, and 0.96 mF/cm², respectively. A detailed explanation of this calculation is reported in the literature [4]. The CV measurements were performed multiple times using different samples prepared under the same SIS conditions. The redox peak positions of the CV curves varied slightly within a ~100 mV range. Therefore, the subtle variation in the peak position observed for the different SIS samples (10, 20, 50, and 100 cy) is considered to be within the experimental error. CV curves collected at different scan rates are provided in Supplementary Materials. In order to investigate the capacitance stability of the sample after prolonged exposure to the electrolyte, we conducted a CV experiment consisting of 1000 cycles (Figure S2). This experiment assessed the evolution of the capacitance over time in response to extended electrolyte exposure. In the electrochemical impedance spectroscopy (EIS) test conducted on the PANI-InOx 50 cycle, pure PANI, and pure InOx samples in a previous study, semicircles were observed at high frequencies and straight lines were observed at low frequencies [4]. Among the three samples, the composite sample exhibited the smallest semicircle, indicating a lower charge transfer resistance.
Further analysis is necessary to identify the redox reactions contributing to the observed CV peaks. However, the conversion of the emeraldine and pernigraniline states was speculated to be the primary redox reaction in the PANI-InOx samples in our previous study. The enhanced peak current of the 50 cy sample might have been related to the larger thickness of the PANI-InOx mixed region (Figure 1b) compared with those of the 10 and 20 cy samples. The 100 cy sample exhibited a sufficiently thick PANI-InOx mixed region; however, the InOx surface layer likely prevented direct contact between the mixed region and the electrolyte [4]. Furthermore, the 50 cy sample had a larger proportion of protonated amine/imine structures (Figure 1c), a fact which is likely related to the PANI-InOx mixed region.

4. Conclusions

It is only recently that SIS was proved to be a promising route for the preparation of organic–inorganic hybrid films for electrochemistry and electrochemical energy storage. The goal of this study was to provide a more comprehensive understanding of the effects of the number of SIS cycles (10, 20, 50, and 100 cycles) on the chemical and electrochemical properties of PANI-InOx composite films. The PANI-InOx films showed a graded composition of InOx within the PANI matrix, with ample concentrations of oxygen vacancies and hydroxide components. The entire film structure was composed of two or three components, including an InOx-rich region, a PANI-InOx mixed region, and a PANI-rich region, and its structure depended on the number of cycles. Combined characterization using HRXPS, UV-vis, Raman, and FTIR spectroscopy consistently revealed the presence of cationic radicals, which might have been related to the transition from the quinonoid structure to the benzenoid structure. The 50 cy samples showed the highest pseudocapacitance among the tested samples, which was likely due to the relatively thick electrochemically active PANI-InOx region and the exposure of the PANI-InOx region to the electrolyte.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym15122616/s1, Figure S1: Tauc plots of PANI-InOx sample prepared with 10, 20, and 50 SIS cycles.; Figure S2: Capacity retention test of PANI-InOx samples prepared with 10, 20, 50, and 100 SIS cycles.; Figure S3: CV curves of 50 cy PANI-InOx sample collected at different scan rates. Reference [24] is cited in the supplementary materials.

Author Contributions

Conceptualization, J.H., H.-U.K. and N.J.; formal analysis, J.H.; investigation, J.H.; Data Curation, H.-U.K.; writing—original draft preparation, J.H.; writing—review and editing, N.J.; visualization, J.H.; supervision, N.J.; project administration, N.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. RS-2023-00214170). This research was supported by Chungnam National University (2021–2022). This work was supported by KIMM institutional program (NK242F) and NST/KIMM.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nemade, K.; Dudhe, P.; Tekade, P. Enhancement of photovoltaic performance of polyaniline/graphene composite-based dye-sensitized solar cells by adding TiO2 nanoparticles. Solid State Sci. 2018, 83, 99–106. [Google Scholar] [CrossRef]
  2. Murugesan, T.; Kumar, R.R.; Anbalagan, A.K.; Lee, C.-H.; Lin, H.-N. Interlinked Polyaniline/ZnO Nanorod Composite for Selective NO2 Gas Sensing at Room Temperature. ACS Appl. Nano Mater. 2022, 5, 4921–4930. [Google Scholar] [CrossRef]
  3. Chen, S.; Huang, D.; Zeng, G.; Xue, W.; Lei, L.; Xu, P.; Deng, R.; Li, J.; Cheng, M. In-situ synthesis of facet-dependent BiVO4/Ag3PO4/PANI photocatalyst with enhanced visible-light-induced photocatalytic degradation performance: Synergism of interfacial coupling and hole-transfer. Chem. Eng. J. 2020, 382, 122840. [Google Scholar] [CrossRef]
  4. Ham, J.; Park, S.; Jeon, N. Conductive Polyaniline–Indium Oxide Composite Films Prepared by Sequential Infiltration Synthesis for Electrochemical Energy Storage. ACS Omega 2023, 8, 946–953. [Google Scholar] [CrossRef] [PubMed]
  5. Leng, C.Z.; Losego, M.D. A physiochemical processing kinetics model for the vapor phase infiltration of polymers: Measuring the energetics of precursor-polymer sorption, diffusion, and reaction. Phys. Chem. Chem. Phys. 2018, 20, 21506–21514. [Google Scholar] [CrossRef]
  6. Ham, J.; Ko, M.; Choi, B.; Kim, H.-U.; Jeon, N. Understanding Physicochemical Mechanisms of Sequential Infiltration Synthesis toward Rational Process Design for Uniform Incorporation of Metal Oxides. Sensors 2022, 22, 6132. [Google Scholar] [CrossRef]
  7. Biswas, M.; Libera, J.A.; Darling, S.B.; Elam, J.W. Polycaprolactone: A Promising Addition to the Sequential Infiltration Synthesis Polymer Family Identified through in Situ Infrared Spectroscopy. ACS Appl. Polym. Mater. 2020, 2, 5501–5510. [Google Scholar] [CrossRef]
  8. Babel, V.; Hiran, B.L. A review on polyaniline composites: Synthesis, characterization, and applications. Polym. Compos. 2021, 42, 3142–3157. [Google Scholar] [CrossRef]
  9. Zare, E.N.; Makvandi, P.; Ashtari, B.; Rossi, F.; Motahari, A.; Perale, G. Progress in Conductive Polyaniline-Based Nanocomposites for Biomedical Applications: A Review. J. Med. Chem. 2020, 63, 1–22. [Google Scholar] [CrossRef]
  10. Stejskal, J.; Trchová, M.; Bober, P.; Humpolíček, P.; Kašpárková, V.; Sapurina, I.; Shishov, M.A.; Varga, M. Conducting Polymers: Polyaniline. In Encyclopedia of Polymer Science and Technology; Wiley: Hoboken, NJ, USA, 2015; pp. 1–44. [Google Scholar] [CrossRef]
  11. Wang, W.; Yang, F.; Chen, C.; Zhang, L.; Qin, Y.; Knez, M. Tuning the Conductivity of Polyaniline through Doping by Means of Single Precursor Vapor Phase Infiltration. Adv. Mater. Interfaces 2017, 4, 1600806. [Google Scholar] [CrossRef]
  12. Wang, W.; Chen, C.; Tollan, C.; Yang, F.; Beltrán, M.; Qin, Y.; Knez, M. Conductive Polymer–Inorganic Hybrid Materials through Synergistic Mutual Doping of the Constituents. ACS Appl. Mater. Interfaces 2017, 9, 27964–27971. [Google Scholar] [CrossRef] [PubMed]
  13. Hoch, L.B.; He, L.; Qiao, Q.; Liao, K.; Reyes, L.M.; Zhu, Y.; Ozin, G.A. Effect of Precursor Selection on the Photocatalytic Performance of Indium Oxide Nanomaterials for Gas-Phase CO2 Reduction. Chem. Mater. 2016, 28, 4160–4168. [Google Scholar] [CrossRef]
  14. Cañón, J.; Teplyakov, A.V. XPS characterization of cobalt impregnated SiO2 and γ-Al2O3. Surf. Interface Anal. 2021, 53, 475–481. [Google Scholar] [CrossRef]
  15. Ko, M.; Kirakosyan, A.; Kim, H.-U.; Seok, H.; Choi, J.; Jeon, N. A new nanoparticle heterostructure strategy with highly tunable morphology via sequential infiltration synthesis. Appl. Surf. Sci. 2022, 593, 153387. [Google Scholar] [CrossRef]
  16. Tantawy, H.R.; Kengne, B.-A.F.; McIlroy, D.N.; Nguyen, T.; Heo, D.; Qiang, Y.; Aston, D.E. X-ray photoelectron spectroscopy analysis for the chemical impact of solvent addition rate on electromagnetic shielding effectiveness of HCl-doped polyaniline nanopowders. J. Appl. Phys. 2015, 118, 175501. [Google Scholar] [CrossRef]
  17. Song, E.; Choi, J.-W. Conducting Polyaniline Nanowire and Its Applications in Chemiresistive Sensing. Nanomaterials 2013, 3, 498–523. [Google Scholar] [CrossRef]
  18. Andriianova, A.N.; Biglova, Y.N.; Mustafin, A.G. Effect of structural factors on the physicochemical properties of functionalized polyanilines. RSC Adv. 2020, 10, 7468–7491. [Google Scholar] [CrossRef]
  19. Agbenyeke, R.E.; Jung, E.A.; Park, B.K.; Chung, T.-M.; Kim, C.G.; Han, J.H. Thermal atomic layer deposition of In2O3 thin films using dimethyl(N-ethoxy-2,2-dimethylcarboxylicpropanamide)indium and H2O. Appl. Surf. Sci. 2017, 419, 758–763. [Google Scholar] [CrossRef]
  20. Maeng, W.J.; Choi, D.-W.; Park, J.; Park, J.-S. Indium oxide thin film prepared by low temperature atomic layer deposition using liquid precursors and ozone oxidant. J. Alloys Compd. 2015, 649, 216–221. [Google Scholar] [CrossRef]
  21. Su, N. Polyaniline-Doped Spherical Polyelectrolyte Brush Nanocomposites with Enhanced Electrical Conductivity, Thermal Stability, and Solubility Property. Polymers 2015, 7, 1599–1616. [Google Scholar] [CrossRef] [Green Version]
  22. Patra, B.N.; Majhi, D. Removal of Anionic Dyes from Water by Potash Alum Doped Polyaniline: Investigation of Kinetics and Thermodynamic Parameters of Adsorption. J. Phys. Chem. B 2015, 119, 8154–8164. [Google Scholar] [CrossRef] [PubMed]
  23. Wong, P.-Y.; Phang, S.-W.; Baharum, A. Effects of synthesised polyaniline (PAni) contents on the anti-static properties of PAni-based polylactic acid (PLA) films. RSC Adv. 2020, 10, 39693–39699. [Google Scholar] [CrossRef] [PubMed]
  24. Yue, J.; Lin, L.; Jiang, L.; Zhang, Q.; Tong, Y.; Suo, L.; Hu, Y.-s.; Li, H.; Huang, X.; Chen, L. Interface Concentrated-Confinement Suppressing Cathode Dissolution in Water-in-Salt Electrolyte. Adv. Energy Mater. 2020, 10, 2000665. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic illustration of structure of PANI-InOx/SiO2/Si samples. InOx content shows graded concentration along film thickness direction. (b) XPS depth profiles showing C, In, N, O, and Si atomic concentrations for four SIS samples with different cycle numbers: 10, 20, 50, and 100 cy. (c) O 1s and N 1s HRXPS data obtained at different locations (i.e., InOx-rich region, PANI-InOx mixed region, and PANI-rich region) in four samples. Each location at which HRXPS data were captured are shown as arrows in (b).
Figure 1. (a) Schematic illustration of structure of PANI-InOx/SiO2/Si samples. InOx content shows graded concentration along film thickness direction. (b) XPS depth profiles showing C, In, N, O, and Si atomic concentrations for four SIS samples with different cycle numbers: 10, 20, 50, and 100 cy. (c) O 1s and N 1s HRXPS data obtained at different locations (i.e., InOx-rich region, PANI-InOx mixed region, and PANI-rich region) in four samples. Each location at which HRXPS data were captured are shown as arrows in (b).
Polymers 15 02616 g001
Figure 2. (a) UV-vis transmittance spectra, (b) ATR-FTIR absorbance spectra, and (c) Raman spectra of PANI-InOx samples at different number of cycles (10, 20, 50, and 100 cy) and PANI-only sample. The PANI-only sample was annealed under the same conditions as the PANI-InOx samples. Inset of (a) shows Tauc plot of 100 cy PANI-InOx sample. The dashed lines in panel (c) mark the location of Raman bands observed in the PANI-InOx film, which are also summarized in Table 1.
Figure 2. (a) UV-vis transmittance spectra, (b) ATR-FTIR absorbance spectra, and (c) Raman spectra of PANI-InOx samples at different number of cycles (10, 20, 50, and 100 cy) and PANI-only sample. The PANI-only sample was annealed under the same conditions as the PANI-InOx samples. Inset of (a) shows Tauc plot of 100 cy PANI-InOx sample. The dashed lines in panel (c) mark the location of Raman bands observed in the PANI-InOx film, which are also summarized in Table 1.
Polymers 15 02616 g002
Figure 3. CV curves of PANI-InOx films prepared with different SIS cycles (10, 20, 50, and 100 cy) and PANI-only film. The CVs were collected at a scan rate of 10 mV/s in an aqueous electrolyte of neutral pH.
Figure 3. CV curves of PANI-InOx films prepared with different SIS cycles (10, 20, 50, and 100 cy) and PANI-only film. The CVs were collected at a scan rate of 10 mV/s in an aqueous electrolyte of neutral pH.
Polymers 15 02616 g003
Table 1. Major Raman bands identified in the PANI-InOx and PANI-only samples.
Table 1. Major Raman bands identified in the PANI-InOx and PANI-only samples.
Raman Shift (cm−1)Assignment
1610C=C stretching vibration of a quinonoid ring
1560N–H bending
1405C–C stretching vibrations in a quinonoid ring
1350C–N+• Radical cation
1240C–N stretching in a benzenoid ring
1170C–H in-plane C–H bending quinonoid ring
817out-of-plane C–H vibration
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

Ham, J.; Kim, H.-U.; Jeon, N. Key Factors in Enhancing Pseudocapacitive Properties of PANI-InOx Hybrid Thin Films Prepared by Sequential Infiltration Synthesis. Polymers 2023, 15, 2616. https://doi.org/10.3390/polym15122616

AMA Style

Ham J, Kim H-U, Jeon N. Key Factors in Enhancing Pseudocapacitive Properties of PANI-InOx Hybrid Thin Films Prepared by Sequential Infiltration Synthesis. Polymers. 2023; 15(12):2616. https://doi.org/10.3390/polym15122616

Chicago/Turabian Style

Ham, Jiwoong, Hyeong-U Kim, and Nari Jeon. 2023. "Key Factors in Enhancing Pseudocapacitive Properties of PANI-InOx Hybrid Thin Films Prepared by Sequential Infiltration Synthesis" Polymers 15, no. 12: 2616. https://doi.org/10.3390/polym15122616

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