Advertisement

Effects of surfactants on the preparation of MnO2 and its capacitive performance

Abstract

Amorphous hydrated manganese dioxide (MnO2) was prepared as an electrode material for supercapacitors by liquid co-precipitation in the presence of polyvinylpyrrolidone (PVP), polyethylene glycol (PEG) and sodium dodecylbenzenesulfonate (SDBS) respectively. The obtained samples were characterized by x-ray diffraction (XRD), field emission scanning electron microscopy (FE-SEM), high-resolution transmission electron microscopy (HRTEM), and electrochemical methods. Physical characterizations confirmed that the addition of surfactants played an important role in the preparation of MnO2. The specific surface areas of MnO2 with the addition of PEG, SDBS and PVP were 169.92 m2/g, 137.40 m2/g and 196.64 m2/g, respectively, and the corresponding capacitances were 207.9 F/g, 187.5 F/g and 238.7 F/g. Compared with the sample without surfactants, the specific surface area and capacitance of the sample with the addition of PVP were improved by 92.2% and 53.1%, respectively. Moreover, the electrode showed good cycle stability at the current density of 120 mA/g, and 91.1% of its specific capacitance still remained after 500 cycles. It was concluded that this performance improvement was attributed to the electrostatic stabilization of the multivariate alkyl residue and cyano group (—NCO) as anchoring group, as well as the steric hindrance effect from lateral polarity groups of pentabasic ring in PVP structure.

J Appl Biomater Funct Mater 2017; 15(Suppl. 1): e7 - e12

Article Type: ORIGINAL RESEARCH ARTICLE

DOI:10.5301/jabfm.5000356

OPEN ACCESS ARTICLE

Authors

Yin Sun, Hangfei Dang, Naibao Huang, Dongchao Wang, Chenghao Liang

Article History

Disclosures

Financial support: This work was financially supported by National Natural Science Foundation of China (21676040, 21276036), the National Key Research and Development Program of China (2016yfb0101200, 2016yfb0101206), the Liaoning Provincial Nature Science Foundation of China (2014025018), and the Fundamental Research Funds for the Central Universities (3132016341), the Foundation of Liaoning Educational Committee (L2014199), the Shanghai Key Laboratory of Digitalized Manufacturing of Complex Sheet Metal Structures (No. 2014004).
Conflict of interest: None of the authors has financial interest related to this study to disclose.

This article is available as full text PDF.

Download any of the following attachments:

Introduction

In recent years, supercapacitors have attracted widespread attentions from all over the world because of their high-power density, quick charge/discharge capability, long cycle life and high safety (1, 2). According to the difference of charge storage mechanism, supercapacitors can be classified into two types: the electric double-layer capacitors (EDLCs) and pseudocapacitors. Since pseudocapacitors store energy on the electrode surface or in the bulk phase through highly reversible chemical adsorption/desorption or oxidation/reduction reactions (3), they usually possess much higher energy density than EDLCs. RuO2 is the most typical representative electrode material for pseudocapacitors. However, the disadvantages of high expense and environmental disruption lead to its low attractiveness in large-scale applications. Instead, materials of high natural abundance and low cost, such as cobalt oxide, nickel oxide, manganese oxide, etc., have become the ideal candidates as substitute materials (4-5-6). Among these materials, manganese dioxide (MnO2) is the most widely studied.

Studies on the performance of MnO2 electrode materials mainly concentrate on the preparation of composite materials and nanostructured materials (7-8-9-10). MnO2 electrode materials can be prepared by various methods (11-12-13-14). It has been confirmed that preparation methods have significant effects on the morphology, structure and capacitance performance of MnO2 electrode materials. The samples prepared by liquid co-precipitation often present a weakly crystalline structure or an amorphous structure (15, 16), which is conducive to the intercalation and deintercalation of electrolyte ions, thus improving the utilization rate of active materials. As a result, these materials always have excellent electrochemical performance.

However, because of the electrostatic interaction between product particles and the existence of non-bridging hydroxyl groups in the produced gel, a large number of product particles are agglomerated together, which would greatly degrade the performance of MnO2 (17). To solve the problem, surfactants are widely used to modify the preparation process of MnO2. Making use of liquid co-precipitation with different concentrations of cationic surfactant CTAB, Zhang et al (18) prepared MnO2 nano-hollow spheres. Compared with MnO2 synthesized without surfactants, the specific capacitance was increased from 126 F/g to 178 F/g at the current density of 500 mA/g. Jiang et al (19) synthesized MnO2 by a similar process with the addition of Pluronic P123 surfactant, and the maximum pseudocapacitance reached 176 F/g. A novel surfactant-assisted dilute polymerization technique was used, Senthilkumar et al (20) have prepared polyaniline nanofibers doped with citric acid, which exhibited a specific capacitance of 298 F/g.

To sum up, the performance of supercapacitor electrode materials can be improved by adding proper surfactants. Polyvinylpyrrolidone (PVP) is a kind of non-ionic polymer surfactant. To the authors’ knowledge, there are no reports about the effects of PVP surfactant on the performance of MnO2 by liquid co-precipitation. In this paper, amorphous hydrated MnO2 was prepared by liquid co-precipitation in the presence of non-ionic surfactant PVP. In order to compare the performance differences caused by different surfactants, amorphous hydrated MnO2 was also prepared by the above-mentioned method through adding another two widely used surfactants, i.e., polyethylene glycol (PEG) and sodium dodecylbenzenesulfonate (SDBS). The effects of different surfactants on the morphology, crystal structure and electrochemical properties of the prepared electrode materials were investigated. The results showed that the performance of the electrode materials prepared with non-ionic surfactants (PEG and PVP) was better than that prepared with ionic surfactant SDBS.

Experimental

Synthesis of MnO2 by liquid co-precipitation

Solutions of 0.3 M MnCl2 and 0.2 M KMnO4 were prepared by dissolving MnCl2·4H2O (5.94 g) and KMnO4 (3.16 g) of analytical grade, respectively, in de-ionized water. Then, MnCl2 solution (with the addition of 0.5 g PVP) was added into KMnO4 solution at 60°C with a speed of 60 drop/min under continuous stirring. After full reaction and subsequent ageing process for 8 h, the suspension was filtered, and washed with de-ionized water and ethanol repeatedly by vacuum filtration. The obtained sample was dried at 100°C for 12 h in an oven, and then ground thoroughly by the mortar to produce MnO2 powder, which was marked as Sample d. For comparison, the MnCl2 solutions without surfactant and with the same mass fraction of PEG or SDBS were also used to prepare MnO2 powder by the same process, and the produced samples were marked as Sample a, Sample b and Sample c, respectively.

Structural characterization

Crystalline structure of the samples was tested by x-ray diffraction (XRD, Rigaku, DMAX- Ultima+ diffraction meter) with Cu Kα radiation (λ = 1.5404 Å). The morphology was observed using field emission scanning electron microscopy (FE-SEM, SUPRA 55 SAPPHHIRE) and transmission electron microscope (TEM, JEOL JEM-2100). The Brunauer-Emmett-Teller (BET) surface area was measured by WBL-8XX BET surface analyzer.

Electrochemical performance

The as-prepared MnO2 powder, activated carbon (XC-72) and binder were mixed according to a weight ratio of 75:15:10. The binder was the mixture of PVDF and N-methyl-2-pyrrolidone (NMP) with a weight ratio of 1:10. The mixture was smeared into a nickel foam, and then dried in vacuum at 100°C for 8 h. Subsequently, the foam was pressed at 10 MPa to obtain a wafer of 20 mm in diameter. Electrochemical test was conducted on the VMP3 (EG&G) electrochemical workstation using a three-electrode system, with the platinum net, a saturated calomel electrode (SCE) and the prepared foam as the counter electrode, reference electrode, and the working electrode, respectively. All electrochemical measurements were carried out in 6 M KOH solution and the galvanostatic charge/discharge test was at the current density of 120 mA/g.

Results and discussion

XRD patterns and BET specific surface area analysis

Figure1 shows the XRD patterns of samples prepared with different surfactants. As can be seen, the four samples have similar diffraction patterns, and present strong diffraction peaks at 37.2° and 66.7° and relatively weak diffraction peaks at 56.0° and 78.9°, which, respectively, correspond to (100) and (110) crystal planes and (102) and (200) planes of α-MnO2 according to the standard cards (JCPDS No. 30-0820). However, the intensity of diffraction peaks is relatively weak and severely broadened. The results indicated that the addition of different surfactants during preparation had no effects on the crystalline structure of the samples. According to the XRD patterns, the prepared products are amorphous α-MnO2·nH2O (21, 22). The amorphous structure is beneficial to the processes of de-embedding and embedding of the proton, which can result in a rapid, reversible chemical absorption/desorption or oxidation/reduction reaction on the electrode surface or within the bulk phase, thereby generating high pseudocapacitance. Because of the amorphous structure, MnO2 is more suitable for the application concerning the supercapacitor electrode material (23).

X-ray (XRD) patterns of the prepared samples by adding different surfactants: (A) without surfactants; (B) adding polyethylene glycol; (C) adding sodium dodecylbenzenesulfonate; (D) adding polyvinylpyrrolidone.

Figure 2 shows the specific surface areas of MnO2 prepared by different surfactants. The specific surface areas for Samples a, b, c and d are 102.32 m2/g, 169.92 m2/g, 137.40 m2/g and 196.64 m2/g, respectively. It can be seen that adding surfactants can effectively promote the dispersion of the samples and thus increase the specific surface area. Compared with Sample a, the specific surface areas of Samples b, c and d increase by 66.1%, 32.3% and 92.2%, respectively. From the perspective of increasing specific surface area, PVP has the best effect, followed by PEG, and SDBS has the worst effect.

Comparison for the specific surface areas of the samples prepared with different surfactants.

Morphology and microstructure analysis

In order to observe the microstructure of the obtained samples by adding different surfactants, the morphology was observed by FE-SEM and HRTEM. As Figure 3 shows, MnO2 particles tend to have smaller particle size, more regular shape and better dispersion after adding surfactants during preparation. The sample without surfactants (Sample a) has irregular particles and presents a serious agglomeration, with an average particle size of about 400 nm. By contrast, the sample with PVP surfactant (Sample d) has irregular spherical particles, and their average particle size is around 200 nm. From the energy-dispersive spectroscopy (EDS) pattern of Sample d (see supplementary figure S1, available online as supplementary material at www.jab-fm.com), only the peaks of Mn and O can be observed. Combined with the XRD patterns, it can be further determined that the product is amorphous MnO2. Figure 4 illustrates the TEM morphology of Sample d. It can be seen that there are many long burrs of about 30-50 nm in length on the surface of the spherical particles. The corresponding selected area electron diffraction (SAED) pattern (the inset shown in Fig. 4) shows a few broad halos without diffraction spots, once again confirming the amorphous structure and low crystallinity of the obtained material. In the process of preparation, the added surfactant could form micelles, reverse micelles, vesicles and other organized assemblies, which could be used as a “micro reactor” to inhibit the agglomeration of nanoparticles (24).

Scanning electron microscopy (SEM) images of the obtained samples: (A) without surfactants, (B) with the addition of polyethylene glycol (PEG), (C) with the addition of sodium dodecylbenzenesulfonate (SDBS), and (D) with the addition of polyvinylpyrrolidone (PVP).

Transmission electron microscope (TEM) image of MnO2 prepared with the addition of polyvinylpyrrolidone (PVP) (inset: the corresponding SAED pattern).

As a typical ionic surfactant, SDBS always takes effect in the form of electrostatic stabilization. According to “DLVO theory” (25), the colloidal particles with negative charges from the sulfonic acid groups in SDBS can be formed. When MnCl2·4H2O solution was added as a reductive agent into SDBS solution, electrostatic attraction of the colloidal particles with Mn2+ ions with positive charges occurred, thus forming a stable double electrical layer. Then, as a “micro reactor”, it could react with oxidizer and generate nanostructured MnO2. Under the action of electrostatic repulsive force, it was difficult for MnO2 particles to agglomerate together. However, PEG and PVP, as non-ionic surfactants, could also produce steric stabilization besides the effect of electrostatic stabilization, and further inhibited the growth of grains, leading to further optimization of the morphology and microstructure.

Electrochemical properties

Cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS) and galvanostatic charge/discharge tests are considered as important tools to study the electrochemical properties of electrode materials. Figure 5 shows the CV curves of the samples with the addition of different surfactants measured in 6 M KOH solution. As can be seen, there are no apparent redox peaks in the CV curves, which are roughly rectangular at lower scan rate. This indicated that the four electrode materials all had good capacitance and fast charge/discharge property. However, with the increase of scan rate, the CV curves deviated from the roughly rectangular shape to a spindle shape. With the increase of scan speed, electrochemical polarization and concentration polarization occurred in the solution, so that the electrolyte ions could only reach the outer surface of the MnO2 electrode during extremely brief time. As a result, the effective interaction between the electrolyte and active material of electrode was greatly reduced, which caused the limited utilization rate of active material and poor electrochemical behavior. From the CV curves, the areas of the samples follow the sequence of Sample d > Sample b > Sample c > Sample a. Furthermore, compared with Sample a, the CV curves of other samples (Samples b, c and d) are excellently symmetrical with the zero-current line, and have less deformation with the increase of the scanning rate. The experimental results showed that the reversibility and rate performance of the electrode materials could be improved by adding surfactants.

Cyclic voltammograms of the samples: (A) without surfactants, (B) with the addition of polyethylene glycol (PEG), (C) with the addition of sodium dodecylbenzenesulfonate (SDBS), and (D) with the addition of polyvinylpyrrolidone (PVP).

In order to investigate the influence of different surfactants on the alternating current (AC) impedance, the EIS of the samples were tested. Figure S2 (available online as supplementary material at www.jab-fm.com) shows the Nyquist plots of different samples. All Nyquist plots show an approximate semicircle in the high frequency range and exhibit a linear characteristic in the low frequency range. Compared with Sample a, the radius of the semicircle decreases gradually while the slope of the straight line increases gradually following the order of Sample c, Sample b and Sample d. The results confirmed the decrease of the electrochemical reaction impedance of the electrode, which corresponded to the increase of the ion diffusion ability in the solution (8, 11). The reason can be attributed to the fact that adding surfactants produced particles of smaller size, which increased the specific surface area and inhibited the agglomeration of nanoparticles.

Figure 6 shows the galvanostatic charge/discharge curves of different samples. As can be seen, all the charge/discharge curves show a good linear characteristic and exhibit an approximate isosceles triangle shape. This indicated that the prepared electrode materials had ideal capacitance characteristics and good reversibility. The following formula can be used to calculate specific capacitance:

Galvanostatic charge/discharge curves of the samples: (A) without surfactants, (B) with the addition of polyethylene glycol (PEG), (C) with the addition of SDBS, and (D) with the addition of polyvinylpyrrolidone (PVP).

C m = i × Δ t m × Δ V

where Cm represents the specific capacitance of the electrode material (F/g); i is the current of discharge (A); Δt is the time of discharge (s); m is the mass of active materials in the working electrode (g); and ΔV is the voltage drop of discharge (V).

According to the formula, the specific capacitances of Samples a, b, c and d at 120 mA/g are 155.9 F/g, 207.9 F/g, 187.5 F/g and 238.7 F/g, respectively. The results indicated that during the preparation process of MnO2, adding surfactants could effectively improve their electrochemical performance. In particular, the samples obtained from nonionic surfactants (PEG and PVP) showed better performance than that from ionic surfactant (SDBS). Since nonionic surfactants played dual roles of electrostatic stabilization and steric stabilization (26, 27), the performance improvement may be explained by the following reasons. PVP is a kind of hyperdispersant, and its molecular structure includes two parts: one is the typical anchoring group and the other is the solvent chain. When it was added as a surface dispersant, on the one hand, the multivariate alkyl residue and cyano group (—NCO) as anchoring group in PVP structure could be tightly adsorbed on the surface of particles by chemical bond, hydrogen bond or ionic bond, and played a role of electrostatic stabilization. On the other hand, the lateral polarity groups of pentabasic ring could act as soluble chain, which fully stretched in the solution and formed a steric hindrance layer with an adequate thickness on the surface of particles. Thus, the steric hindrance effect could be produced, which hindered the flocculation and agglomeration of nanoparticles.

In order to further study the stability of Sample d, cyclic charge/discharge tests were carried out at 120 mA/g. Figure 7 shows the specific capacitance of the electrode material keeps stable at the initial stage, and still remains 91.1% of the initial value or 217.46 F/g after 500 cycles. This further confirmed that the prepared amorphous hydrated MnO2was a good electrode material for supercapacitors.

Cyclic performance of MnO2 prepared by adding polyvinylpyrrolidone (PVP) at the current density of 120 mA/g (inset: charge/discharge curves in the cyclic performance test).

Conclusions

Amorphous hydrated MnO2 was prepared for the application as supercapacitor electrode materials by liquid co-precipitation in the presence of different surfactants. The results indicated that during the process of preparing MnO2, adding surfactants (PEG, SDBS and PVP) had significant effects on the morphology and particle size of MnO2 as well as its electrochemical performance. The specific surface areas of the samples with the addition of PEG, SDBS and PVP were 169.92 m2/g, 137.40 m2/g and 196.64 m2/g, respectively, and the corresponding capacitances were 207.9 F/g, 187.5 F/g and 238.7 F/g. Compared with the sample prepared without surfactants, the specific surface area and capacitance of the sample prepared with PVP were improved by 92.2% and 53.1%, respectively. Moreover, the electrode prepared with PVP showed good cycle stability at the current density of 120 mA/g, and 91.1% of its specific capacitance still remained after 500 cycles. It was believed that this material was a good candidate for supercapacitor electrode materials.

Disclosures

Financial support: This work was financially supported by National Natural Science Foundation of China (21676040, 21276036), the National Key Research and Development Program of China (2016yfb0101200, 2016yfb0101206), the Liaoning Provincial Nature Science Foundation of China (2014025018), and the Fundamental Research Funds for the Central Universities (3132016341), the Foundation of Liaoning Educational Committee (L2014199), the Shanghai Key Laboratory of Digitalized Manufacturing of Complex Sheet Metal Structures (No. 2014004).
Conflict of interest: None of the authors has financial interest related to this study to disclose.
References
  • 1. Yuan GH Electrochemical capacitor. 3rd ed. Beijing,PA: Chemical Industry Press 2006 1 10 Google Scholar
  • 2. Zhang ZA Yang BC Deng MG Hu YD Wang BH Synthesis and characterization of nanostructured MnO2 for supercapacitor. Acta Chimi Sin 2004 62 17 1617 1620 Google Scholar
  • 3. Conway BE Birss V Wojtowicz J The role and utilization of pseudocapacitance for energy storage by supercapacitors. J Power Sources 1997 66 1-2 1 14 Google Scholar
  • 4. Lin C Ritter JA Popov BN Characterization of sol-gel-derived cobalt oxide xerogels as electrochemical capacitors. J Electrochem Soc 1998 145 12 4097 4103 Google Scholar
  • 5. Venkat S Weidner JW An electrochemical route for making porous nickel oxide electrochmical capacitors. Electrochem Soc 1997 144 8 210 213 Google Scholar
  • 6. Pang SC Aederson MA Chapman TW Novel electrode materials thin-film ultracapacitors: comparison of sol-gel-derived and electrodeposited manganese dioxide. Electrochem Soc 2000 147 2 444 450 Google Scholar
  • 7. Yu M Zhai T Lu X et al. Manganese dioxide nanorod arrays on carbon fabric for flexible solid-state supercapacitors. J Power Sources 2013 239 64 71 Google Scholar
  • 8. Zhai T Xie S Yu M et al. Oxygen vacancies enhancing capacitive properties of MnO2 nanorods for wearable asymmetric supercapacitors. Nano Energy 2014 8 255 263 Google Scholar
  • 9. Xiao X Li T Yang P et al. Fiber-based all-solid-state flexible supercapacitors for self-powered systems. ACS Nano 2012 6 10 9200 9206 Google Scholar
  • 10. Yuan L Lu XH Xiao X et al. Flexible solid-state supercapacitors based on carbon nanoparticles/MnO2 nanorods hybrid structure. ACS Nano 2012 6 1 656 661 Google Scholar
  • 11. Gao Y Lin Y Chen J et al. Three-dimensional nanotube electrode arrays for hierarchical tubular structured high-performance pseudocapacitors. Nanoscale 2016 8 27 13280 13287 Google Scholar
  • 12. Gao Y Jin H Lin Q et al. Highly flexible and transferable supercapacitors with ordered three-dimensional MnO2/Au/MnO2 nanospikes arrays. J Mater Chem A 2015 3 19 10199 10204 Google Scholar
  • 13. Hu P Yan M Wang X et al. Single-nanowire electrochemical probe detection for internally optimized mechanism of porous graphene in electrochemical devices. Nano Lett 2016 16 3 1523 1529 Google Scholar
  • 14. Mai L Dong F Xu X et al. Cucumber-like V2O5/poly(3,4-ethylenedioxythiophene) & MnO2 nanowires with enhanced electrochemical cyclability. Nano Lett 2013 13 2 740 745 Google Scholar
  • 15. Parida KM Kanungo SB Sant BR Studies on chemically precipitated Mn(IV) oxides. Electrochim Acta 1981 26 2 435 443 Google Scholar
  • 16. Hong MS Lee SH Kim SW Use of KCl aqueous electrolyte for 2 V manganese oxide/activated carbon hybrid capacitor. Electrochem Solid-State Lett 2002 5 10 A227 Google Scholar
  • 17. Li SG Preparation of ultrafine oxide powder by chemical liquid phase coprecipitation. Rare Metals and Cemented Carbides 1998 03 42 43 Google Scholar
  • 18. Zhang H Wang Y Liu C Jiang H Influence of surfactant CTAB on the electrochemical performance of manganese dioxide used as supercapacitor electrode material. Alloys Compd 2012 517 3 1 8 Google Scholar
  • 19. Jiang RR Huang T Liu JL Zhuang JH Yu AS A novel method to prepare nanostructured manganese dioxide and its electrochemical properties as a supercapacitor electrode. Electrochim Acta 2009 54 11 3047 3052 Google Scholar
  • 20. Senthilkumar B Thenamirtham P Selvan RK Structural and electrochemical properties of polythiophene. Appl Surf Sci 2011 257 21 9063 9067 Google Scholar
  • 21. Mi J Wang YT Gao PC et al. Effects of thermal treatment on the electrochemical behavior of manganese dioxide. Acta Phys Chim Sin 2011 27 4 893 899 Google Scholar
  • 22. Zhang ZA Yang BC Deng MG Hu YD Amorphous MnO2 as electrode material for supercapacitor. Funct Mater Devices 2005 11 1 58 62 Google Scholar
  • 23. Lee HY Goodenough JB Supercapacitor behavior with KCl electrolyte. J Solid State Chem 1999 144 1 220 223 Google Scholar
  • 24. Song JM Zhang YX Zhu SQ Wang YH Piao CA Application of surfactant in battery material. Mod Chem Ind 2011 31 1 28 31 Google Scholar
  • 25. Zheng Z Hu JH Stability of colloid. 1st edition. Guang zhou, PA: Guangzhou Science & Technology Press 1993 12 22 Google Scholar
  • 26. Zhou TS Li QH Zheng KY Zhu J Dispersant and its application in the preparation of ceramics. Journal of Hubei University 2001 23 4 331 333 Natural Science. Google Scholar
  • 27. Cui HM Liu H Wang JY Li X Han F Agglomeration and dispersion of nano-scale powders. Materials for Mechanical Engineering 2004 28 8 38 41 Google Scholar

Authors

Affiliations

  • Transportation Equipment and Ocean Engineering College, Dalian Maritime University, Dalian - China

Article usage statistics

The blue line displays unique views in the time frame indicated.
The yellow line displays unique downloads.
Views and downloads are counted only once per session.

This article has supplementary materials available to download.