Next Article in Journal
Particle-Reinforced Ceramic Matrix Composites—Selected Examples
Previous Article in Journal
Developing Wound Moisture Sensors: Opportunities and Challenges for Laser-Induced Graphene-Based Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of High Energy Ball Milling and Dispersant on Capacitive Properties of Fe2O3—Carbon Nanotube Composites

Department of Materials Science and Engineering, McMaster University, Hamilton, ON L8S 4L7, Canada
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2022, 6(6), 177; https://doi.org/10.3390/jcs6060177
Submission received: 25 May 2022 / Revised: 10 June 2022 / Accepted: 15 June 2022 / Published: 17 June 2022
(This article belongs to the Special Issue Carbon Composites for Energy Conversion and Storage)

Abstract

:
This investigation is motivated by increasing interest in ferrimagnetic materials and composites, which exhibit electrical capacitance. It addresses the need for the development of magnetic materials with enhanced capacitive properties and low electrical resistance. γ-Fe2O3-multiwalled carbon nanotube (MWCNT) composites are developed by colloidal processing and studied for energy storage in negative electrodes of supercapacitors. High energy ball milling (HEBM) of ferrimagnetic γ-Fe2O3 nanoparticles results in enhanced capacitive properties. The effect of HEBM on particle morphology is analyzed. Gallocyanine is used as a co-dispersant for γ-Fe2O3 and MWCNTs. The polyaromatic structure and catechol ligand of gallocyanine facilitated its adsorption on γ-Fe2O3 and MWCNTs, respectively, and facilitated their electrostatic dispersion and mixing. The adsorption mechanisms are discussed. The highest capacitance of 1.53 F·cm−2 is achieved in 0.5 M Na2SO4 electrolyte for composites, containing γ-Fe2O3, which is high energy ball milled and co-dispersed with MWCNTs using gallocyanine. HEBM and colloidal processing strategies allow high capacitance at low electrical resistance, which facilitates efficient charge–discharge. Obtained composites are promising for fabrication of multifunctional devices based on mutual interaction of ferrimagnetic and capacitive properties.

1. Introduction

Many recent investigations have focused on the design and optimization of processing techniques for the fabrication of advanced composites [1,2,3,4]. Significant interest has been generated in the development of colloidal techniques [5,6,7,8]. The design of electrostatic self-assembly methods resulted in the fabrication of novel nanocomposites with high performance [9,10,11,12]. Advanced nanocomposites have been developed, combining functional properties of inorganic materials, polymers and carbon nanotubes [13,14,15,16]. A technological problem in the development of such composites by colloidal techniques is related to dispersion of carbon nanotubes, and their efficient co-dispersion and mixing with other materials. Various techniques have been reported for the dispersion of carbon nanotubes [13,17,18,19]. However, their co-dispersion with metal oxide nanoparticles presents difficulties due to the lack of efficient co-dispersants. Therefore, there is a need for the development of co-dispersant molecules, which must be adsorbed on both materials and facilitate their mixing on the nanometric scale. It is challenging to co-disperse carbon nanotubes with nanoparticles of magnetic oxides, which are prone to agglomeration due to van der Waals and magnetic forces.
Another challenge is related to the development of advanced techniques for fabrication of nanoparticles. High energy ball milling (HEBM) is a promising technique for sustainable production of different nanoparticles, such as metals, alloys and oxides [20,21,22]. Particle shape and properties can be modified by HEBM [23]. Aluminum nanoparticles with enhanced reactivity were prepared by this method [24]. Moreover, Al-Ni mixed powders with enhanced reactivity were fabricated [25].
HEBM is under development for the mechanochemical synthesis of materials [26,27,28]. This technique offers advantages of waste-free eco-friendly processing [26]. Mechanochemical alloying is one of the promising applications of HEBM [29,30]. The use of HEBM allowed the fabrication of advanced composites with uniform dispersion of ceramic particles in a metal alloy matrix [31]. Such composites exhibited advanced mechanical properties [31]. Mg alloys showed significant enhancement of H-storage properties after HEBM [32]. This resulted from reduction in grain size and improved distribution of a catalyst material [32]. HEBM allowed modification of microstructure and properties of magnetic materials and facilitated the fabrication of advanced alloys with hard magnetic properties [33]. HEBM is a promising technique for the fabrication of particles for energy storage applications in supercapacitors and batteries [34,35,36]. The reduction in particle size allowed an improved rate performance of Li4Ti5O12 electrodes [34]. HEBM treatment of graphite created abundant oxygen functional groups on the particle surface [37] and supercapacitor devices with an enlarged voltage window and high capacitance were obtained. However, the effect of HEBM on Ni(OH)2 battery electrode material performance is not well understood. The structure changes in Ni(OH)2 electrodes showed deterioration of their electrochemical performance [38]. Such controversial and diverging results generate a need for further investigation of HEBM for fabrication of active materials for energy storage devices.
The goal of this investigation was fabrication and testing of Fe2O3–carbon nanotube composite electrodes for supercapacitors. The approach was based on synergy of two methods, colloidal processing and HEBM. An important finding was that gallocyanine is a promising co-dispersant which adsorbs on Fe2O3 and carbon nanotubes and allows for their efficient co-dispersion and mixing. We analyzed the influence of HEBM on morphology of Fe2O3 particles. Electrochemical testing results revealed benefits of HEBM and colloidal processing in the presence of gallocyanine, which facilitated the fabrication of the Fe2O3–carbon nanotube composite electrodes with enhanced capacitance.

2. Materials and Methods

Multiwalled carbon nanotubes (MWCNTs, OD 13 nm, ID 4 nm, length 1–2 μm, Bayer, Leverkusen, Germany), Ni foams (thickness 1.6 mm, 95% porosity, Vale, Mississauga, Canada), Fe2O3 nanopowder (size < 50 nm), Na2SO4, poly(vinyl butyral) (PVB), gallocyanine (MilliporeSigma, Oakville, Canada) were used. The mass ratio Fe2O3:MWCNT:PVB was 80:20:3 in composites 1–3. The composite electrodes were prepared by impregnation of Ni foam current collectors with mixed Fe2O3 and MWCNT slurries in ethanol, containing dissolved PVB. The slurries were ultrasonicated for 30 min before impregnation of the current collectors. As-received Fe2O3 powder was used for the preparation of composite 1, whereas high energy ball milled Fe2O3 was utilized for composites 2 and 3. Composite 2 was prepared without gallocyanine, whereas gallocyanine was used for preparation of composite 3 using a preliminary co-dispersion procedure, described below. High energy ball milling was performed using a Mixer Mill MM 500 Nano (Retsch GmbH, Haan, Germany) at a frequency of 15 Hz. Milling time was 30 min. A preliminary co-dispersion and mixing procedure was used for the preparation of composite 3. In this preliminary procedure, the ball milled Fe2O3 powder was mixed with MWCNTs in ethanol, containing dissolved gallocyanine. The mass ratio of gallocyanine:Fe2O3 was 0.15. After ultrasonication for 30 min, filtering for removal of non-adsorbed gallocyanine and drying, the obtained mixture was used for preparation of a slurry in ethanol, containing PVB for impregnation of the current collectors. The impregnated electrodes were dried in air at 60 °C and pressed to 30% of original thickness. The total mass of impregnated material after drying (mass loading) was 40 mg·cm−2.
Electron microscopy investigations were performed using a JSM-7000F, (JEOL, Tokyo, Japan) scanning electron microscope (SEM) and Talos 200 (Thermo Fisher Scientific, Waltham, MA, USA) transmission electron microscope (TEM). X-ray diffraction (XRD) analysis (diffractometer Bruker D8, Bruker, Billerica, MA, USA) was performed at the rate of 0.01 degrees per second using Cu-Kα radiation. Fourier transform infrared spectroscopy (FTIR) studies were performed using a Bruker Vertex 70 spectrometer (Bruker, Billerica, MA, USA). The IR spectra were registered between 400 and 4000 cm−1 as a result of the average of 3 scans with a resolution of 0.5 cm−1. Zeta potential measurements were performed by a mass transfer method [39].
The capacitive behavior of the electrodes was analyzed in aqueous 0.5 M Na2SO4 electrolyte using a PARSTAT 2273 (Ametek, Berwyn, PA, USA) potentiostat for cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS). A BioLogic VMP 300 potentiostat (Biologic, Willow Hill, IL, USA) was used for galvanostatic charge–discharge (GCD) investigations. A 3-electrode electrochemical cell was used, which contained a working electrode, counter-electrode (Pt mesh) and a reference electrode (SCE, saturated calomel electrode). Gravimetric (Cm, F·g−1) and areal (CS, F·cm−2) capacitances were calculated from the CV, EIS and GCD data as was described in previous investigations [40,41].
The capacitance was obtained from CV data using the following equation:
C = Δ Q Δ U = | 0 t ( U max ) Idt | + | t ( U max ) 0 Idt | 2 U max
where ΔQ is charge, I—current, t—time and ΔU—the potential range. GCD data were used for the calculation of capacitance by the following equation:
C = IΔt/ΔU.
The complex capacitance C*(ω) = C′(ω) − iC″(ω) was calculated at different frequencies (ω) from the complex impedance Z*(ω) =Z′(ω) + iZ″(ω) data:
C ( ω ) =   Z ( ω ) ω | Z ( ω ) | 2 ,
C ( ω ) =   Z ( ω ) ω | Z ( ω ) | 2 .
CS and Cm obtained from CV and GCD data in a potential window of 0–0.9 V represented integral capacitances. EIS data obtained in an open circuit potential at voltage amplitude of 5 mV provided differential capacitances. The EIS spectra were acquired in the frequency range of 10 Hz–1 kHz.

3. Results

Figure 1 shows the X-ray diffraction pattern of as-received Fe2O3 powder. The diffraction pattern shows peaks of γ-Fe2O3 phase. Labeled peaks in Figure 1 correspond to JCPDS file 025-1402 of γ-Fe2O3.
Figure 2 shows TEM images of as-received and HEBM nanoparticles at different magnifications. The particle size of as-received Fe2O3 powder was typically below 50 nm, is agreement with the information provided by the powder manufacturer. However, TEM observations revealed a small number of larger spherical particles with a particle size of 200–250 nm (Figure 2A). Such particles were not observed after HEBM (Figure 2B). Therefore, HEBM eliminated such relatively large particles. The TEM images at a higher magnification (Figure 2C,D) did not show a significant difference in the size of nanoparticles before and after HEBM. It is in this regard that HEBM over 90 h resulted in a similar size of Fe2O3 nanoparticles [42]. However, TEM analysis of HEBM nanoparticles showed significant particle grinding, which resulted in a rough particle surface (Figure 3).
It is known that γ-Fe2O3 phase (maghemite) is a ferrimagnetic material with relatively high spontaneous magnetization [43,44]. The fabrication of stable suspensions of such nanoparticles for colloidal processing presents difficulties due to van der Waals and magnetic interactions, which promote particle aggregation and sedimentation. Another problem is related to poor electronic conductivity of γ-Fe2O3, which is detrimental for applications of this material in supercapacitors.
In this investigation, gallocyanine was used as a dispersion agent for Fe2O3. A sedimentation test of as-received Fe2O3 powder in ethanol showed fast precipitation. The addition of gallocyanine with mass ratio gallocyanine: Fe2O3 = 0.15 resulted in colloidal stability for 2 days for as-received Fe2O3 and more than 10 days for HEBM Fe2O3. It was hypothesized that gallocyanine adsorbed on Fe3O4 particles and provided electrostatic stabilization. Figure 4A shows the chemical structure of gallocyanine. It contains a catechol ligand, which can facilitate gallocyanine adsorption on inorganic surfaces (Figure 4B).
Catecholate molecules are gaining attention for surface modification and dispersion of different materials. One of the greatest drivers of catecholate dispersant development is investigation of strong mussel adsorption on different surfaces, which is based on bidentate bonding of catechol ligands of mussel adhesive proteins [45,46,47]. Recent studies of monoaromatic charged molecules from the catechol family led to the development of advanced dispersants [48]. Caffeic acid, tiron, dopamine and other cationic and anionic catecholates were used for dispersion of ZrO2, TiO2, MnO2, ZnO and other metal oxides in ethanol [48,49]. Adsorption of caffeic acid on positively charged MnO2 particles in ethanol resulted in charge reversal [50]. Therefore, it was not surprising that gallocyanine adsorbed on Fe2O3 and facilitated particle dispersion. The gallocyanine adsorption on Fe2O3 nanoparticles was confirmed by FTIR studies (Figure 5).
The FTIR spectrum of as-received Fe2O3 showed absorption at 1404 cm−1, which can result from vibrations of adsorbed CO2 [51]. The spectrum of gallocyanine showed absorption at 1722 cm−1 related to C=O stretch [52]. Absorptions at 1591 and 1555 cm−1 are related to C=C–C stretch [18,19]. Absorptions at 1385, 1312 and 1119 cm−1 are related to C─O, C–N–C and C–O–C stretches, respectively [18,52]. The spectrum of Fe2O3 dispersed using gallocyanine showed similar peaks and confirmed gallocyanine adsorption on Fe2O3 particles. The spectra of pure Fe2O3 and Fe2O3 dispersed using gallocyanine (Supplementary Information, Figure S1) showed broad absorptions at 550 cm−1 related to Fe–O stretching [53].
The larger size of gallocyanine, compared to the size of monoaromatic catecholates, such as caffeic acid, tiron and dopamine, is beneficial for dispersion of Fe2O3. Figure 4 indicates that gallocyanine is a cationic molecule. Therefore, gallocyanine was used as a cationic dispersant for Fe2O3. Electrophoretic experiments in an electric field of 30 V cm−1 for Fe2O3 suspensions, containing gallocyanine, confirmed that Fe2O3 particles were positively charged, as they moved toward the cathode under the influence of the electric field and deposited electrophoretically on the cathode surface. Zeta potential of Fe2O3 particles was found to be +8.2 mV. Figure 4 indicates that gallocyanine is a polyaromatic molecule. It was found that gallocyanine allowed for efficient dispersion of MWCNTs in ethanol. The adsorption mechanism involved π–π interactions of polyaromatic gallocyanine with side walls of MWCNTs. The MWCNT suspensions prepared without gallocyanine were unstable and showed rapid precipitation. In contrast, the MWCNT suspension containing gallocyanine showed colloidal stability for more than 2 weeks. Therefore, the adsorbed gallocyanine provided electrostatic dispersion of MWCNTs and electrostatic co-dispersion of Fe2O3 and MWCNTs in mixed suspensions.
Figure 6 shows SEM images of composites 1–3 impregnated into Ni foam current collectors. Composite 1 contained mainly nanoparticles of Fe2O3 and a small number (about 1%) of spherical particles with a typical size of 200–250 nm in agreement with the results of TEM studies (Figure 6). HEBM resulted in crashing and elimination of such particles. The SEM images of composites 2 and 3 showed only nanoparticles of Fe2O3, which were mixed with MWCNTs (Figure 6B,C). The porosity of the electrode materials was beneficial for electrolyte access to the active material.
The electrochemical performance of the composite electrodes was tested in 0.5 M Na2SO4 electrolyte. Figure 7A–C show CVs for composites 1–3.
The CVs for composites 1–3 showed pseudocapacitive behavior of the electrodes. Composite 2 showed improved shape of CVs, which were nearly of rectangular shape. Composite 3 showed enlarged CV area, which indicated higher capacitance. The capacitances calculated from the CV data were 0.59 F·cm−2 (14.75 F·g−1), 0.88 F·cm−2 (22.0 F·g−1) and 1.53 F·cm−2 (38.25 F·g−1) at a scan rate of 2 mV·s−1. The higher capacitance of composite 2 compared to composite 1 can be attributed to effect of HEBM, such as crushing of large particles and increase in surface roughness of nanoparticles. It should be noted that capacitance increase cannot be correlated with surface area. Previous investigations did not show correlations between surface area and capacitance of different pseudocapacitive oxides [54,55,56,57]. Composite 3 showed the highest capacitance due to the beneficial effect of HEBM and enhanced mixing of Fe2O3 and MWCNT, which was achieved using gallocyanine as a co-dispersant for colloidal processing. The observed capacitances were mainly attributed to pseudocapacitive properties of Fe2O3 due to low capacitance of MWCNTs and low (20%) content of MWCNTs in the composites. In our approach, MWCNTs were used as conductive additives, which improved electronic conductivity of the composites. We found that the mass ratio of conductive MWCNTs must be 20% of the total mass of Fe2O3 and MWCNTs in order to achieve low impedance and high capacitance. Further increase in the MWCNT content in the composite resulted in reduced capacitance.
Figure 8 shows electrochemical impedance spectroscopy testing results. Composites 1–3 showed a relatively low real part of impedance, which indicated low electrode resistance. Composites 2 and 3 showed higher capacitance, compared to composite 1 at low frequencies, in agreement with CV data. However, composite 2 showed higher capacitance, compared to composite 3. The capacitances obtained from the impedance data were lower than capacitances derived from CVs. The difference may have resulted from different measurement conditions. The AC capacitance represents a differential capacitance measured at voltage amplitude of 5 mV. The differential capacitance depends on electrode potential. In contrast, the capacitance calculated from the CV data represents an integral capacitance, measured in a voltage window of 0.8V.
Figure 9A–C shows GCD data for composites 1–3 at different current densities. The charge–discharge curves had a nearly triangular shape. The charge–discharge time increased with decreasing current density. Composite 3 showed longer discharge times, compared to composites 1 and 2 at fixed current densities. The capacitances of 0.65, 1.09 and 1.26 F·cm−2 were obtained at a current density of 3 mA cm−2 for composites 1, 2 and 3, respectively. The capacitances decreased with increasing current density (Figure 9D).
Areal capacitance is an important characteristic of electrodes with high active mass loading [40]. Table 1 compares areal capacitance of the composite 3 electrode with literature data on areal capacitance of Fe2O3 electrodes, containing conductive additives. It is seen that previous investigations of negative electrodes based on Fe2O3 in Na2SO4 electrolyte were mainly focused on testing of α-Fe2O3, which is an antiferromagnetic material. In contrast, γ-Fe2O3 is a ferrimagnetic material, which showed relatively high capacitance (Table 1). Despite the difficulties related to colloidal processing of ferrimagnetic γ-Fe2O3 particles, we achieved higher capacitance of γ-Fe2O3-based composites. This resulted from the effect of HEBM and the use of gallocyanine as a co-dispersant.
The combination of ferrimagnetic and capacitive properties in a single material is very promising, because reduction of Fe3+ ions results in changing of their magnetic moments. Therefore, interesting effects can potentially be observed in such materials due to the influence of the charging process on magnetization or the influence of magnetic field on the charging process. Pseudocapacitive ferrimagnetics or ferromagnetics represent an important alternative to magnetically ordered ferroelectrics [58], which exhibit interesting magnetoelectric effects. It is in this regard that capacitance of γ-Fe2O3 is significantly higher than that of magnetically ordered ferroelectrics [58]. Moreover, magnetically ordered ferroelectrics exhibit antiferromagnetic or weak ferri/ferromagnetic properties. In contrast, γ-Fe2O3 is a ferrimagnetic material, which has a high magnetization. Therefore, this material, which exhibits advanced capacitive and magnetic properties, is promising for potential applications in multifunctional devices based on mutual interactions of magnetic and capacitive properties.
Table 1. Capacitance of Fe2O3-based composites containing conductive carbon additives in Na2SO4 electrolyte in negative potential range.
Table 1. Capacitance of Fe2O3-based composites containing conductive carbon additives in Na2SO4 electrolyte in negative potential range.
MaterialMass LoadingVoltage WindowCSReference
mg·cm−2VmF cm−2
α-Fe2O3/graphene1.0−1.0–0.0286[59]
vs. SCEat 2 mV·s−1
α-Fe2O3/carbon2.4−1.0–0.0862[60]
vs. SCEat 1 mA·cm−2
α-Fe2O3/carbon1.1−1.0–0.0430.8[61]
nanoarray vs. Ag/AgClat 1 mA·cm−2
α-Fe2O3/carbonN/A−0.8–0.0340[62]
vs. Ag/AgClat 1 mA·cm−2
γ-Fe2O3/carbon40−0.8–0.01530this work
vs. SCEat 2 mV·s−1

4. Conclusions

Composites of γ-Fe2O3 and MWCNTs with active mass loading of 40 mg·cm−2 were prepared by colloidal processing for energy storage in negative electrodes of supercapacitors. The composites combined advanced ferrimagnetic properties with high capacitance. The high capacitance was achieved at a low electrical resistance, which is beneficial for efficient charge–discharge. HEBM exerted influence on powder morphology and facilitated the fabrication of composites with enhanced capacitance. Gallocyanine allowed efficient co-dispersion of γ-Fe2O3 and MWCNTs. The catechol group of this dispersant facilitated its adsorption on γ-Fe2O3, whereas the polyaromatic structure allowed adsorption on MWCNTs. The electrostatic co-dispersion allowed for efficient mixing of individual components of the composite material. The highest capacitance of 1.53 F·cm−2 was achieved in Na2SO4 electrolyte for composites, containing γ-Fe2O3, which was high energy ball milled and co-dispersed with MWCNTs using gallocyanine. Obtained composites are promising for fabrication of multifunctional devices based on mutual interaction of ferrimagnetic and capacitive properties.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jcs6060177/s1, Figure S1: FTIR spectra of (a) as-received Fe2O3, (b) gallocyanine, (c) Fe2O3, containing adsorbed gallocyanine.

Author Contributions

Conceptualization, C.Z. and I.Z.; methodology, C.Z.; software, C.Z.; validation, C.Z. and I.Z.; formal analysis, C.Z.; investigation, C.Z.; resources, I.Z.; data curation, C.Z.; writing—original draft preparation, C.Z. and I.Z.; writing—review and editing, C.Z. and I.Z.; visualization, C.Z. and I.Z.; supervision, I.Z.; project administration, I.Z.; funding acquisition, I.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Sciences and Engineering Research Council of Canada, grant number RGPIN-2018-04014, and CRC program.

Data Availability Statement

Data contained within the article and Supplementary Information are available.

Acknowledgments

Electron microscopy investigations were performed at the Canadian Centre for Electron Microscopy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, S.-J.; Mose, M.P.; Kannaiyan, S. Artificial Intelligence Application in Solid State Mg-Based Hydrogen Energy Storage. J. Compos. Sci. 2021, 5, 145. [Google Scholar] [CrossRef]
  2. Kausar, A. Green Nanocomposites for Energy Storage. J. Compos. Sci. 2021, 5, 202. [Google Scholar] [CrossRef]
  3. Wang, Y.; Wang, Y.; Wang, C.; Wang, Y. A Hierarchical Architecture of Functionalized Polyaniline/Manganese Dioxide Composite with Stable-Enhanced Electrochemical Performance. J. Compos. Sci. 2021, 5, 129. [Google Scholar] [CrossRef]
  4. Bharti; Ahmed, G.; Kumar, Y.; Bocchetta, P.; Sharma, S. Determination of Quantum Capacitance of Niobium Nitrides Nb2N and Nb4N3 for Supercapacitor Applications. J. Compos. Sci. 2021, 5, 85. [Google Scholar] [CrossRef]
  5. Sorkhi, L.; Farrokhi-Rad, M.; Shahrabi, T. Electrophoretic Deposition of Hydroxyapatite–Chitosan–Titania on Stainless Steel 316 L. Surfaces 2019, 2, 458–467. [Google Scholar] [CrossRef] [Green Version]
  6. Song, Y.; Zhao, R.; Zhang, K.; Ding, J.; Lv, X.; Chen, M.; Xie, J. Facile synthesis of Mn3O4/double-walled carbon nanotube nanocomposites and its excellent supercapacitive behavior. Electrochim. Acta 2017, 230, 350–357. [Google Scholar] [CrossRef]
  7. Zhitomirsky, I.; Petric, A. Electrochemical deposition of yttrium oxide. J. Mater. Chem. 2000, 10, 1215–1218. [Google Scholar] [CrossRef]
  8. Pang, X.; Zhitomirsky, I.; Niewczas, M. Cathodic electrolytic deposition of zirconia films. Surf. Coat. Technol. 2005, 195, 138–146. [Google Scholar] [CrossRef]
  9. Beladi-Mousavi, S.M.; Sadaf, S.; Hennecke, A.K.; Klein, J.; Mahmood, A.M.; Rüttiger, C.; Gallei, M.; Fu, F.; Fouquet, E.; Ruiz, J. The Metallocene Battery: Ultrafast Electron Transfer Self Exchange Rate Accompanied by a Harmonic Height Breathing. Angew. Chem. 2021, 133, 13666–13670. [Google Scholar] [CrossRef]
  10. Beladi-Mousavi, S.M.; Sadaf, S.; Mahmood, A.M.; Walder, L. High performance poly (viologen)–graphene nanocomposite battery materials with puff paste architecture. ACS Nano 2017, 11, 8730–8740. [Google Scholar] [CrossRef]
  11. Beladi-Mousavi, S.M.; Klein, J.; Ciobanu, M.; Sadaf, S.; Mahmood, A.M.; Walder, L. Flexible, Self-Supported Anode for Organic Batteries with a Matched Hierarchical Current Collector System for Boosted Current Density. Small 2021, 17, 2103885. [Google Scholar] [CrossRef]
  12. Zhang, H.; Xu, D.; Wang, L.; Ye, Z.; Chen, B.; Pei, L.; Wang, Z.; Cao, Z.; Shen, J.; Ye, M. A Polymer/Graphene Composite Cathode with Active Carbonyls and Secondary Amine Moieties for High-Performance Aqueous Zn-Organic Batteries Involving Dual-Ion Mechanism. Small 2021, 17, 2100902. [Google Scholar] [CrossRef]
  13. Kausar, A.; Bocchetta, P. Poly(methyl methacrylate) Nanocomposite Foams Reinforced with Carbon and Inorganic Nanoparticles—State-of-the-Art. J. Compos. Sci. 2022, 6, 129. [Google Scholar] [CrossRef]
  14. Rodriguez-Uicab, O.; Tayyarian, T.; Abot, J.L. Effect of Curing Temperature of Epoxy Matrix on the Electrical Response of Carbon Nanotube Yarn Monofilament Composites. J. Compos. Sci. 2022, 6, 43. [Google Scholar] [CrossRef]
  15. Lemartinel, A.; Castro, M.; Fouché, O.; De-Luca, J.-C.; Feller, J.-F. A Review of Nanocarbon-Based Solutions for the Structural Health Monitoring of Composite Parts Used in Renewable Energies. J. Compos. Sci. 2022, 6, 32. [Google Scholar] [CrossRef]
  16. Sikkema, R.; Baker, K.; Zhitomirsky, I. Electrophoretic deposition of polymers and proteins for biomedical applications. Adv. Colloid Interface Sci. 2020, 284, 102272. [Google Scholar] [CrossRef]
  17. Rajendran, D.; Ramalingame, R.; Adiraju, A.; Nouri, H.; Kanoun, O. Role of Solvent Polarity on Dispersion Quality and Stability of Functionalized Carbon Nanotubes. J. Compos. Sci. 2022, 6, 26. [Google Scholar] [CrossRef]
  18. Su, Y.; Zhitomirsky, I. Electrophoretic deposition of graphene, carbon nanotubes and composite films using methyl violet dye as a dispersing agent. Colloids Surf. A Physicochem. Eng. Asp. 2013, 436, 97–103. [Google Scholar] [CrossRef]
  19. Shi, K.; Zhitomirsky, I. Electrophoretic nanotechnology of graphene–carbon nanotube and graphene–polypyrrole nanofiber composites for electrochemical supercapacitors. J. Colloid Interface Sci. 2013, 407, 474–481. [Google Scholar] [CrossRef]
  20. Muñoz, J.E.; Cervantes, J.; Esparza, R.; Rosas, G. Iron nanoparticles produced by high-energy ball milling. J. Nanoparticle Res. 2007, 9, 945–950. [Google Scholar] [CrossRef]
  21. Indris, S.; Bork, D.; Heitjans, P. Nanocrystalline oxide ceramics prepared by high-energy ball milling. J. Mater. Synth. Processing 2000, 8, 245–250. [Google Scholar] [CrossRef]
  22. Helle, S.; Pedron, M.; Assouli, B.; Davis, B.; Guay, D.; Roué, L. Structure and high-temperature oxidation behaviour of Cu–Ni–Fe alloys prepared by high-energy ball milling for application as inert anodes in aluminium electrolysis. Corros. Sci. 2010, 52, 3348–3355. [Google Scholar] [CrossRef]
  23. Tsuzuki, T.; Schäffel, F.; Muroi, M.; McCormick, P.G. α-Fe2O3 nano-platelets prepared by mechanochemical/thermal processing. Powder Technol. 2011, 210, 198–202. [Google Scholar] [CrossRef]
  24. André, B.; Coulet, M.-V.; Esposito, P.-H.; Rufino, B.; Denoyel, R. High-energy ball milling to enhance the reactivity of aluminum nanopowders. Mater. Lett. 2013, 110, 108–110. [Google Scholar] [CrossRef] [Green Version]
  25. Rogachev, A.; Shkodich, N.; Vadchenko, S.; Baras, F.; Kovalev, D.Y.; Rouvimov, S.; Nepapushev, A.; Mukasyan, A. Influence of the high energy ball milling on structure and reactivity of the Ni+ Al powder mixture. J. Alloy. Compd. 2013, 577, 600–605. [Google Scholar] [CrossRef]
  26. Kumar, M.; Xiong, X.; Wan, Z.; Sun, Y.; Tsang, D.C.; Gupta, J.; Gao, B.; Cao, X.; Tang, J.; Ok, Y.S. Ball milling as a mechanochemical technology for fabrication of novel biochar nanomaterials. Bioresour. Technol. 2020, 312, 123613. [Google Scholar] [CrossRef]
  27. Kong, L.; Ma, J.; Zhu, W.; Tan, O. Preparation of Bi4Ti3O12 ceramics via a high-energy ball milling process. Mater. Lett. 2001, 51, 108–114. [Google Scholar] [CrossRef]
  28. Bid, S.; Pradhan, S. Preparation of zinc ferrite by high-energy ball-milling and microstructure characterization by Rietveld’s analysis. Mater. Chem. Phys. 2003, 82, 27–37. [Google Scholar] [CrossRef]
  29. Hadef, F. Effect of high-energy ball milling on structure and properties of some intermetallic alloys: A mini review. Metallogr. Microstruct. Anal. 2019, 8, 430–444. [Google Scholar] [CrossRef]
  30. Rodríguez, V.P.; Rojas-Ayala, C.; Medina, J.M.; Cabrera, P.P.; Quispe-Marcatoma, J.; Landauro, C.; Tapia, J.R.; Baggio-Saitovitch, E.; Passamani, E. Fe50Ni50 synthesized by high energy ball milling: A systematic study using X-ray diffraction, EXAFS and Mössbauer methods. Mater. Charact. 2019, 149, 249–254. [Google Scholar] [CrossRef]
  31. Cabeza, M.; Feijoo, I.; Merino, P.; Pena, G.; Pérez, M.; Cruz, S.; Rey, P. Effect of high energy ball milling on the morphology, microstructure and properties of nano-sized TiC particle-reinforced 6005A aluminium alloy matrix composite. Powder Technol. 2017, 321, 31–43. [Google Scholar] [CrossRef]
  32. Révész, Á.; Gajdics, M. Improved H-Storage Performance of Novel Mg-Based Nanocomposites Prepared by High-Energy Ball Milling: A Review. Energies 2021, 14, 6400. [Google Scholar] [CrossRef]
  33. Liu, W.; Yue, M.; Cui, B.; Hadjipanayis, G.C. Permanent magnetic nanoparticles and nanoflakes prepared by surfactant-assisted high-energy ball milling. Rev. Nanosci. Nanotechnol. 2014, 3, 259–275. [Google Scholar] [CrossRef]
  34. Wang, G.; Xu, J.; Wen, M.; Cai, R.; Ran, R.; Shao, Z. Influence of high-energy ball milling of precursor on the morphology and electrochemical performance of Li4Ti5O12–ball-milling time. Solid State Ion. 2008, 179, 946–950. [Google Scholar] [CrossRef]
  35. Nandhini, R.; Mini, P.; Avinash, B.; Nair, S.; Subramanian, K. Supercapacitor electrodes using nanoscale activated carbon from graphite by ball milling. Mater. Lett. 2012, 87, 165–168. [Google Scholar] [CrossRef]
  36. Kakaei, K.; Akbarpour, M.R.; Liyaghi, Z. Synthesis of partially oxidized NiTi nanostructure by high energy ball milling and its application to supercapacitor. J. Mater. Sci. Mater. Electron. 2018, 29, 2222–2227. [Google Scholar] [CrossRef]
  37. Wang, Y.; Cao, J.; Zhou, Y.; Ouyang, J.-H.; Jia, D.; Guo, L. Ball-milled graphite as an electrode material for high voltage supercapacitor in neutral aqueous electrolyte. J. Electrochem. Soc. 2012, 159, A579. [Google Scholar] [CrossRef]
  38. Chen, H.; Wang, J.; Pan, T.; Xiao, H.; Zhang, J.; Cao, C. Effects of high-energy ball milling (HEBM) on the structure and electrochemical performance of nickel hydroxide. Int. J. Hydrog. Energy 2003, 28, 119–124. [Google Scholar] [CrossRef]
  39. Hashiba, M.; Okamoto, H.; Nurishi, Y.; Hiramatsu, K. The zeta-potential measurement for concentrated aqueous suspension by improved electrophoretic mass transport apparatus—application to Al2O3, ZrO3 and SiC suspensions. J. Mater. Sci. 1988, 23, 2893–2896. [Google Scholar] [CrossRef]
  40. Chen, R.; Yu, M.; Sahu, R.P.; Puri, I.K.; Zhitomirsky, I. The Development of Pseudocapacitor Electrodes and Devices with High Active Mass Loading. Adv. Energy Mater. 2020, 10, 1903848. [Google Scholar] [CrossRef]
  41. Nawwar, M.; Poon, R.; Chen, R.; Sahu, R.P.; Puri, I.K.; Zhitomirsky, I. High areal capacitance of Fe3O4-decorated carbon nanotubes for supercapacitor electrodes. Carbon Energy 2019, 1, 124–133. [Google Scholar] [CrossRef] [Green Version]
  42. Wang, L.-L.; Jiang, J.-S. Preparation of α-Fe2O3 nanoparticles by high-energy ball milling. Phys. B Condens. Matter 2007, 390, 23–27. [Google Scholar] [CrossRef]
  43. Nazari, M.; Ghasemi, N.; Maddah, H.; Motlagh, M.M. Synthesis and characterization of maghemite nanopowders by chemical precipitation method. J. Nanostructure Chem. 2014, 4, 99. [Google Scholar] [CrossRef] [Green Version]
  44. Talgatov, E.T.; Auyezkhanova, A.S.; Seitkalieva, K.S.; Tumabayev, N.Z.; Akhmetova, S.N.; Zharmagambetova, A.K. Co-precipitation synthesis of mesoporous maghemite for catalysis application. J. Porous Mater. 2020, 27, 919–927. [Google Scholar] [CrossRef]
  45. Lee, B.P.; Messersmith, P.B.; Israelachvili, J.N.; Waite, J.H. Mussel-inspired adhesives and coatings. Annu. Rev. Mater. Res. 2011, 41, 99–132. [Google Scholar] [CrossRef] [Green Version]
  46. Lee, H.; Lee, B.P.; Messersmith, P.B. A reversible wet/dry adhesive inspired by mussels and geckos. Nature 2007, 448, 338–341. [Google Scholar] [CrossRef]
  47. Xu, N.; Li, Y.; Zheng, T.; Xiao, L.; Liu, Y.; Chen, S.; Zhang, D. A mussel-inspired strategy for CNT/carbon fiber reinforced epoxy composite by hierarchical surface modification. Colloids Surf. A Physicochem. Eng. Asp. 2022, 635, 128085. [Google Scholar] [CrossRef]
  48. Ata, M.; Liu, Y.; Zhitomirsky, I. A review of new methods of surface chemical modification, dispersion and electrophoretic deposition of metal oxide particles. Rsc. Adv. 2014, 4, 22716–22732. [Google Scholar] [CrossRef]
  49. Wu, K.; Wang, Y.; Zhitomirsky, I. Electrophoretic deposition of TiO2 and composite TiO2–MnO2 films using benzoic acid and phenolic molecules as charging additives. J. Colloid Interface Sci. 2010, 352, 371–378. [Google Scholar] [CrossRef]
  50. Wang, Y.; Zhitomirsky, I. Bio-inspired catechol chemistry for electrophoretic nanotechnology of oxide films. J. Colloid Interface Sci. 2012, 380, 8–15. [Google Scholar] [CrossRef]
  51. Chernyshova, I.V.; Ponnurangam, S.; Somasundaran, P. Linking interfacial chemistry of CO2 to surface structures of hydrated metal oxide nanoparticles: Hematite. Phys. Chem. Chem. Phys. 2013, 15, 6953–6964. [Google Scholar] [CrossRef] [PubMed]
  52. Ulusoy, H.İ.; Şimşek, S. Removal of uranyl ions in aquatic mediums by using a new material: Gallocyanine grafted hydrogel. J. Hazard. Mater. 2013, 254, 397–405. [Google Scholar] [CrossRef] [PubMed]
  53. Ahangaran, F.; Hassanzadeh, A.; Nouri, S. Surface modification of Fe3O4@ SiO2 microsphere by silane coupling agent. Int. Nano Lett. 2013, 3, 23. [Google Scholar] [CrossRef]
  54. Reddy, R.N.; Reddy, R.G. Sol–gel MnO2 as an electrode material for electrochemical capacitors. J. Power Sources 2003, 124, 330–337. [Google Scholar] [CrossRef]
  55. Jeong, Y.; Manthiram, A. Nanocrystalline manganese oxides for electrochemical capacitors with neutral electrolytes. J. Electrochem. Soc. 2002, 149, A1419. [Google Scholar] [CrossRef]
  56. Dong, W.; Rolison, D.R.; Dunn, B. Electrochemical properties of high surface area vanadium oxide aerogels. Electrochem. Solid State Lett. 2000, 3, 457. [Google Scholar] [CrossRef]
  57. Dong, W.; Sakamoto, J.S.; Dunn, B. Electrochemical properties of vanadium oxide aerogels. Sci. Technol. Adv. Mater. 2003, 4, 3–11. [Google Scholar] [CrossRef] [Green Version]
  58. Venevtsev, Y.N.; Gagulin, V.V.; Zhitomirsky, I.D. Material science aspects of seignette-magnetism problem. Ferroelectrics 1987, 73, 221–248. [Google Scholar] [CrossRef]
  59. Liu, Y.; Luo, D.; Shi, K.; Michaud, X.; Zhitomirsky, I. Asymmetric supercapacitor based on MnO2 and Fe2O3 nanotube active materials and graphene current collectors. Nano-Struct. Nano-Objects 2018, 15, 98–106. [Google Scholar] [CrossRef]
  60. Kang, M.; Zhou, S.; Zhang, J.; Ning, F.; Ma, C.; Qiu, Z. Facile fabrication of oxygen vacancy-rich α-Fe2O3 microspheres on carbon cloth as negative electrode for supercapacitors. Electrochim. Acta 2020, 338, 135820. [Google Scholar] [CrossRef]
  61. Chen, D.; Zhou, S.; Quan, H.; Zou, R.; Gao, W.; Luo, X.; Guo, L. Tetsubo-like α-Fe2O3/C nanoarrays on carbon cloth as negative electrode for high-performance asymmetric supercapacitors. Chem. Eng. J. 2018, 341, 102–111. [Google Scholar] [CrossRef]
  62. Liang, H.; Xia, C.; Emwas, A.-H.; Anjum, D.H.; Miao, X.; Alshareef, H.N. Phosphine plasma activation of α-Fe2O3 for high energy asymmetric supercapacitors. Nano Energy 2018, 49, 155–162. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction pattern of as received Fe2O3 powder (♦—peaks corresponding to JCPDS file 025-1402 of γ-Fe2O3 phase).
Figure 1. X-ray diffraction pattern of as received Fe2O3 powder (♦—peaks corresponding to JCPDS file 025-1402 of γ-Fe2O3 phase).
Jcs 06 00177 g001
Figure 2. TEM images (AD) at different magnifications for Fe2O3 nanoparticles: (A,C) before and (B,D) after HEBM. Arrow shows large particle in as-received Fe2O3.
Figure 2. TEM images (AD) at different magnifications for Fe2O3 nanoparticles: (A,C) before and (B,D) after HEBM. Arrow shows large particle in as-received Fe2O3.
Jcs 06 00177 g002
Figure 3. (A,B) High magnification TEM images of Fe2O3 nanoparticles after HEBM, which show increased surface roughness of the particles.
Figure 3. (A,B) High magnification TEM images of Fe2O3 nanoparticles after HEBM, which show increased surface roughness of the particles.
Jcs 06 00177 g003
Figure 4. (A) Chemical structure of gallocyanine, (B) adsorption of gallocyanine on Fe2O3 particle involving catecholate-type complexation of Fe atom on the particle surface.
Figure 4. (A) Chemical structure of gallocyanine, (B) adsorption of gallocyanine on Fe2O3 particle involving catecholate-type complexation of Fe atom on the particle surface.
Jcs 06 00177 g004
Figure 5. FTIR spectra of (a) as-received Fe2O3, (b) gallocyanine, (c) Fe2O3, containing adsorbed gallocyanine.
Figure 5. FTIR spectra of (a) as-received Fe2O3, (b) gallocyanine, (c) Fe2O3, containing adsorbed gallocyanine.
Jcs 06 00177 g005
Figure 6. SEM images of (A) composite 1, (B) composite 2 and (C) composite 3 impregnated into Ni foam current collectors. Arrows show large particles in composite 1. Composite 1 was prepared using as-received Fe2O3. HEBM was used for preparation of composites 2 and 3. Composite 2 was prepared without dispersants, composite 3 was prepared using gallocyanine as a co-dispersant.
Figure 6. SEM images of (A) composite 1, (B) composite 2 and (C) composite 3 impregnated into Ni foam current collectors. Arrows show large particles in composite 1. Composite 1 was prepared using as-received Fe2O3. HEBM was used for preparation of composites 2 and 3. Composite 2 was prepared without dispersants, composite 3 was prepared using gallocyanine as a co-dispersant.
Jcs 06 00177 g006
Figure 7. (AC) CVs for (A) composite 1, (B) composite 2 and (C) composite 3 at scan rates of (a) 5, (b) 10, (c) 20 mV·s−1, (D) capacitance versus scan rate for (a) composite 1, (b) composite 2, (c) composite 3.
Figure 7. (AC) CVs for (A) composite 1, (B) composite 2 and (C) composite 3 at scan rates of (a) 5, (b) 10, (c) 20 mV·s−1, (D) capacitance versus scan rate for (a) composite 1, (b) composite 2, (c) composite 3.
Jcs 06 00177 g007
Figure 8. (A) Nyquist plot of Z*, (B,C) frequency dependences of (B) Cs′ and (C) Cs″ for (a) composite 1, (b) composite 2 and (c) composite 3.
Figure 8. (A) Nyquist plot of Z*, (B,C) frequency dependences of (B) Cs′ and (C) Cs″ for (a) composite 1, (b) composite 2 and (c) composite 3.
Jcs 06 00177 g008
Figure 9. (AC) GCD data for (A) composite 1, (B) composite 2 and (C) composite 3 at current densities of (a) 3, (b) 5, (c) 7 and (d) 10 mA.cm−2 and (D) capacitance versus current density for (a) composite 1, (b) composite 2 and (c) composite 3.
Figure 9. (AC) GCD data for (A) composite 1, (B) composite 2 and (C) composite 3 at current densities of (a) 3, (b) 5, (c) 7 and (d) 10 mA.cm−2 and (D) capacitance versus current density for (a) composite 1, (b) composite 2 and (c) composite 3.
Jcs 06 00177 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, C.; Zhitomirsky, I. Influence of High Energy Ball Milling and Dispersant on Capacitive Properties of Fe2O3—Carbon Nanotube Composites. J. Compos. Sci. 2022, 6, 177. https://doi.org/10.3390/jcs6060177

AMA Style

Zhang C, Zhitomirsky I. Influence of High Energy Ball Milling and Dispersant on Capacitive Properties of Fe2O3—Carbon Nanotube Composites. Journal of Composites Science. 2022; 6(6):177. https://doi.org/10.3390/jcs6060177

Chicago/Turabian Style

Zhang, Chengwei, and Igor Zhitomirsky. 2022. "Influence of High Energy Ball Milling and Dispersant on Capacitive Properties of Fe2O3—Carbon Nanotube Composites" Journal of Composites Science 6, no. 6: 177. https://doi.org/10.3390/jcs6060177

Article Metrics

Back to TopTop