Next Article in Journal
Coal Properties and Coalbed Methane Potential in the Southern Part of the Upper Silesian Coal Basin, Poland
Previous Article in Journal
State-of-the-Art Review of Small Modular Reactors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enhancement of Electrochemical Performance of Aqueous Zinc Ion Batteries by Structural and Interfacial Design of MnO2 Cathodes: The Metal Ion Doping and Introduction of Conducting Polymers †

by
Mikhail A. Kamenskii
,
Filipp S. Volkov
,
Svetlana N. Eliseeva
,
Elena G. Tolstopyatova
* and
Veniamin V. Kondratiev
Institute of Chemistry, Saint Petersburg State University, 7/9 Universitetskaya nab, 199034 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Dedicated to the 300th Anniversary of Saint Petersburg University.
Energies 2023, 16(7), 3221; https://doi.org/10.3390/en16073221
Submission received: 28 February 2023 / Revised: 29 March 2023 / Accepted: 1 April 2023 / Published: 3 April 2023
(This article belongs to the Section D1: Advanced Energy Materials)

Abstract

:
Aqueous zinc-ion batteries (AZIBs) and, in particular, Zn//MnO2 rechargeable batteries have attracted great attention due to the abundant natural resources of zinc and manganese, low cost, environmental friendliness, and high operating voltage. Among the various ways to improve the electrochemical performance of MnO2-based cathodes, the development of MnO2 cathodes doped with metal ions or composites of MnO2 with conducting polymers has shown such advantages as increasing the specific capacity and cycling stability. This mini-review focuses on the strategies to improve the electrochemical performance of manganese-based cathodes of AZIBs.

1. Introduction

Since 1991, lithium-ion batteries have been at the forefront of modern science and technology in the field of energy storage materials due to their high energy density. Nevertheless, their recycling is still a global problem due to the high cost and insufficient efficiency of these processes [1,2]. In fact, the development of clean and safe metal-ion batteries with a low cost, high energy efficiency and cycle stability, and environmental friendliness, the so-called “beyond Li-ion batteries”, has attracted global attention. Among such batteries, aqueous zinc-ion batteries (AZIBs) are state-of-the-art due to their high zinc abundance, high volumetric and gravimetric theoretical capacity values (5854 mAh·cm–3 and 820 mAh·g−1), safety, and non-flammability [3,4,5,6,7]. In addition, metallic Zn has a comparatively low redox potential (−0.76 V vs. Standard Hydrogen Electrode (SHE)), is non-toxic, and is inexpensive.
The key problem for multivalent metal-ion batteries is the selection of a suitable cathode material with high specific capacity and stable cycling performance. Bivalent zinc ions Zn2+ have a strong electrostatic interaction with the host lattice, which leads to moderate Zn2+ diffusion in the crystal lattice of the cathode material and consequently to unsatisfactory electrochemical properties. Among the studied cathode materials, manganese compounds have been the most widely used in the development of rechargeable AZIBs since their appearance in 2011 [8,9,10,11,12,13,14]. In particular, manganese dioxide MnO2 occupies a leading position due to its wide variety of polymorphic phases, which allow the construction of materials with a given morphology, high theoretical specific capacity (308 mAh·g−1 per one electron transfer Mn4+-Mn3+ and 616 mAh·g−1 per two-electron transfer Mn4+-Mn2+) and relatively high redox transition potential (≈1.3–1.5 V vs. Zn/Zn2+). Despite all the advantages of MnO2 as a cathode material, there are two main problems that limit its further large-scale development [14,15]. First, the electrical conductivity of MnO2 is relatively low (≈5 × 10−6 S∙cm−1). The low electronic conductivity value affects the electron transport in the cathode material, reducing the rate capability and cycling stability during cycling. Second, during the one-electron redox process, MnO2 transforms to Mn3+, which is unstable due to the Jahn–Teller effect and undergoes a disproportionation reaction to Mn4+ and Mn2+. The second product has a high solubility in the aqueous solutions, leading to dissolution of the cathode material during the charge/discharge process. Third, MnO2-based materials tend to undergo structural transformations due to the high number of polymorphs and electrostatic repulsion between MnO6 octahedra and Zn2+ ions. All of these drawbacks have a critical impact on the electrochemical performance, especially the cycling stability of MnO2-based cathodes.
To overcome structural instability and suppress manganese dissolution during cycling, the addition of manganese salts to the aqueous electrolyte is widely used [16,17,18,19]. The mixed Zn-Mn aqueous electrolyte is commonly used to evaluate the electrochemical performance of MnO2 cathodes. However, in aqueous electrolytes at high potentials (E > 1.7 V vs. Zn/Zn2+), a competitive reaction of oxygen evolution due to water oxidation may occur. A detailed study of Pourbaix diagrams of Zn-Mn aqueous solutions showed that the optimal composition of the electrolyte solution is 2 M ZnSO4 and 0.1 M MnSO4 (Figure 1) [20].
To date, the studies of rechargeable aqueous Zn//MnO2 batteries have accumulated a huge amount of experimental data on the electrochemical behavior of MnO2 cathodes, their structural and chemical characterization, and the approaches to improve their functional characteristics. Several strategies have been used to improve the properties of MnO2 cathodes, such as the synthesis of nanosized MnO2 with a given morphology, doping or pre-intercalation of metal ions into the manganese dioxide, oxygen defect engineering, and the development of composite materials with carbonaceous materials or conducting polymers, among others.
This mini-review focuses on a thorough analysis of the two hottest and most promising strategies for the modification of MnO2. The first strategy is doping with metal ions, which has been the subject of a large number of recently published works. As reported in most of them, this approach resulted in an increased interlayer spacing, which allowed for improved ion intercalation kinetics. The second equally interesting strategy is the development of MnO2-based composites with conducting polymers, which can also facilitate Zn2+ diffusion and increase the structural stability of the MnO2 cathodes. In addition, the combination of these strategies was analyzed, and the structural and electrochemical properties of the modified materials were thoroughly compared.
The authors of this review deliberately limited themselves to reviewing a selected specific class of MnO2-based cathode materials, since the review is part of a series of mini-reviews in which the role of metal ion pre-intercalation and conducting polymer can be presented as the main topic. The specific focus of this review, as opposed to the multidirectionality and similarity of many other reviews, is precisely on these aspects of cathode modification. We believe that this allows for a more in-depth analysis of the effects of the two aforementioned strategies of improvement of MnO2-based composite cathodes, thereby emphasizing the novelty of the review.
Another important problem considered here is the issue of the complex mechanism of electrochemical transformations in aqueous Zn//MnO2 batteries. Such complexity arises due to the involvement of three types of species (Zn2+, Mn2+, and H+ or water molecules). A high complexity is the presence of reversible Zn2+ movement into the crystal lattice of MnO2 due to high electrostatic repulsion. The most commonly described pathways of the cathodic reaction mechanism were compiled and are carefully analyzed here to find the most probable path of the electrochemical reaction.

2. Metal-Ions-Doped Manganese-Dioxide-Based Cathode Materials

Doping is usually defined as the introduction of a heteroatom into the structure of a host material in order to enhance its properties. It can result in interlayer doping or the substitution of Mn atoms in the lattice, depending on the initial material structure.
Metal ion doping is typically used for materials with good crystallinity, while nonmetal doping is associated with carbon-based materials [21]. In many cases, metal-ion-doped materials have increased interlayer spaces or stronger bonds between structural fragments, superior conductivity, and reduced electrostatic interaction forces, resulting in increased specific capacitance and cycling stability [22,23,24,25,26,27,28,29,30,31]. Various metal ions, such as alkali and alkaline earth metal ions (Li+, Na+, K+, and Ca2+); transition metal ions, including VIII group metals (Zn2+, Co2+, and Cu2+); rare-earth-metal ions (La3+ and Ce3+), and post-transition metals (Al3+, Bi3+, and Sn4+), have been used as dopants for MnO2-based cathodes.

2.1. Alkali and Alkaline-Earth Metal Ions

Light metal ions (Li+, Na+, and K+) were chosen for the preparation of alkali-doped layered δ-MnO2 cathode material [23] to decrease the additional mass of pre-intercalated cations, which can reduce the specific capacity of the material. All three materials (denoted as Li-δ-MnO2, Na-δ-MnO2, and K-δ-MnO2) were prepared by the precipitation method from the manganese chloride solution treated with H2O2 in an alkaline environment formed by the corresponding hydroxide (LiOH, NaOH, or KOH) of the same concentration. Based on the XRD patterns of the synthesized materials, the layered crystal structure of birnessite (δ-MnO2) was confirmed for Li-δ-MnO2, Na-δ-MnO2, and K-δ-MnO2, with a small shift of the main peak to lower 2θ values, which is related to the increase of the interlayer spaces, as shown in Figure 2a. The molar ratios of the metals determined by ICP mass spectrometry were 0.42:1, 0.43:1, and 0.52:1 for K:Mn, Na:Mn, and Li:Mn pairs, respectively. The valence state of manganese in these doped materials determined from XPS spectra was 3.40, 3.42, and 3.46 for K-doped, Na-doped, and Li-doped, respectively. Scanning and transmission electron microscopy data showed that all alkali-doped materials have a two-dimensional nanosheet morphology; the structure of birnessite was further confirmed by HR-TEM and atomic force microscopy; and the interlayer spacings were determined to be 0.64, 0.55, and 0.52 nm for K-, Na-, and Li-doped materials, respectively. The electrochemical performance of these materials was tested in an aqueous solution of 2 M ZnSO4/0.1 M MnSO4. For Li-δ-MnO2, the lowest specific capacity values were obtained at a current density of 0.3 A∙g−1 (192 mAh∙g−1 after 100 cycles). At the same time, K-δ-MnO2 delivered the highest specific capacity (282 mAh∙g−1 after 100 cycles at 0.3 A∙g−1). In addition, the capacity retention for K-δ-MnO2 and Na-δ-MnO2 was greater than 50%, while this value was only 30% for Li-δ-MnO2 after 1000 cycles at 2.0 A∙g−1 (see Figure 2b). The rate performance evaluated in the current range 0.1–3.0 A∙g−1 was the highest for K-δ-MnO2: 270 mAh∙g−1 at 0.1 A∙g−1 and 95 mAh∙g−1 at 3.0 A∙g−1. The reason for the superior performance of the K-doped material is the most significant interlayer expansion, which leads to facilitated ion diffusion through the crystal lattice.
Despite the rather modest electrochemical performance of lithium-doped layered-type MnO2 cathode materials, Li+ ions have been used as a dopant for α-MnO2 [32]. First, MnO2 was obtained hydrothermally from KMnO4 and MnSO4 mixed solution, and then MnO2 was transferred to organic solvent and N-butyllithium was used as lithium source to obtain Li-MnO2. The crystal structure of the tetragonal α-MnO2 was confirmed by XRD; the molar ratio of Li:Mn was close to 0.023:1 based on ICP emission spectroscopy. The valence state of manganese decreased from +4 to +3 after Li+ insertion and more oxygen defects were observed in Li-MnO2 from the XPS data. The morphology of the Li-MnO2 was described as nanorods with slight lattice expansion. The crystal space in Li-MnO2 expanded from 0.305 to 0.463 nm. The specific capacity values for the Li-doped material were even lower than those for the undoped manganese dioxide (185 mAh∙g−1 and 225 mAh∙g−1 at 0.1 A∙g−1, respectively), while the capacity retention for Li-MnO2 was close to 100% after 100 cycles at 0.1 A∙g−1 (Figure 2c). The artificial increase in specific capacity for Li-MnO2 is associated with the deposition of an electroactive MnOx layer on the electrode surface. The charge transfer resistance evaluated from the impedance spectra was higher for Li-MnO2 (Figure 2d inset). Li-MnO2 electrodes exhibited high reversibility of the charge/discharge processes at different current densities in the rate capability tests; the specific capacity recovered to 156 mAh g−1 at 0.1 A∙g−1 after the current density of 2 A∙g−1 (Figure 2e). Over 1000 cycles at 1.0 A∙g−1, the capacity retention for the Li-MnO2 cathode was 89%, while for MnO2, only 50% of the capacity was retained; meanwhile, the coulombic efficiency was about 100% (Figure 2f).
Due to the more stable crystal lattice and enlarged crystal spaces of the Li-MnO2 material, the diffusion of Zn2+ and H+ was facilitated despite the lower initial specific capacity and higher charge transfer resistance of the material due to the insufficient increasing of the interlayer distances and, as a consequence, high electrostatic repulsion of Zn2+ ions [23].
The electrochemical deposition of Na-doped manganese oxide (Na:MnO2) on the graphene-like carbon film (GCF) was performed in [33]. The GCF was synthesized from the pristine graphite film via the electrochemical intercalation process, as shown in Figure 3a. The electrodeposition of Na:MnO2 was performed at E = 1.2 V (vs. Ag/AgCl) for 400 s in a three-electrode cell from 0.1 M Na2SO4 and 0.05 M MnSO4 aqueous solution. The SEM images showed the uniform distribution of the deposited Na:MnO2 nanosheets on the graphene sheets (Figure 3b), and the uniform distribution was also confirmed by an EDX analysis. The layered structure of the doped material was confirmed by the XRD diffraction pattern of the Na:MnO2/GCF, which was indexed to birnessite-type MnO2 (Figure 3c). The peak at 1070.9 eV in the Na 1s XPS spectra (Figure 3d) confirmed the doping of MnO2 by Na. The Na:Mn ratio was 0.15:1 based on ICP/atomic-emission-spectroscopy data. The resulting Na:MnO2/GCF composite was used as a binder-free electrode for further electrochemical tests. The binder-free Na:MnO2/GCF electrode in 2 M ZnSO4/0.1 M MnSO4 aqueous electrolyte achieved the specific capacity of 382 mAh∙g−1 at 0.1 A∙g−1, and at the high current density of 3.0 A∙g−1, the specific capacity was 95 mAh∙g−1. The capacity retention of Na:MnO2/GCF was 83% over 100 cycles at 0.1 A∙g−1, and 75% over 1000 cycles at 1.0 A∙g−1. This outstanding performance was associated with the highly porous 3D structure of the Na:MnO2/GCF composite and stronger electrostatic attractions between the Zn2+ and MnO2 lattice due to the presence of Na, which was confirmed by first-principles DFT calculations.
The doping of MnO2 with potassium ions is the most commonly used strategy, especially for materials with a birnessite structure [28,34,35,36], because KMnO4, as a raw material, acts as a source of K+ and Mn4+ at the same time. Due to this fact, almost all birnessite-type structures have potassium ions in the interlayer spaces. By introducing the K+ ions and oxygen defects into the crystal lattice, the electrochemical performance of MnO2-based materials was improved; e.g., the specific capacity was increased to 270–300 mAh∙g−1 at 0.1 A∙g−1 [34,35].
One of the strategies to prepare potassium-enriched MnO2 (KMO) is the hydrothermal treatment of as-synthesized birnessite in KOH solution [36]. This allowed the crystal lattice of birnessite to be preserved, while the K:Mn ratio was 0.29:1, and the interlayer spacing was increased from 6.8 to 7.4 Å. The specific capacity values of KMO at I = 0.2 A∙g−1 and I = 3.0 A∙g−1 were 300 and 136 mAh∙g−1, respectively. The capacity retention of KMO at I = 0.2 A∙g−1 was 92% during 500 cycles and 98% after 12,000 cycles at I = 2 A∙g−1.
Another crystal structure that can be easily doped by K+ ions is α-MnO2. Hydrothermal or polyol syntheses have been used to deposit K-doped manganese dioxide with a K+ content less than 0.2 on different carbon substrates [37,38,39]. The morphology of the samples mainly depended on the substrate: nanotubes [37], nanoflowers [38], or nanowires [39]. It is interesting to note that α-K0.19MnO2 nanotubes grown on the carbon nanofibers delivered 270 mAh∙g−1 at 1 C (1 C = 0.308 A∙g−1) in 3 M Zn(CF3SO3)2/0.2 M Mn(CF3SO3)2 electrolyte, while at 20 C, the specific capacity was as high as 113 mAh∙g−1, and the capacity retention over 50 cycles at 1 C was 78%. The presence of K+ ions in the electrolyte was shown to improve the cycling stability. The addition of 3 M K(CF3SO3) to the electrolyte allowed the researchers to increase the cycling stability at moderate C-rates: 98.5% of capacity retention was achieved after 50 cycles at 1 C, 92% after 200 cycles at 2 C, and 90% after 400 cycles at 5 C [37].
K-doped MnO2 (K0.15MnO2 and K0.23MnO2) was grown via an in situ hydrothermal reaction on layered graphite paper (LGP) as a reaction template (Figure 4a). The conductive LGP substrate was preliminarily obtained by electrochemical treatment of the graphite paper. LGP@K0.15MnO2 exhibited a nanoflower morphology consisting of interconnected nanosheets that were deposited in the interlayer spaces of LGP (Figure 4b,c). According to BET measurements, the LGP@K0.15MnO2 had a higher specific surface area (20.881 m2/g) and pore volume (0.192 cm3/g) than K0.23MnO2. The resulting binder-free LGP@KxMnO2 cathodes were tested in a 2 M ZnSO4/0.1 M MnSO4 aqueous electrolyte. To investigate the effect of the conductive substrate on the electrochemical performance of the binder-free electrode, KxMnO2 was also deposited on the carbon cloth to obtain CC@KxMnO2. The LGP@K0.15MnO2 was shown to outperform K0.23MnO2 and CC@KxMnO2 in rate capability tests (Figure 4d). At a low current density of 0.05 A∙g−1, the specific capacity value of LGP@K0.15MnO2 was 403 mAh∙g−1, and at I = 2.0 A∙g−1, it was 116.1 mAh∙g−1. The capacity retention over 100 cycles at I = 0.2 A∙g−1 was 92.5% (Figure 4e) [38]. Potassium-doped MnO2 (KMO) nanowires grown on carbon nanotubes mixed with graphene substrate, when used as a binder-free electrode, delivered 360 mAh∙g−1 at 0.05 A∙g−1 due to the presence of highly conductive substrate and high concentration of Mn2+ (0.4 M MnSO4) in the electrolyte [39].
Calcium-doped MnO2 was prepared by hydrothermal method from mixed manganese solution (KMnO4 + MnSO4) in the presence of calcium chloride [40]. The crystal structure of the obtained product (CaMnO) corresponded to layered-type δ-MnO2 with increased interlayer space due to the presence of Ca2+ ions. The Ca:Mn ratio calculated from the XPS data was 0.24:1, and 33% of the Mn was in the +3 valence state. The more accurate Ca:Mn ratio determined by ICP-OES was 0.28:1. According to the SEM-characterization, CaMnO was composed of nanorods and nanospheres that were formed by uniform interconnected nanoflakes. The initial specific capacity of CaMnO electrodes in 1 M ZnSO4/0.1 M MnSO4 aqueous solution at a current density of 0.35 A∙g−1 was 277.7 mAh∙g−1, and at 3.5 A∙g−1, it was 124.5 mAh∙g−1. After 5000 cycles at 3.5 A∙g−1, the retained capacity was 100.9 mAh∙g−1 (81%), which is quite good for AZIBs. The electrolyte pH was shown to significantly affect the electrochemical performance of CaMnO. In acetate solution (1 M ZnAc2/0.1 M MnAc2) at I = 1.5 A∙g−1, the cycling stability was much lower than in sulfate-based solution (only 39.4% of capacity retention was observed over 500 cycles).
Thus, alkali and alkaline earth metal ions are quite suitable dopants for MnO2-based cathode materials due to their ability to increase the interlayer spacing in tunnel or layered crystal structures. As can be seen from the ionic radii, Li+ ions have an insufficient effect on the electrochemical performance of cathode materials due to their negligible effect on the crystal lattice spacing, while the use of sodium or potassium ions endowed the doped MnO2-based cathodes with superior properties. The use of divalent Ca2+ ions with an ionic radius close to that of K+ as a dopant allows a significant improvement in cycling stability due to the stronger bonding of the dopant ion to the oxygen in the MnO6 octahedra. In this case, the increase in interlayer spacing due to doping resulted in superior capacity values, and the interaction of the dopant with the host lattice influenced the stability during long-term cycling.

2.2. Transition and Rare-Earth-Metal Ions

2.2.1. Light Transition Metal Ions

Among the different metal ions, the pre-intercalation of Zn2+ into the MnO2 crystal lattice is an attractive way to improve the electrochemical performance due to Zn2+ intended reversible insertion/extraction into the MnO2 host during the cycling process in Zn-containing electrolytes. In particular, Zn2+ ions were inserted into layered MnO2 to stabilize the host structure by a simple reaction between KMnO4 solution and Zn powder in acidic media [41]. The birnessite-type δ-MnO2 structure of the obtained product was confirmed by XRD, and a typical flower-like morphology of the sample consisting of ultrathin nanosheets was observed. The incorporation of Zn2+ ions into the tunnels between MnO2 layers was confirmed by the XPS data, and the Zn:Mn ratio was 0.16:1. The specific capacities of Zn-doped MnO2 in the conventional mixed 2 M ZnSO4/0.1 M MnSO4 aqueous electrolyte were 275 mAh∙g−1 at the low current density of 0.3 A∙g−1, and 121 mAh∙g−1 after a tenfold increase in current density. This excellent rate performance was attributed to the nanostructure of the material, with nanospheres having a high specific surface area (178 m2∙g−1) and providing more active sites and diffusion channels. The cyclic stability of the material was investigated at current densities of 0.3, 1.0, and 3.0 A∙g−1 during 100, 500, and 2000 cycles, respectively. In the absence of the Mn2+ additive in the solution, a rapid capacity fading (more than 50%) was observed at 0.3 A∙g−1 during 100 cycles. However, in the presence of MnSO4, a gradual increase in capacity was observed under the same conditions, and this is associated with the reoxidation of Mn2+ from the electrolyte solution. Increasing the current allowed the stabilization of the cell, and the capacity retention was greater than 90% in both cases. Thus, the presence of pre-intercalated Zn2+ ions led to the stability of the host material with the formation of a mesoporous flower-like morphology that facilitated reversible Zn2+/H+ insertion/extraction.
To improve the electronic conductivity of the Zn-doped MnO2 cathode, the oxide was deposited from KMnO4 solution with HCl and Zn(NO3)2 in the presence of graphene quantum dots to obtain a GQDs@ZnxMnO2 composite, as shown in Figure 5a [42]. According to the SEM image (Figure 5b), the composite material consisted of monodispersed and uniform nanoflowers, GQDs were formed on the Zn-doped MnO2 surface. The crystal structure of the ternary composite matched to the layered-type δ-MnO2. During the insertion of Zn2+ into the MnO2 lattice, a partial decrease of the valence state from Mn4+ to Mn3+ was observed from the XPS data. The electrochemical performance of GQDs@ZnxMnO2 in 1 M ZnSO4 aqueous solution was remarkable compared to other Mn-based cathodes: 404 mAh∙g−1 at 0.3 A∙g−1, 211 mAh∙g−1 at 4.0 A∙g−1, and the capacity retention of 88% after 500 cycles at 1.0 A∙g–1 was observed (Figure 5c). The polarization value, defined as the difference between the cathodic and anodic peak potentials on the cyclic voltammograms, was lower in the case of the GQDs@ZnxMnO2 composite compared to the pristine MnO2 (Figure 5d). Thus, the modification of MnO2 by GQDs and the Zn intercalation are responsible for providing active sites and conductive media for electron/ion transfer and for such high specific capacity values of the GQDs@ZnxMnO2 composite.
The mixed Mn3O4-MnO2 oxide was modified with Zn2+ ions to induce the electronic and ionic conductivity of the composite material [27]. This composite material was electrodeposited on the Ni foam coated with silver-nanoparticle-modified carbon nanotubes (AgCNT-Ni). The AgCNT coating was used to enhance the interactions between the substrate and the deposited manganese oxides; the structure of the resulting composite is shown in Figure 5e. Electrodeposition was performed in the mixed solution of MnSO4 and ZnSO4 at E = −1.8 V (vs. Ag/AgCl). As a result, Zn-doped Mn3O4-MnO2 vertical nanosheets were formed, as was observed in the SEM images. The analysis of the crystalline phase composition showed that the material consisted of Mn3O4 and γ-MnO2, without any impurities. The valence states of manganese (Mn2+, Mn3+, and Mn4+) and the presence of Zn2+ ions were confirmed by XPS data. The specific capacity of this binder-free composite cathode material, obtained in aqueous solution of 2 M ZnSO4/0.1 M MnSO4 at a low current density (0.3 A∙g−1), was 562.1 mAh∙g−1. This value surpassed almost all results obtained for MnO2-based cathodes, and even at I = 6.0 A∙g−1, the composite retained the specific capacity of 272.7 mAh∙g−1 (Figure 5f). Despite this brilliant rate performance, the cycling stability of the composite at 3.0 A∙g−1 was much lower than reported for other materials: only 69.4% of capacity retention after 200 cycles. However, it was much better than for the Zn2+-undoped material, where the capacity fading after 200 cycles was 39%. These remarkable properties of the Zn2+-doped Mn3O4-MnO2 composite cathode were attributed to the complex 3D porous structure of the binder-free electrode, which facilitates Zn2+ transport due to zinc doping and multiple valence states of manganese, favoring fast redox reactions and a high rate performance.
Vanadium ions can be easily inserted into the MnO2 lattice due to the close ionic radii of V5+ (0.53 Å) and Mn4+ (0.54 Å), and V-doped MnO2 was successfully synthesized via ambient redox reaction [25]. The solution of KMnO4 was mixed with Mn(CH3COO)2 containing a small amount of dissolved V2O5 under continuous stirring. The crystal structure of the V-doped material was defined as tetragonal α-MnO2, with a slight shift of the characteristic XRD peaks to lower angles. Based on ICP-AES measurements, the V:Mn ratio was calculated to be 0.05:1. The valence state of vanadium was defined as V5+ according to the XPS results. The electrochemical tests performed in 1 M ZnSO4 showed that the initial specific capacity value for V-doped MnO2 was 266 mAh∙g−1 at I = 0.066 A∙g−1, while for undoped MnO2, 213 mAh∙g−1 was achieved at the same current density. Increasing the current to 1.06 A∙g−1 resulted in a significant decrease in the capacity of V-doped MnO2 to 67 mAh∙g−1. The capacity retention at the low current of 0.066 A∙g−1 after 100 cycles was close to 50%. The improved specific capacity and rate performance of V-doped MnO2 were explained by an increase in the conductivity of MnO2 after doping, and this was further evidenced by a decrease of the charge-transfer resistance in the impedance spectra.
V-doped MnO2 aerogel was obtained via the coprecipitation method from dissolved V2O5 gel solution with MnSO4, which was then oxidized with (NH4)2S2O8, under continuous stirring [43]. The resulting aerogel (A-MnO2) had a crystalline structure of birnessite and was composed of ultrathin nanosheets with many wrinkles. According to the XPS survey spectrum, A-MnO2 consisted of Mn, V, and O elements, and the molar ratio of V:Mn was 0.07:1. The valence state of vanadium was V5+, and manganese had two valence states, Mn3+ and Mn4+, that are typical of doped materials. The electrochemical performance was evaluated in 2 M ZnSO4 aqueous solution without manganese additive. The initial specific capacity of A-MnO2 at 0.2 A∙g−1 was 194 mAh∙g−1, and at I = 2.0 A∙g−1, the capacity was 74 mAh∙g−1. The cyclic performance of A-MnO2 evaluated at 0.3 A∙g−1 during 100 cycles showed the capacity retention of 71%, with a small capacity growth attributed to the activation of the cathode, including a better infiltration of the electrolyte into the porous material. The addition of 0.1 M MnSO4 to the electrolyte solution increased the battery lifetime by a factor of six, indicating that the Mn2+ additive has a crucial effect on the manganese dissolution during the cycling process. The improvement of the V-doped material was mainly explained by the oxygen defects, high conductivity, and porosity of A-MnO2.
Group VIII metals (Fe, Co, and Ni) can be used as dopants for MnO2 due to their ability to suppress the Jahn–Teller effect, which leads to a decrease in manganese disproportionation and consequent dissolution [44]. Fe-doped manganese oxide (FMO) was obtained hydrothermally from potassium permanganate solution in the presence of iron nitrate. According to X-ray diffraction data, the product had a birnessite-type δ-MnO2 crystalline structure, and the characteristic peaks of FMO were broadened and lower in intensity compared to those of undoped MnO2 due to an increase in lattice disorder caused by Fe doping (Figure 6a). However, the nanoflower-like morphology, which is typical of birnessite-type structures, was still maintained after Fe-doping (Figure 6b). The EDX analysis showed a uniform distribution of Mn, O, and Fe elements in the FMO sample. The HRTEM images (Figure 6c) show the distinct lattice fringes with a spacing of 0.714 nm, which can be indexed to the (001) plane of MnO2. The survey XPS spectra showed the presence of Mn and Fe peaks, confirming the intercalation of Fe into the MnO2 structure. The analysis of the valence states of manganese and iron had shown that only a small part of Mn4+ was reduced to Mn3+ after the insertion of Fe3+. The electrochemical properties of FMO as a cathode material were investigated in the aqueous solution of 2 M ZnSO4/0.1 M MnSO4. The initial specific capacity of FMO at a current density of 0.1 A∙g−1 was 170 mAh∙g−1, and it increased twice by the third cycle due to the activation process. The maximum specific capacity delivered by FMO at I = 0.1 A∙g−1 was 390 mAh∙g−1, and after increasing the current to 3.0 A∙g−1, this value was 160 mAh∙g−1. This was explained by the inhibition of Jahn–Teller distortion due to the increase in the mean valence of manganese after Fe intercalation and the suppression of manganese dissolution. The influence of this effect on the cycling stability was demonstrated at the current density of 1.0 A∙g−1 during 200 cycles, when the capacity retention of 86.3% was observed. Thus, pre-intercalation of MnO2 with Fe ions stabilized its crystal structure and resulted in superior electrochemical performance.
Cobalt ions are the most widely used dopant for MnO2 cathodes among other Group VIII metals [26,29,45,46]. For chemical modification, the as-synthesized MnO2 powder was treated with aqueous CoCl2 solution under continuous stirring [45]. The XRD pattern of the obtained Co-doped MnO2 product was consistent with the δ-MnO2 crystal structure, and the product had a flower-like morphology and uniform distribution of the Mn, O, and Co elements according to the SEM and EDX results. The rate performance tests in mixed zinc–manganese aqueous electrolyte showed that Co-doped MnO2 had the lowest initial capacity values in the potential range of 1.0–1.8 V compared to unmodified δ-MnO2 (196 and 252 mAh∙g−1 at 0.1 A∙g−1, respectively), but the most intensive capacity growth was observed during continuous cycling at 0.3 A∙g−1. It was shown that cobalt ions catalyzed the MnO2 electrodeposition during cycling, resulting in a stabilization of the specific capacity over 5000 cycles at I = 2.0 A∙g−1 with a capacity retention of 80%. This catalytic mechanism was proposed based on the comproportionation between Mn2+ and Mn4+ and the reversible oxidation of Co2+ to Co3+ at E < 1.85 V (Figure 6d).
Another way to obtain Co-doped MnO2 is a hydrothermal synthesis from KMnO4 solution in the presence of Co(NO3)2 on a carbon cloth in an acidic environment; the product was further treated by plasma to obtain vacancy-rich Co-MnO2 [29]. The crystal structure of Co-MnO2 was defined as α-MnO2 before and after plasma treatment. The Co-MnO2 had a nanowire-like morphology and was uniformly distributed on the carbon cloth substrate. The valence states of manganese evaluated from the XPS data were Mn4+ and Mn3+, and for Co, the mixed-valence state, consisting of Co2+ and Co3+, was detected. The concentration of oxygen vacancies after the plasma treatment, calculated from O 1s spectra, was 19–29%, depending on the treatment time (1–3 min), with 2 min being the optimal time (29%). The electrochemical performance of vacancy-rich binder-free Co-MnO2 cathodes was tested in 2 M ZnSO4/0.1 M MnSO4 aqueous solution. The initial specific capacity at a current density of 0.5 A∙g−1 was 511 mAh∙g−1. At a high current density of 5.0 A∙g−1, the delivered specific capacity of plasma-treated Co-MnO2-2 was 100 mAh∙g−1, which is about 25% higher than that of Co-doped MnO2 without additional oxygen defects (Co-MnO2-0) (Figure 6e). The capacity retention for plasma-treated Co-MnO2-2 over 1000 cycles at I = 3.0 A∙g−1 was 98% (Figure 6f). Therefore, Co-doping, together with the formation of oxygen vacancy-rich structures, had the synergistic effect of inhibiting the Jahn–Teller effect and increasing the conductivity and performance of the material.
Electrodeposition of Co-doped MnO2 has also been used to develop binder-free cathodes for Zn//MnO2 batteries [26,46]. To overcome the poor conductivity of MnO2, Co-doped oxide was grown on the carbon cloth from a solution of MnSO4, ZnSO4, and CoSO4 at E = 1.0 V (vs. Ag/AgCl) [26]. Under these conditions, Zn2+-intercalated Co-doped δ-MnO2 with a birnessite-type structure was obtained. The XPS analysis confirmed that Co and Zn atoms were incorporated into the material structure, the valence state of Co was Co3+, and these Co3+ ions were inserted into the layers of MnO6 octahedra, not between them. The intercalation of Co3+ significantly improved the rate capability of the Co-MnO2 film: 280 mAh∙g−1 at 1.2 A∙g−1 and 205 mAh∙g−1 at I = 3.0 A∙g−1 were achieved. During 200 cycles at 1.2 A∙g−1, a rapid capacity increase from 60 to 290 mAh∙g−1 was observed, with a consequent capacity fading of 8% only in the last few cycles. The charge-transfer resistance of the Co-MnO2 material, evaluated from the impedance spectra, was five times lower than that of the undoped material. The reason for these remarkable electrochemical properties of Co-doped MnO2 is the increased electrical conductivity of the material after Co3+ intercalation.
The electrochemical synthesis on the stainless-steel mesh from solution of MnSO4 and Na2SO4 in the presence of Co(CH3COO)2 at E = −1.8 V (vs. SCE) allowed researchers to obtain layered-type Co-MnO2 [46]. The valence states of Mn and Co were investigated by XPS, and, after doping, the manganese had a mixed valence of Mn3+ and Mn4+, while for cobalt, only one valence state, Co3+, was observed. The electrochemical tests were carried out in a mixed zinc–manganese electrolyte, with and without the addition of 0.02 M Co(CH3COO)2, to check the possibility of additional electrochemical intercalation of Co. The presence of Co2+ ions in the electrolyte solution had a positive effect on the specific capacity in the current range of 1.0–3.0 A∙g−1: 310 mAh∙g−1 and 175 mAh∙g−1 at I = 1.0 and 3.0 A∙g−1, respectively, while the same cathode in the solution without cobalt additive showed 260 and 145 mAh∙g−1 at the same current densities. The capacity retention in the ternary electrolyte solution over 1000 cycles at I = 1.0 A∙g−1 was 91.78%. Thus, a synergistic combination of a Co-doped MnO2 cathode and a Co2+ electrolyte additive allowed the rate and cycle performance of the electrodeposited material to be significantly increased.
Nickel has also been used to develop doped MnO2-based cathodes [44,47]. Ni-doped MnO2 was prepared hydrothermally from a mixed solution of KMnO4 and Ni(NO3)2 in an acidic medium [47]. The XRD analysis showed that the product (NKMO) had a tetragonal crystal lattice corresponding to α-MnO2, in which 5% of the Mn was substituted by Ni. The Ni:Mn elemental ratio was defined as 0.055:1 from the ICP-OES results. The specific discharge capacity of the NKMO cathode at 0.05 C (where 1 C = 0.308 A∙g−1) was 303 mAh∙g−1 in the mild mixed aqueous 2 M ZnSO4/0.2 M MnSO4 electrolyte, which was 29% higher than that of the undoped MnO2-based cathode. Even at the high current density of 4 C, the specific capacity was 190 mAh∙g−1, but the capacity retention over 2000 cycles was slightly lower (71%) than for MnO2 (73%). Theoretical investigations by DFT calculations showed that the structural transformation from the tetragonal to the orthorhombic phase is more effective in the presence of Ni, leading to the enhancement of H+ diffusion into the MnO2 tunnels and, consequently, to a superior rate performance.
Copper, another rather popular dopant for various cathode materials, has also been used for MnO2 modification [48,49]. To obtain Cu-doped MnO2 (CuMO) materials, a hydrothermal synthesis from KMnO4 and CuSO4 solutions in the presence of citric acid was used [49]. The crystal structure of the as-synthesized product was matched to the birnessite-type δ-MnO2, and the morphology of this material was described as nanoflowers consisting of nanoflakes. The presence of Cu 2p peaks in the XPS spectra confirmed the intercalation of Cu2+ ions into the MnO2 lattice, and the valence state of the manganese was slightly reduced after the doping. The Cu:Mn ratio in the CuMO determined by the ICP-OES was 0.06:1. The specific capacity values of the CuMO obtained in 2 M ZnSO4/0.1 M MnSO4 aqueous electrolyte at I = 0.1 and 2.0 A∙g−1 were 493.3 and 125.8 mAh∙g−1, respectively, which was much higher than those of unmodified MnO2. This remarkable rate performance was related to the high conductivity of the Cu-doped material. The capacity retention for 150 cycles at I = 0.5 A∙g−1 was close to 95%, with capacity growth observed during continuous cycling due to MnOx electrodeposition from the electrolyte solution. Thus, CuMO undergoes a displacement mechanism in which Cu2+ is reduced to metallic form and extracted from the crystal lattice.
Cu2+ pre-intercalated δ-MnO2 (CuMO) was also hydrothermally synthesized on the carbon cloth from the KMnO4 solution mixed with CuCl2 to fabricate a flexible binder-free electrode [48]. The morphology and microstructure of the as-prepared sample were described as nanowires which are homogeneously anchored on carbon fibers (Figure 7a). Based on the XRD patterns, the crystalline structure of CuMO was shown to be δ-MnO2. XPS peaks of Cu 2p in the 935 eV binding energy region confirmed that Cu ions were intercalated into the MnO2 host structure. This led to an increase in the Mn3+ content in the CuMO material by creating active oxygen defects. The molar ratio of Cu:Mn was determined to be 0.1:1 by ICP-OES. The higher conductivity of the CuMO material was confirmed by DFT calculations of the bandgap, which was lower for the Cu-doped sample. The rate performance of the CuMO electrode was investigated in an aqueous solution of 2 M ZnSO4/0.2 M MnSO4 in the current range of 0.1–5.0 A∙g−1. At a low current density of 0.1 A∙g−1, the specific capacity value was 398 mAh∙g−1, and at I = 5.0 A∙g−1, it was 125 mAh∙g−1. As observed for another Cu-modified electrode [49], this remarkable rate performance was due to the high conductivity of the material. The cycling performance of the CuMO electrode was evaluated at current densities of 0.2 and 5.0 A∙g−1 over 90 and 700 cycles, respectively. At a low current density of 0.2 A∙g−1, a slight increase in specific capacity up to 280 mAh∙g−1 was observed during 50 cycles, and then a decrease in capacity to 240 mAh∙g−1 occurred. As shown in the Figure 7b, at the high current density of 5.0 A∙g−1, the capacity retention was 90.1% without significant capacity growth. Thus, rapid and reversible Zn2+ storage was achieved in a structurally modulated δ-MnO2 cathode with high conductivity of CuMO and stability of CuMO capacitive response during long-term cycling.

2.2.2. Heavy Transition Metal Ions

Silver ions were used as dopants for MnO2-based cathodes to develop flexible quasi-solid-state AZIBs [50]. Ag-doped MnO2 was prepared hydrothermally from MnSO4 solution in the presence of (NH4)2S2O8 and AgNO3 at different concentrations. The crystal structure of the as-synthesized product was determined to be α-MnO2 from the XRD patterns. After doping, increased lattice distortion was observed in the MnO2 cell. The valence state of manganese was slightly reduced after the incorporation of Ag+ ions, as was confirmed by XPS spectroscopy. The content of oxygen defects calculated from the O 1s XPS spectra was 23%, slightly higher than for pure MnO2 (19%). Electrochemical tests in an electrolyte solution of 2 M ZnSO4/0.1 M MnSO4 showed that the Ag-doped material had a specific capacity value of 315 mAh∙g−1 at I = 0.05 A∙g−1, and at a high current density of 2.0 A∙g−1, the retained capacity was only 85 mAh∙g−1. The capacity retention during 500 cycles at I = 0.5 A∙g−1 was 94%. During continuous cycling, a slight capacity increase was observed from the 50th to the 300th cycle, and then a gradual capacity decrease took place. According to the impedance spectroscopy data, Ag+ ions intercalated in the MnO2 host decreased the charge-transfer resistance and facilitated Zn2+ migration due to the expanded crystal lattice.
Molybdenum ions are high-valence particles that can form chemical bonds with oxygen atoms and stabilize the Mn3+ to improve the structural stability of MnO2 [51]. To prepare Mo-doped MnO2, KMnO4 solution was mixed with Na2MoO4 solution, and then the resulting mixture was acidified and hydrothermally treated. As a result, a tunnel-type α-MnO2 lattice structure was obtained, and the morphology of the Mo-MnO2 was described as nanorods with a diameter of 44 nm. The XPS data confirmed that manganese had two valence states, Mn3+ and Mn4+, while for molybdenum, Mo5+ and Mo6+ valence states were observed. Electrochemical tests in a conventional zinc–manganese electrolyte in the potential range of 0.5–1.8 V showed that the initial capacity at I = 0.1 A∙g−1 for the Mo-MnO2 cathode was 200 mAh∙g−1, while for MnO2, it was 250 mAh∙g−1, but at a high current density (5 A∙g−1) and during 100 cycles, the specific capacity of Mo-MnO2 was higher than that of the unmodified MnO2 cathode. The better rate capability of Mo-MnO2 was clearly observed at current densities higher than 0.5 A∙g−1 (Figure 7c). The capacity retention of Mo-MnO2 at I = 2.0 A∙g−1 was 82.6%, while for MnO2, only 22% of the capacity was retained. It was also observed that the valence state of Mo changes during cycling, and the partial substitution of Mn3+ by Mo5+ decreased the rate of the disproportionation process, resulting in the improved cycling stability of the Mo-doped material.
The hydrothermal method was also used to obtain Mo-doped MnO2 with a birnessite-type structure by the reaction between KMnO4 and MnSO4 in the presence of (NH4)6Mo7O24 [52]. The morphology of the as-obtained product was typical of layered manganese oxide and was described as irregular microparticles consisting of interconnected nanosheets. The crystal structure of δ-MnO2 was confirmed by XRD, and the expansion of the interlayer spacing from 0.631 to 0.674 nm due to Mo intercalation was observed by XRD and TEM analyses. The valence states of Mo4+ and Mn3+/Mn4+ and the higher concentration of oxygen defects were determined by XPS. The Mo-doped MnO2 cathode in a conventional 2 M ZnSO4/0.1 M MnSO4 electrolyte showed capacity growth at a current density of 0.1 A∙g−1 up to 327 mAh∙g−1, while at the high current density of 3.0 A∙g−1, the specific capacities of MnO2 and Mo-MnO2 were almost equal (107 mAh∙g−1). Moreover, the stability of the Mo-doped material was the same as that of the undoped material (Figure 7d), while the specific capacity of the Mo-MnO2 cathode after 1000 cycles at 1.0 A∙g−1 was 57% higher. Mo ions serve as “pillars” for layered structures which stabilize the crystal lattice by forming Mo–O bonds. In addition, oxygen defects provide more sites for Zn2+/H+ intercalation, which explains the superior performance of Mo-MnO2 in the current range of 0.2–2.0 A∙g−1.
Thus, the transition metal ions could facilitate the diffusion of Zn2+ ions due to their large size or interactions between the metal ion and MnO2 host; however, not all metal ions, such as Ni, Mo, and some others, could be effective dopants for MnO2, while Zn, Fe, Co, and Cu ions are rather good for improving the electrochemical properties of cathode materials, especially for binder-free cathodes.

2.2.3. Rare-Earth-Metal Ions

Rare earth metals have also been successfully intercalated into the MnO2-based cathodes. Lanthanum-doped MnO2 was obtained by a simple precipitation method from mixed solutions of MnSO4 and KMnO4 with the addition of La(NO3)3 [53]. XRD and SEM analyses of La-doped MnO2 (LMO) showed that the oxide had a flower-like morphology that is typical of layered-type materials, and the XRD patterns confirmed that birnessite-type δ-MnO2 was obtained. The interlayer distance in LMO observed from the TEM and XRD data was 0.76 nm, while for undoped MnO2, it was 0.68 nm, indicating the expansion of the interlayer space due to La intercalation. The Mn:O element ratio determined from the XPS data was 0.97:2, so the main valence state of manganese was Mn4+. The LMO cathode in 1 M ZnSO4/0.4 M MnSO4 electrolyte reached a high specific capacity of 275 mAh∙g−1 at I = 0.1 A∙g−1, and after increasing the current to 1.6 A∙g−1, the specific capacity was 121.8 mAh∙g−1, which is much better than the almost-zero capacity of the unmodified MnO2 cathode. The probable explanation for this remarkable rate performance is that the pre-intercalated pillars of La3+ promoted Zn2+ diffusion due to the expanded interplanar spaces. The capacity retention of the LMO cathode after 200 cycles at 0.2 A∙g−1 was 71%, which is associated with the high reversibility of the cathodic reaction in the presence of La3+ ions.
Figure 7. (a) SEM image of CuMO [48]. (b) Cycling performance of MnO2 and CuMO electrodes at 5 A∙g−1 [48]. (c) Rate performance of MnO2 and Mo-MnO2 cathodes [51]. (d) Cycling performance of MnO2 and Mo-MnO2 cathodes [52]. (e) Rate capability of Ce-doped MnO2 [54]. (f) GITT curve and diffusivity vs. state of discharge for Ce-doped MnO2 [54].
Figure 7. (a) SEM image of CuMO [48]. (b) Cycling performance of MnO2 and CuMO electrodes at 5 A∙g−1 [48]. (c) Rate performance of MnO2 and Mo-MnO2 cathodes [51]. (d) Cycling performance of MnO2 and Mo-MnO2 cathodes [52]. (e) Rate capability of Ce-doped MnO2 [54]. (f) GITT curve and diffusivity vs. state of discharge for Ce-doped MnO2 [54].
Energies 16 03221 g007
As lanthanum, cerium ions have been used to modify MnO2-based cathode materials [54,55]. Chemically synthesized Ce-doped MnO2 was obtained hydrothermally from MnSO4 solution in the presence of cerium (III) nitrate additive, and (NH4)2S2O8 was used as the oxidant [54]. The diffraction peaks on the XRD patterns of the Ce-doped product corresponded to tunnel-type α-MnO2, while MnO2 obtained in the same conditions without Ce additive corresponded to β-MnO2. Thus, cerium ions enlarged 1 × 1 tunnels of β-MnO2 to 2 × 2 tunnels, corresponding to α-MnO2. The valence states of manganese and cerium evaluated by XPS were +3.75 and +3, respectively. The morphology of Ce-doped MnO2 was uniform nanorods, the length of which decreased with increasing amount of cerium. The specific capacity of Ce-doped MnO2 in aqueous 2 M ZnSO4/0.1 M MnSO4 electrolyte was higher than that of undoped β-MnO2 in the current range of 1–5 C (1 C = 0.308 A∙g−1) (Figure 7e). At I = 1 C, the specific capacity of Ce-doped MnO2 was 160 mAh∙g−1, and at 5 C, the capacity was 134 mAh∙g−1. The cyclic stability over 100 cycles at 2 C was higher for Ce-doped MnO2 than for β-MnO2: the capacity retention was 80% and 50%, respectively. The galvanostatic intermittent titration technique (GITT) test showed that the chemical diffusion coefficient of Zn2+ in Ce-doped MnO2 is significantly increased; it was determined to be 10−10–10−13 cm2 s−1 during the entire Zn intercalation process (Figure 7f), which is 10–100 times higher than that of the pristine β-MnO2 cathode. The result clearly showed that the 2 × 2 tunnel structure of Ce-doped MnO2 facilitated fast and reversible Zn2+ migration, resulting in better rate capability.
Ce-doped MnO2 was potentiostatically deposited at E = 1.2–1.7 V (vs. Ag/AgCl) on the carbon cloth from the MnSO4 solution in the presence of surfactants and cerium nitrate to obtain a binder-free cathode [55]. The resulting Ce-MnO2@CC composite showed a uniform coating with a porous lamellar structure on the substrate surface. The XRD patterns showed that the resulting Ce-MnO2 material was amorphous, which was confirmed by field-emission SEM. XPS spectroscopy was used to confirm the Ce doping of the MnO2 structure and to determine the Mn4+ valence state. The specific capacity of the Ce-MnO2@CC electrode in the mixed zinc–manganese sulfate electrolyte was 292 mAh∙g−1 at 0.1 A∙g−1, while only 179 mAh∙g−1 was achieved for the undoped material. Even at a high current density of 2 A∙g−1, the Ce-doped cathode material still maintained a considerable capacity of 106 mAh∙g−1. During 450 cycles at I = 0.1 A∙g−1, a sharp capacity decay was observed during the first 40 cycles, and then almost-stable capacities were observed with an insignificant increase from 190 to 200 mAh g−1. It was found that cerium ions, together with a highly conductive carbon substrate, accelerate the ionic transport through the cathode material, resulting in a superior rate performance.
Thus, rare-earth-metal ions can be used as dopants for MnO2-based cathodes. However, large ionic radii of La3+, Ce3+, and their analogues resulted in insufficient cycling stability (more than 20% of capacity decay after 100–200 cycles), which is lower than the values for MnO2-based cathodes doped with transition metal cations with smaller ionic radii.

2.3. Post-Transition Metal Ions

Aluminum is one of the most attractive non-alkali and non-transition metals because Al3+ ions can combine with the O atom to form a stable Al-O chemical bond. Al-doped MnO2 has been successfully used as a cathode material for AZIBs [24,56,57]. Al-doped MnO2 with different crystal phases and morphologies was obtained hydrothermally from the solutions of KMnO4 in the presence of Al(NO3)3 [24,57] or Al2(SO4)3 [56].
Al-doped MnO2 (AMO) with an urchin-like morphology consisting of numerous nanorods was obtained by self-reduction of KMnO4 in the presence of aluminum nitrate [24]. The XRD analysis showed that tetragonal α-MnO2 was synthesized, and the observed shift of the most intensive peak to lower 2θ angles (Figure 8a) suggested the change of the space parameter of the crystal lattice due to heteroatom intercalation. Based on a high-resolution TEM analysis, the interplanar distance was close to 0.29 nm, compared to 0.24 nm for undoped material. A decrease in the valence state of manganese to Mn3+ was also observed, resulting in a change of the length of the Mn–O bond. The initial specific capacity of the AMO cathode material in 2 M ZnSO4 without the Mn2+ additive was 401 mAh∙g−1 at a current density of 0.1 A∙g−1, and at high I = 4.0 A∙g−1, the capacity was 229 mAh∙g−1, much higher than that of pristine MnO2 (Figure 8b). This remarkable capacity performance at high currents may be due to the decrease of polarization in the AMO which was observed from the cyclic voltammetry data (Figure 8c). During continuous cycling (over 2000 cycles at a current density of 2.0 A∙g−1) in the absence of MnSO4 additive in the electrolyte solution, the capacity retention was 94.5%. The concentrations of dissolved manganese at different stages of cell life, as determined by the ICP analysis of AMO electrodes, were extremely low (<0.3 mg∙L−1), and the Al3+ concentration was also low (≈0.6 mg∙L−1). Thus, highly stable Al–O chemical bonds not only increase the interplanar distances but also stabilize the interactions within the crystal lattice, preventing MnO2 dissolution.
To stabilize the metal ions in the MnO2 structure, coatings that interact with the intercalated metal ions can be used. In the case of Al3+, lignin was used as such a coating to obtain lignin-coated Al-doped MnO2 (L+Al@MnO2) [56]. During the synthesis, one-dimensional nanorods with the crystal structure of α-MnO2 were obtained. According to the EDX mapping of the main elements (Mn, O, Al, and C), the lignin coating was uniformly distributed on the Al@MnO2 surface. The Al:Mn atomic ratio was close to 1:22 according to ICP-OES. A partial reduction of Mn4+ during the doping process was observed from the XPS spectra of manganese, suggesting that some of the Mn+4 ions were exchanged with Al3+ ions without significant changes in valence state with the introduction of the oxygen defects. Electrochemical tests in 2 M ZnSO4/0.2 M MnSO4 showed that the doping of MnO2 with Al3+ ions improved the diffusion of H+, increasing the specific capacities up to 188 mAh∙g−1, and the lignin coating stabilized the material structure during cycling and prevented the dissolution of manganese, leading to an increase in the cycling stability (66.7% of capacity retention after 3000 cycles at a current density of 1.5 A∙g−1).
Al3+ ions have also been used as dopants for birnessite-type MnO2 with pre-intercalated K+ ions in the interlayer spaces (KMO) [57]. Hydrothermally obtained Al-doped MnO2 with K+ ions (Al-KMO) with a layered structure had a smaller grain size (6.7 nm instead of 8.4 nm), and the pre-intercalation of Al3+ improved the nucleation rate of the grains, resulting in a large specific surface area (44.1 m2∙g−1). Al-KMO had a microflower morphology composed of nanosheets. Compared with the KMO material, the particles of the Al-doped material were more spherical due to the prevention of the agglomeration by stabilizing the internal structure. The XPS analysis showed that Al3+ and K+ were both intercalated in the MnO2 structure, and Al3+ ions exchanged both K+ and Mn4+ ions. This reduced the electrostatic repulsions between intercalated Zn2+ ions and the MnO2 host lattice. The specific capacity of Al-KMO in a zinc–manganese mixed aqueous electrolyte at a low current density of 0.1 A∙g−1 was 327 mAh∙g−1, and at I = 2.0 A∙g−1, it was 85 mAh∙g−1, higher than the capacities of as-prepared KMO and the samples calcined at 300 °C (KMO-300 and Al-KMO-300) (Figure 8d). After 300 cycles at I = 0.5 A∙g−1, the retained capacity of Al-KMO was 205 mAh∙g−1, corresponding to 90% of the capacity retention. However, the Al-KMO had the longest activation process (Figure 8e). In conclusion, the intercalation of Al3+ and co-intercalation of water molecules stabilized the interlayer spaces, thus enhancing the diffusion of Zn2+ and H+ ions and reducing the dissolution of manganese. It was also shown that calcined samples without intercalated water molecules had a lower electrochemical performance than Al-KMO before calcination.
The one-step pre-intercalation of K+, Co2+, and Al3+ ions was used to enhance the diffusion Zn2+ ions and the conductivity of MnO2-based electrode materials [58]. This tri-ion-doped electrode material (α-MnO2@KCoAl) was synthesized by a one-step precipitation method, without heating. The crystal structure of the as-synthesized product determined by XRD corresponded to tetragonal α-MnO2. The morphology of the doped sample was described as irregular lumpy small particles with high dispersity and a specific surface area of ≈6 m2∙g−1. After doping, the valence state of Mn+4 was partially decreased, confirming the intercalation of all heterovalent atoms. The specific capacity of α-MnO2@KCoAl in the 2 M ZnSO4/0.05 M MnSO4 electrolyte at a current density of 0.5 A∙g−1 was 524 mAh∙g−1, which is an extremely high value. Even at the high current density of 5.0 A∙g−1, the specific capacity delivered was 221 mAh∙g−1. This outstanding rate performance of α-MnO2@KCoAl can be explained by the significant decrease of the internal resistance in the ternary-doped composite. Nevertheless, the capacity retention of α-MnO2@KCoAl at 0.5 A∙g−1 was 66% over 100 cycles, and at I = 5.0 A∙g−1, it was only 34% over 800 cycles, which is comparable to the unmodified MnO2 cathode. Despite the high conductivity achieved, the problem of manganese dissolution was not solved by tri-ion doping.
Another trivalent ion with a high ionic radius that has been successfully used for MnO2 modification is the bismuth ion Bi3+ [49,59]. A hydrothermal synthesis of Bi-doped MnO2 (BiMO) was carried out from KMnO4 and Bi(NO3)3 solutions in the presence of citric acid [49]. The layered MnO2 structure was confirmed by XRD, and the patterns of the BiMO were of lower intensity compared to the pristine MnO2. The morphology of the BiMO particles was nanoflowers consisting of interconnected nanoflakes, with uniform distribution of Mn, O, and Bi, as was shown by the EDX analysis. The pre-intercalation of Bi3+ was confirmed by the appearance of Bi 4f peaks in the XPS spectra. The Bi:Mn ratio determined by ICP-OES was 0.09:1. Despite the presence of a dopant, the specific capacity of the BiMO cathode in the 2 M ZnSO4/0.1 M MnSO4 electrolyte at low current densities (0.1–0.4 A∙g−1) was lower than that of unmodified MnO2 (175.5 and 273.4 mAh∙g−1 at I = 0.1 A∙g−1, respectively). Nevertheless, the cyclic stability of the BiMO material at I = 0.5 and 1.0 A∙g−1 was high. The capacity retention over 500 cycles at 0.5 A∙g−1 was 72.3%, and at I = 1.0 A∙g−1, it was 98.6% over 1100 cycles. In addition, at the high current density of 2.0 A∙g−1, the specific capacity of BiMO was 66 mAh∙g−1. Thus, doping with bismuth ions affected the structural stability of the MnO2 crystal lattice and prevented the dissolution of manganese, but the conductivity of this composite was relatively low.
Figure 8. (a) XRD patterns of MnO2 and AMO [24]; (b) GCD profiles of AMO cathode under different current densities [24]. (c) Cyclic voltammograms of MnO2 and AMO cathodes at scan rate 0.2 mV∙s−1 [24]. (d) Rate performance of KMO, Al-KMO, KMO-300 and Al-KMO-300 [57]. (e) Cycling performance of KMO and Al-KMO materials [57]. (f) Schematic relaxed structure of Bi-doped MnO2 [59]. (g) C-rate capability and (h) long-term cycle performance of pristine and Bi-doped MnO2 cathodes [59].
Figure 8. (a) XRD patterns of MnO2 and AMO [24]; (b) GCD profiles of AMO cathode under different current densities [24]. (c) Cyclic voltammograms of MnO2 and AMO cathodes at scan rate 0.2 mV∙s−1 [24]. (d) Rate performance of KMO, Al-KMO, KMO-300 and Al-KMO-300 [57]. (e) Cycling performance of KMO and Al-KMO materials [57]. (f) Schematic relaxed structure of Bi-doped MnO2 [59]. (g) C-rate capability and (h) long-term cycle performance of pristine and Bi-doped MnO2 cathodes [59].
Energies 16 03221 g008
Bi-doped MnO2 (BMO) was also prepared via precipitation from KMnO4 and Bi(NO3)3 solutions with further annealing [59]. After synthesis, nanoparticles with an average size of 50 nm and uniform distribution of Mn, Bi, and O elements were obtained. The crystal lattice of BMO corresponded to α-MnO2. The valence state of manganese defined from the XPS spectra was Mn4+ with small amounts of Mn3+, so the intercalation of Bi affected the electronic structure of MnO2. First-principles calculations were also used to determine the changes in the crystal lattice after doping. It was shown (Figure 8f) that the presence of Bi3+ ions increased the Mn–O bond, causing the expansion of the inner spaces. The electrochemical tests in 2 M ZnSO4/0.2 M MnSO4 aqueous electrolyte showed that BMO had a higher rate capability than undoped MnO2 (Figure 8g), and the specific capacity of BMO at low current density (0.1 A∙g−1) increased continuously due to MnOx deposition and reached 365 mAh∙g−1, while the specific capacity of pristine MnO2 gradually decreased to 200 mAh∙g−1 (Figure 8h). After increasing the current to 3.0 A∙g−1, the delivered capacity was 103 mAh∙g−1, which is a relatively high value. The capacity retention of the BMO electrodes over 10,000 cycles at I = 1.0 A∙g−1 was 93%, indicating that the bismuth ions, which were incorporated into the MnO2 crystal lattice, had a greater effect on the structural stability than on the conductivity of the material.
As shown by the example of Sn doping, MnO2 can also be doped with the tetravalent ions [30]. Manganese dioxide was synthesized by the hydrothermal method from a mixed solution of KMnO4 and MnSO4 in the presence of an acidic SnCl2 solution, and then the product was calcined. The crystal lattice of the obtained product (α-SM) was defined as the tunnel-type α-MnO2 without the presence of tin oxide. The material had a rod-like morphology with a tendency to agglomerate. Tin doping was additionally confirmed by the presence of Sn 3d3/2 and 3d5/2 peaks in the XPS spectra. The valence states of manganese and tin determined by XPS were +4, +3, and +2 for Mn and +4 for Sn. Thus, during the synthesis, Sn2+ was oxidized by potassium permanganate with intercalation of the resulting Sn+4 into the lattice of MnO2, with the formation of oxygen defects. The cyclic performance of α-SM was studied in the current range of 0.1–2.0 A∙g−1. The specific capacities obtained were 240 mAh∙g−1 (0.1 A∙g−1) and 45 mAh∙g−1 (2.0 A∙g−1), with the initial capacity growth during 20 cycles at 0.1 A∙g−1 and subsequent sharp capacity fading. During continuous cycling at I = 0.1 A∙g−1, the specific capacity increased from 250 to 339 m Ah∙g−1 in 34 cycles, and then it decreased to 160 mAh∙g−1 over 70 cycles. Increasing the current density to 1.0 A∙g−1 resulted in a decrease in capacity, but the increase in stability at the high current density (the capacity retention over 500 cycles was 80%) was explained by the rapid dissolution of manganese at low current densities and the inhibition of this process at the high discharge rates.
Trivalent and tetravalent ions intercalated into the MnO2 crystal lattice have a crucial effect on the stability of the crystal lattice and oxygen defects. Large Sn+4 ions significantly improved the conductivity, but they also destabilized the lattice of MnO2 and promoted the Mn2+ dissolution. On the other hand, trivalent ions have strong chemical bonding with oxygen atoms, and this stabilizes the structure and, in the case of Al3+, improves the conductivity, so Al-doped MnO2 materials are the most attractive for further development.
The above reviewed results on the electrochemical properties of the metal-doped MnO2-based cathode materials are summarized in Table 1. The observed effects are related to the improved stability of the materials and the enhanced reversibility of the intercalation processes due to the increased interlayer/channel spacing, which favors ionic diffusion. The stabilizing effect of the introduction of metal ions (probably combined with H2O molecules) into the host structure of MnO2 adjusts the crystal spacing for easier insertion/extraction of H+ and Zn2+ ions.

3. Manganese-Oxide-Conducting Polymer Composite Cathodes for AZIBs

Intrinsically conducting polymers are promising materials for improving the electrochemical performance of cathodes due to their high electronic and ionic conductivity, chemical inertness, and low solubility in aqueous media [61,62]. The most commonly used conducting polymers are polyaniline, polypyrrole, and poly(3,4-ethylenedioxythiophene) due to the many possible polymerization routes that can be easily controlled. It should also be noted that conducting polymers could be used to coat the electroactive grains or to intercalate polymer fragments into the manganese dioxide lattice. On the other hand, it has been reported that coating the zinc anode with conducting polymers also stabilizes the Zn//MnO2 cell [63].

3.1. Polyaniline-Modified MnO2 Cathodes

Polyaniline (PANI) is a frequently investigated conducting polymer in the field of energy storage materials due to several advantages, such as ease of synthesis, environmental friendliness, low cost, and high conductivity [64]. Generally, composites of MnO2 with PANI have been prepared by chemical polymerization reaction between aniline and potassium permanganate or by self-polymerization of aniline in the presence of MnO2 in acidic environment. Such syntheses allow researchers to obtain polymer coatings on the surface of electroactive grains, interfacial 3D networks, and PANI-intercalated materials.
One of the ways to obtain PANI–manganese-oxide composites is the reaction between KMnO4 and aniline without the addition of an acid. In the case of an interfacial gas/liquid reaction of evaporated aniline and the solution of potassium permanganate, layered MnO2@PANI nanohybrids were obtained by layer-by-layer formation [65]. During the reaction, black powder was formed on the surface of the KMnO4 solution. The XRD pattern of the sample exhibited several peaks that are consistent with a 2H birnessite structure. From the XPS spectra of the manganese, it was confirmed that the valence state of Mn is close to +4, and the nitrogen XPS spectra correspond to a PANI structure. The weight percentage of PANI in the composite was ≈4.5% according to the TGA analysis. The morphology of the MnO2@PANI composites, as observed from the large-scale SEM images, was a 3D mesoporous framework with a high surface area (255 m2∙g−1 from the BET measurements). High-resolution TEM characterization revealed two phases: the ordered core associated with MnO2 and the amorphous shell consisting of PANI. The lattice spacing in the MnO2 was close to 0.37 nm, which correlated well with the XRD patterns. The electrochemical properties of the composite cathodes were evaluated in an aqueous solution of 2 M ZnSO4 with an additive of 0.2 M MnSO4. At the low current density of 0.1 A∙g−1, MnO2@PANI nanohybrids delivered a specific capacity as high as 310 mAh∙g−1, and 109 mAh∙g−1 at 2.0 A∙g−1. The cycling stability of the MnO2@PANI nanohybrids was evaluated at 0.5 A∙g−1 for 500 cycles. For the first 20 charge/discharge cycles, a gradual increase in capacity was observed; then, after 100 cycles, the capacity fading was 30%; and then the capacity stabilized. After 500 cycles, the electrode material exhibited a specific capacity of 188 mAh∙g−1. Such a good electrochemical performance of MnO2@PANI can be explained by the polyaniline surface coating, which prevents the dissolution and phase transformation of MnO2 during cycling. The increase in capacity during the first cycles can be related to the presence of Mn2+ ions in the electrolyte solution and the electrodeposition of a MnOx electroactive layer on the electrode surface.
Another method to carry out the chemical polymerization reaction is the interfacial process between reagents in aqueous and organic media (aniline solution in CCl4 and aqueous solution of KMnO4) (Figure 9a). This process allowed researchers to obtain PANI-intercalated manganese dioxide [66]. The mechanism of this process involves the oxidation of the aniline monomer, reduction of the permanganate anion, and diffusion of PANI into the crystal lattice of MnO2, resulting in a layer-by-layer formation of a composite material with a mesoporous structure. The composite material obtained had a porous structure with polycrystalline fragments, as observed from the TEM data. The HR-TEM study after heat treating showed that the interlayer distance in the MnO2 structure was expanded to nearly 1.0 nm. From the XRD patterns, it was concluded that the material was highly amorphous, with many large crystalline fragments. This was related to the expansion of the MnO2 structure by PANI. At the same time, small diffraction peaks at 2θ 12.04, 36, and 65° were associated with the birnessite structure (or δ-MnO2). The specific surface area of the composite was 277 m2∙g−1. Thus, the reaction between aniline and potassium permanganate allows researchers to obtain composites with a mesoporous morphology and high specific surface area. The electrochemical properties of this composite material were tested in aqueous 2 M ZnSO4//0.1 M MnSO4 solution. In the first discharge cycle at 0.05 A∙g−1, two main regions were observed: a sloping discharge profile from 1.55 to 1.33 V with a capacity of 50 mAh∙g−1 and the voltage plateau at E = 1.36 V with a capacity of 210 mAh∙g−1 (Figure 9b). In the second cycle, the capacity of PANI-intercalated MnO2 was 298 mAh∙g−1, which is close to the theoretical value (308 mAh∙g−1). Even at the high current density (3.0 A∙g−1), this material still delivered the specific capacity as high as 110 mAh∙g−1. Over 200 cycles at a current density of 0.2 A∙g−1, the capacity retention was close to 90%, with excellent coulombic efficiency. It has been suggested [66] that the reinforced structure of manganese oxide, achieved via intercalation of polyaniline oligomers between MnO2 layers, is more stable to H+/Zn2+ co-insertion, and no collapse of the structure occurred during repeated insertion/extraction of hydrated cations. It was also demonstrated that PANI-intercalated MnO2 can be successfully cycled in solution without the manganese sulfate additive. Capacity decay was observed during several initial cycles due to manganese dissolution; however, from the 5th to the 200th cycle, the capacity decay was very slow. Thus, PANI fragments between MnO2 layers allow researchers to stabilize the crystal lattice and increase the cycle life of the composite materials.
The development of composite materials with conducting polymers and carbon additives is aimed at improving the overall electronic and ionic conductivity of MnO2 and its mechanical stability. A polyaniline-coated aerogel of MnO2 and reduced graphene oxide (rGO), MnO2/rGO/PANI, was prepared by a two-step process of mechanical mixing of MnO2 and rGO and hydrothermal treatment, followed by in situ polymerization of aniline at a low temperature in the presence of ammonium persulfate as oxidant in hydrochloric acid [67]. The resulting MnO2/rGO/PANI aerogel had a dense and compact structure, and the uniform distribution of Mn, O, C, and N elements demonstrated by the EDX analysis confirmed the complete coating of the aerogel by PANI. The specific surface area of the composite was 142 m2∙g−1. The XRD patterns of the MnO2/rGO/PANI composite material clearly showed that β-MnO2 retained its crystal structure throughout the synthetic operations (Figure 9c). An additional broad peak at 26° on the XRD patterns of MnO2/rGO and MnO2/rGO/PANI composites was attributed to the presence of graphene oxide, and the resulting composite was more amorphous due to the presence of PANI. The XPS characterization confirmed the structure and composition of the aerogel. The electrochemical properties of the composite were evaluated in the aqueous 2 M ZnSO4 electrolyte without Mn2+ additive, and at the current density of 0.1 A∙g−1, the MnO2/rGO/PANI delivered the specific capacity of 241 mAh∙g−1, while at 1.0 A∙g−1, the reversible capacity value of 111.7 mAh∙g−1 was achieved (Figure 9d). Both values were 45% higher than those of a pristine MnO2 or MnO2/rGO composite. Furthermore, the capacity decay of the ternary composite was the lowest (13.3% per 600 cycles). The suppression of the manganese dissolution during the charge/discharge by the polyaniline layer is the main reason for the high electrochemical performance. An evaluation of the diffusion rate by GITT showed high diffusion coefficients in the range of 10−9–10−11 cm2∙s−1, which is two orders higher than that of binary composite or pristine MnO2. Such high values were achieved due to the high electronic and ionic conductivity of rGO and PANI, which accelerate the diffusion of Zn2+ ions.
MnO2 that was obtained hydrothermally on carbon cloth substrate was dispersed in acidic aniline solution for several hours to obtain PANI-MnO2/CC three-dimensional networks [68]. During the synthesis, MnO2 nanosheets were formed on the carbon cloth’s surface. Initially, MnO2/CC 3D structures were formed by growing MnO2 nanosheets on the carbon cloth’s surface, and the addition of PANI led to an increase in the roughness of the surface. According to the EDX mapping of Mn, O, C, and N, the inorganic components and PANI coating were homogeneously dispersed in the composite. From the HR-TEM and XRD measurements, it was found that layered δ-MnO2 was formed during the synthesis. The valence state of manganese (Mn+4) and the polymer structure were also confirmed by XPS and Raman spectroscopy. Charge/discharge tests of the PANI-MnO2/CC binder-free electrode in the mixed 2 M ZnSO4/0.1 M MnSO4 electrolyte showed an excellent specific capacity of 286 mAh∙g−1 at a current density of 0.5 A∙g−1, and at the high current density of 4.0 A∙g−1, the delivered capacity was 177 mAh∙g−1 (Figure 9e). Such large capacity values can be attributed to the high electrical and ionic conductivity of PANI, the high porosity of the 3D networks with a greater number of adsorption sites, and the pseudocapacitive charge storage in the PANI itself, as seen in the charge/discharge profiles (Figure 9f). The capacity retention of the PANI-MnO2/CC composite was 90.2% after 1800 cycles at a current density of 2.0 A∙g−1, and 96.9% after 9000 cycles at I = 4.0 A∙g−1, based on the initial specific capacity value. The changes in capacity during continuous cycling were non-linear: first, a decrease in capacity was observed, and then the capacity increased and decreased back to near the initial values. Such changes are associated with the presence of Mn2+ ions in the electrolyte solutions. This high stability allowed the authors [68] to conclude that the PANI coating prevents the dissolution of manganese during cycling, while the artificial increase in the specific capacity occurs due to the Mn2+-Mn+4 redox reaction and the formation of MnOx. PANI-MnO2/CC also exhibited the lowest charge-transfer resistance, as compared to MnO2/CC and pristine PANI (Figure 9g). The slope in the low frequency region of the EIS spectra of PANI-MnO2/CC was higher than that of MnO2/CC and PANI, indicating that the PANI effectively enhances the zinc-ion diffusion rate in the composite. The Zn2+ diffusion coefficient of the PANI-MnO2/CC (2.33·10−11 cm2·s−1), as determined by GITT, was higher than that of MnO2/CC (5.06·10−12 cm2·s−1), indicating a faster Zn2+ transport rate in PANI-MnO2/CC.
The in situ polymerization of PANI on the MnO2 ultralong nanowires was proposed in [69] to obtain a core–shell MnO2@PANI composite material, which was used as a binder-free cathode for AZIBs. As a result of the redox process between aniline monomer and MnO2, ultralong nanowires interconnected in a highly porous architecture were obtained. No significant changes in the morphology of the material were observed after polymerization, so the PANI coating was considered homogeneous. Based on HR-TEM and XRD patterns, tetragonal α-MnO2 was obtained during the hydrothermal synthesis and was retained after aniline polymerization. The structure of PANI was confirmed by FTIR and Raman spectroscopy, and the amount of polymer coating was ≈10.5% according to TGA analysis. XPS measurements also confirmed the structure of PANI due to the presence of nitrogen peaks in the survey spectra and the +4 valence of manganese. The electrochemical measurements were carried out in an aqueous solution of 2 M ZnSO4/0.1 M MnSO4. The polarization value for the MnO2@PANI composite electrode was lower than that of the electrode prepared from unmodified MnO2 nanowires. At the low current of 0.2 A∙g−1, the capacity at the 50th cycle was 343 mAh∙g−1 due to MnOx electrodeposition during the cycling process. At the high current density of 3.0 A∙g−1, the delivered capacity was 100 mAh∙g−1. The long-term stability of the MnO2@PANI hybrid cathodes was investigated at I = 0.5 and I = 2.0 A∙g−1 over 300 and 2000 cycles, and the capacity retention was close to 80% regardless of the applied current. It was suggested that the main role of the PANI coating is to improve the mechanical integrity and conductivity of the composite material.
Thus, PANI-coated MnO2 cathodes for AZIBs have better stability because PANI prevents the dissolution of Mn2+ from the electrode surface and provides higher conductivity within the electrode material. It is interesting to note that PANI can be intercalated into the layered MnO2, as has been described for vanadium-based materials [70,71,72]. The intercalation of PANI between the crystal lattice layers also improves the mechanical stability.

3.2. Polypyrrole-Modified MnO2 Cathodes

Polypyrrole (PPy) is a conducting polymer that is widely used in composite materials due to its low solubility in various solvents, large surface area, high conductivity, non-toxicity, and chemical and electrochemical stability [73]. In particular, PPy can be used in energy storage materials [74,75], and it is more stable compared to PANI [76]. Like any other conducting polymer, the properties of PPy depend on the polymerization reaction conditions and methods [77]. PPy is typically used as a coating on the MnO2 surface, which is obtained by different methods, including self-polymerization and chemical or electrochemical oxidation of pyrrole monomer.
The chemical polymerization of pyrrole in solution with dispersed MnO2 was performed at low temperatures by adding aqueous FeCl3 solution to obtain a MnO2/PPy composite (Figure 10a) [78]. The composite maintained the morphology of the initial MnO2 nanorods (Figure 10b). Based on density functional theory calculations and HR-TEM images, it was concluded that the crystal lattice of the MnO2/PPy composite is more disordered than that of pristine MnO2. An amorphous coating of PPy was detected on the tetragonal α-MnO2 nanorods by the XRD and TEM. Strong interactions between the MnO2 lattice and the pyrrole rings were confirmed by the presence of the Mn-N bond in the XPS spectra of N 1s at 399.2 eV (Figure 10c) and additionally confirmed by DFT calculations. The electrochemical performance of the MnO2/PPy cathode material was evaluated in 2 M ZnSO4 aqueous solution in the presence and absence of 0.1 M MnSO4. The remarkable specific capacity of 256 mAh∙g−1 was observed after 100 cycles at I = 0.1 A∙g−1, with a gradual increase of the capacity, while for bare MnO2 nanorods, the capacity fading under the same conditions was 45%. The MnO2/PPy composite showed a high rate performance (90 mAh∙g−1 at 1.5 A∙g−1). Moreover, in the absence of the MnSO4 additive, 66% capacity retention was observed for the composite material (Figure 10d). One of the explanations for the increase in capacity is the additional pseudocapacitive charge storage in the PPy coating. The study of manganese dissolution by ICP analysis had shown that it was mitigated by the presence of a protective PPy shell on the surface of MnO2 grains that was chemically bound to the MnO2 core. The kinetic parameters (activation energy and diffusion coefficients) obtained by impedance spectroscopy and cyclic voltammetry showed that the diffusion of Zn2+ ions was more effective for the MnO2/PPy composite.
Pre-synthesized PPy nanowires were added to MnSO4 solution in the presence of ammonium persulfate to obtain a β-MnO2/PPy composite [79]. The diffraction peaks of the as-synthesized product corresponded well to tunnel-type β-MnO2. The morphology of the β-MnO2/PPy composite was described as microspheres formed by PPy nanowires and MnO2 nanorods. However, the specific surface area of the composite calculated from BET measurements was only 47.3 m2∙g−1, which is much lower than for other composites with PANI [65]. The valence state of manganese (+4) was calculated from XPS spectra, and the interactions between N and Mn elements were also observed. The rate performance studied in conventional electrolyte (2 M ZnSO4/0.1 M MnSO4) showed that the specific capacity of MnO2/PPy was 215 mAh∙g−1 at 0.1 A∙g−1 and 105 mAh∙g−1 at 1.0 A∙g−1, which is 30–40% higher than that of unmodified MnO2. The specific capacity during long-term cycling at 0.2 A∙g−1 (160 cycles) gradually increased due to the MnOx electrodeposition during cycling, while the capacity degradation for unmodified manganese oxide was about 50%. Such advantages of the MnO2/PPy composite may be related to the randomly oriented morphology of the material, which consists of PPy nanowires and MnO2 nanorods. Such porous architecture of the composite material, which improves its ionic conductivity, facilitates Zn2+ and H+ diffusion into the material. In addition, highly conductive PPy nanowires located between MnO2 nanorods provide better electronic contact within the composite material.
Composite materials with polypyrrole and carbon additives have also been used for MnO2-based cathodes to improve their electrical conductivity. Ternary composites based on MnO2 with carbon nanotubes (CNTs) coated with PPy have been successfully applied in AZIBs [80,81] with superior properties. PPy was obtained by in situ oxidation of pyrrole by MnO2 nanowires in the presence of CNTs (CNT/MnO2-PPy). As a result, α-MnO2 was obtained, and its structure was preserved after the polymerization of pyrrole. The X-ray diffraction data with a peak at 25.84° indicated the presence of PPy in the composite, and the lower intensity of the XRD peaks suggested that the MnO2 in the composite material was more amorphous than the initial oxide. A three-dimensional interconnected network, in which all components are uniformly distributed, was formed during the reaction. The morphology of the CNT/MnO2-PPy composite was porous, with macropores of several micrometers in size. Highly dispersed MnO2/PPy nanowires were surrounded by CNT networks. The cycling stability and rate performance of the CNT/MnO2/PPy cathode material were investigated in the conventional electrolyte with a Mn2+ additive. At the low current density (0.1 A∙g−1) the specific capacity value was ≈250 mAh∙g−1, and at I = 2.0 A∙g−1, it was 84.2 mAh∙g−1. The cycling stability was investigated at the current densities of 0.3 A∙g−1 (200 cycles) and 1.0 A∙g−1 (1000 cycles). In both cases, the capacity decay was less than 15% [81]. At I = 0.3 A∙g−1, the initial gradual increase in capacity was observed. The kinetic investigations showed that the charge-transfer resistance of the ternary composite is lower, and the diffusion coefficient is twice as high for both H+ and Zn2+ cations. These improved properties can be explained by the high porosity of the network, the conductivity of the CNTs and PPy, and the formation of a protective layer that prevents manganese dissolution. Such a unique architecture provides abundant sites for Zn2+ and H+ adsorption, resulting in additional pseudocapacitive charge storage.
Another carbon additive, graphene oxide, has also been used to develop ternary composites with MnO2 and PPy [82]. For this purpose, an α-MnO2/rGO composite was prepared hydrothermally from MnSO4, KMnO4, and graphene oxide. Then the chemical polymerization of PPy was carried out in a solution of as-prepared α-MnO2/rGO material in an acidic environment, with ammonium persulfate as the oxidizing agent. No significant structural changes occurred after the polymerization: the position of the main peaks in the XRD patterns of pristine α-MnO2 and α-MnO2/rGO-PPy did not change. The composites had the morphology of nanowires coated with conducting polymer, forming continuous conductive networks. As seen from the XPS spectra, the valence state of manganese is dual: Mn3+ and Mn4+, which can be related to the oxidation of pyrrole not only by (NH4)2S2O8 but also by MnO2 itself (Figure 10e). The cathode material based on the ternary α-MnO2/rGO/PPy composite delivered 438 mAh∙g−1 at I = 0.1 A∙g−1 and 190.2 mAh∙g−1 at 1.0 A∙g−1. Such a high capacity can be explained by the conductive network formed by rGO and PPy and also by the pseudocapacitive properties of both components. The capacity decay for the ternary composite was the lowest (14.1%), while for the binary α-MnO2/rGO composite, it was 36.5%. Thus, the main contributor to the electrochemical performance improvement is the carbon additive, while the conducting polymer enhances its benefits.
The combination of two different manganese oxides with the PPy layer has also been discussed as a way to improve the electrochemical performance of Zn//MnO2 batteries. The molten salt synthesis of MnO2/Mn2O3 (Mn-O-3) nanocomposite and the self-polymerization of PPy were performed in [83]. During the synthesis, nanobelts and nanoparticles of Mn-O-3 coated with a dense shell of amorphous polypyrrole were obtained. According to the XRD patterns, the nanocomposite consisted of two oxides (α-MnO2 and Mn2O3), and more diffraction peaks of α-MnO2 could be indexed. Based on the ICP analysis, the content of Mn2O3 in the composite was about 34.1%. Depending on the synthesis time, the weight content of the PPy shell in the composite varied from 0.64% to 1.96%, as was shown by TGA. At the low current density of 0.2 A∙g−1, the specific capacity of cathodes with a Mn-O-3 composite with thin PPy coating was 289.8 mAh∙g−1, while at I = 3.0 A∙g−1, the capacity was 199.8 mAh∙g−1. Increasing the thickness of the PPy coating resulted in an insignificant decrease in capacity over the entire current range. These excellent values were explained by the high conductivity of the PPy film (21.4 S∙cm−1), which is much higher than that of MnO2. The cycling stability was evaluated in 2 M ZnSO4 electrolytes with/without manganese sulfate additive. At I = 3.0 A∙g−1 in a 2 M ZnSO4/0.2 M MnSO4 electrolyte, the capacity retention during 1000 cycles was over 100% due to the capacity increase caused by the formation of electroactive MnOx layer. In pure 2 M ZnSO4 solution at I = 1.0 A∙g−1, the capacity retention over 1000 cycles was 80.5%, which is a remarkable value. It is related to the protective PPy layer on the Mn-O-3 composite surface, which prevents manganese dissolution during the cycling process. Kinetic studies showed that the polarization of the cells and the charge-transfer resistance decreased for PPy-coated Mn-O-3, while the diffusion of Zn2+ and H+ was faster than for the uncoated Mn-O-3 composite.
Conducting polymer coatings can also be applied to modified oxides, e.g., for oxides doped with heterovalent atoms. Iron-doped MnO2 coated with PPy was proposed as a cathode in [84]. First, Fe3+-doped MnO2 was prepared from Mn(CH3COO)2, KMnO4, and a small amount of Fe(NO3)3 via the chemical precipitation method. Then the self-polymerization of PPy was performed at 0 °C for 5 h. The samples had a highly porous morphology consisting of microspheres. Based on the EDX analysis, uniform distribution of Mn, O, C, N, and Fe elements was observed, so it was concluded that the uniform polymer coating was formed on the whole surface. The specific surface area of the material was 56 m2∙g−1. A tetragonal α-MnO2 crystal structure was found from the XRD patterns, and for Fe3+-doped α-MnO2 samples, the shift of the diffraction peak at 2θ = 37.52° to lower 2θ angles was probably due to the partial substitution of Mn4+ by Fe3+. The PPy-coated sample was more amorphous, but the α-MnO2 crystal structure was maintained. As shown by XPS, the valence state of manganese and iron was +4 and +3, respectively, so Fe3+ ions did not react with pyrrole monomer or manganese precursors. The initial capacity of the Fe-doped MnO2@PPy composite at a current density of 0.1 A∙g−1 was 270 mAh∙g−1. However, at the high current density of 1.0 A∙g−1, it was only 73 mAh∙g−1. The capacity fading over 100 cycles at I = 0.1 A∙g−1 was ≈75%. This electrochemical performance is mainly attributed to the increased interlayer spacing of manganese oxide and oxygen defects due to the substitution of Mn+4 to Fe3+. At the same time, the PPy coating improves the conductivity of the composite. Thus, the iron doping and PPy conductive network can not only facilitate zinc ion migration and electron transport but also reduce manganese dissolution from the material.
Finally, the MnO2/PPy composite was used for the development of flexible zinc-ion batteries [85]. For this purpose, the cathode was electrodeposited on stainless-steel yarns from manganese acetate solution in the presence of sodium sulfate. According to the XRD patterns, α-MnO2 was the main crystalline phase of the electrodeposited cathode material. After the deposition of MnO2, the surface of individual stainless-steel yarn fibers became rougher, and after the coating of MnO2@SS with PPy, the morphology was further changed, and the higher surface area of the resulting material was favorable for the charging/discharging process. The presence of PPy in the composite was confirmed by Raman spectroscopy. To assemble the cell, the cathode yarn was wound around a Zn@Nitinol wire, which served as the anode. Gelatin–borax flexible gel electrolyte containing 1 M ZnSO4/0.1 M MnSO4 encapsulated the electrodes and served as a separator (Figure 10f). The rate performance tests showed that the specific capacity of this battery was 174.2 mAh∙g−1 at 0.5 C and 60 mAh∙g−1 at 4 C, and the capacity retention at 2 C over 1000 cycles was >60%. In addition, this battery was stable to mechanical deformation, such as bending, without significant capacity degradation. This was probably achieved due to the PPy coating on the surface of the MnO2-electrode, which provided not only electrical but also mechanical support for the active material and improved the conductivity and flexibility of the cell.
Figure 10. (a) Schematic diagram of synthesis of MnO2/PPy composite [78]. (b) SEM image of MnO2/PPy nanorod [78]. (c) High-resolution N 1s XPS spectrum of MnO2/PPy composite [78]. (d) Continuous cycling performance of MnO2/PPy electrode in 2 M ZnSO4 aqueous electrolyte [78]. (e) High-resolution Mn 2p and N 1s XPS spectra of α-MnO2/rGO-PPy composite [82]. (f) The construction of the flexible Zn//MnO2 device with shape memory function [85].
Figure 10. (a) Schematic diagram of synthesis of MnO2/PPy composite [78]. (b) SEM image of MnO2/PPy nanorod [78]. (c) High-resolution N 1s XPS spectrum of MnO2/PPy composite [78]. (d) Continuous cycling performance of MnO2/PPy electrode in 2 M ZnSO4 aqueous electrolyte [78]. (e) High-resolution Mn 2p and N 1s XPS spectra of α-MnO2/rGO-PPy composite [82]. (f) The construction of the flexible Zn//MnO2 device with shape memory function [85].
Energies 16 03221 g010
Thus, polypyrrole-modified MnO2-cathodes have advantageous properties due to the chemical bonding of the MnO2 core and the PPy shell, the high conductivity of the polymer, and the simplicity of the synthesis. However, the porosity and surface area of these composites are lower than those of PANI-modified cathodes, and PPy-intercalated MnO2 materials have not yet been obtained.

3.3. Poly(3,4-Ethylenedioxythiophene)-Modified MnO2 Cathodes

Poly(3,4-ethylenedioxythiophene) (PEDOT) is a one of the most stable derivatives of polythiophene due to the presence of two alkoxy substituents attached to the 3- and 4- positions of the thiophene ring. PEDOT itself and the aqueous dispersion of PEDOT with polystyrene sulfonate (PEDOT:PSS) are widely used for various applications, such as electrode materials for secondary power sources, solar cells, electrochromic devices, etc. [86,87]. PEDOT exhibits high stability in the doped state and excellent conductivity [88,89]; thermal, chemical, and electrochemical stability; and insolubility in aqueous and organic media. Typically, PEDOT-modified materials are obtained by electrodeposition or chemical oxidation of the monomer. PEDOT:PSS can be applied as received.
For manganese-based cathodes for AZIBs, the chemical or electrochemical deposition of manganese compounds, followed by deposition of the polymer, is a popular method to prepare MnO2/PEDOT composites [90,91,92], which can be further used as binder-free electrodes. In several cases, MnO2-based materials have also been electrodeposited on the substrate. A flexible quasi-solid battery with electrodeposited MnO2@PEDOT cathode, zinc anode, and 2 M ZnCl2/0.4 M MnSO4 gel polymer electrolyte based on polyvinyl alcohol was developed in [90]. During the synthesis, polycrystalline α-MnO2 coated by the amorphous polymer layer was obtained. The EDX analysis showed the uniform distribution of Mn, O, C, and S elements, thus indicating that the composite coating was homogeneous. The predominant valence state of manganese was determined as Mn4+ from the XPS spectra. The electrochemical performance of the composite was evaluated in liquid aqueous and gel polymer electrolytes. In the liquid aqueous electrolyte, the specific capacity was 366.6 mAh∙g−1 at a current density of 0.74 A∙g−1 and 143 mAh∙g−1 at I = 7.43 A∙g−1. This excellent performance can be explained by the thin MnO2 layer, the significant effect of the additional MnOx deposition, and the protective properties of the PEDOT layer on the composite surface. The cycling stability of the composite at a current density of 1.11 A∙g−1 was better in the presence of PEDOT due to the alleviation of manganese dissolution. The capacity retention of the composite electrode after 300 cycles was 83.7% compared to 47% for pristine MnO2. It was also found that PEDOT was not involved in the electrochemical reaction, and the full capacity of the electrode was provided by MnO2. The functional properties of the flexible quasi-solid cell with gel electrolyte were also remarkable: the specific capacities of 284 mAh∙g−1 at I = 0.37 A∙g−1 and 76 mAh∙g−1 at I = 5.58 A∙g−1 and capacity retention of 77% over 300 cycles at I = 1.86 A∙g−1. The capacity decrease at high current densities was explained by the increase in charge-transfer resistance in the gel polymer electrolyte. The capacity degradation was associated with the slow dissolution and destruction of the composite material.
The above work [90] was further developed by synthesizing MnO2/PEDOT on carbon cloth modified with carbon nanotubes (CMOP) [91], which were chosen for their high electrical conductivity. CNTs were deposited on the carbon cloth surface via the chemical vapor deposition method, followed by electrodeposition of MnO2 and PEDOT. The morphology of the composite was described as a coaxial-cable structure based on interconnected CNTs, with MnO2 and a rough, thick layer of PEDOT deposited on each CNT array (Figure 11a). The crystal structure of the CMOP composite was α-MnO2, as previously observed in [90]. The addition of CNTs to the composite allowed the researchers to achieve the specific capacity of 306.1 mAh∙g−1 at I = 1.1 A∙g−1, and at the high current density of 10.8 A∙g−1, the capacity was 176.8 mAh∙g−1. In addition, it was shown that the capacity retention of the Zn//MnO2 cell is not only determined by the cathode material, but also by the zinc anode: with the same cathode, replacing the zinc anode with the fresh one led to an increase in capacity and stability after 1000 cycles (Figure 11b) [91]. After the electrochemical tests, cracking and pulverization were observed on the SEM images of the MnO2/CNT electrode, while only negligible tiny cracks were observed for PEDOT-coated material. The capacity retention of the PEDOT-modified electrode was 71% before and 81% after replacement of the zinc anode. A flexible quasi-solid-state battery was assembled by sandwiching an aqueous poly(vinyl alcohol) (PVA) hydrogel electrolyte between the CMOP cathode and Zn anode. The device demonstrated high-capacity retention (77.0%) after 500 cycles at a high current density of 5.4 A g−1 and close to 100% Coulombic efficiency (Figure 11c).
Another way to obtain MnO2/PEDOT composites is the chemical deposition of MnO2 on a conductive substrate, followed by electrodeposition of the conducting polymer, which has been used for Co2+-doped MnO2 grown on nickel foam [92]. In this work, a Co-MnO2 composite was prepared by precipitation from an acidic manganese acetate solution with a cobalt acetate additive, which was treated with sodium hydroxide and potassium persulfate with further annealing. The PEDOT layer was deposited potentiostatically from the solution containing EDOT, SDS, and LiClO4 as an electrolyte. The Co-MnO2 product consisted of uniform nanoflakes that were horizontally interconnected to form a highly open and porous structure. The presence of Co2+ ions increased the dimensions of the flakes, resulting in more effective electronic and ionic contact between the particles. The resulting PEDOT@Co-MnO2 material had a rough surface with irregular bulges. The main elements (Mn, O, Co, and S) were uniformly distributed in the composite material. No clear diffraction peaks were detected in the XRD patterns, indicating that the material was highly amorphous. The electrochemical properties of PEDOT@Co-MnO2 were investigated in 2 M ZnSO4 aqueous solution without manganese additive. Nevertheless, the specific capacity of PEDOT@Co-MnO2 was 298.9 mAh∙g−1 at I = 1 A∙g−1, while at I = 10 A∙g−1, it was 150 mAh∙g−1. The absence of the Mn2+ additive explained the abrupt capacity drop for MnO2 and Co-MnO2 materials during continuous cycling, while for the PEDOT-coated composite, the capacity fading over 1000 cycles was only 7.7%. Thus, in the case of the PEDOT@Co-MnO2 material, the enlarged crystal lattice due to Co doping and conducting and protective PEDOT layer mitigate the structural deformations and facilitate Zn2+ and H+ transfer in the material.
An interesting approach to obtain a PEDOT-modified MnO2-based cathode by electrodeposition of polymer on the electrode prepared by mechanical mixing of conventional components was proposed in [93]. For this purpose, α-MnO2 and δ-MnO2 were synthesized separately and then mixed in a 1:2 weight ratio to further prepare the electrode material by mixing manganese oxides with Super P carbon and polyvinylidene fluoride in a 7:2:1 weight ratio. Then the PEDOT layer was coated on the electrode surface by electrodeposition to prepare the PEDOT-coated electrode (DMOP). In the composite electrode, the particles of δ-MnO2 were homogeneously distributed and interconnected by α-MnO2 nanowires under the dense and smooth PEDOT layer. The EDX analysis showed a uniform distribution of Mn, O, C, and S elements. Structural analysis showed that no changes were detected after the electrode material preparation. The electrochemical study of the electrodes in 2 M ZnSO4/0.1 M MnSO4 aqueous electrolyte showed that pristine δ-MnO2 electrodes were unstable due to manganese dissolution and structural collapse, and the addition of a tunnel-type α-MnO2 stabilized the electrode. The addition of PEDOT further improved the electrochemical performance of the cathode material (360.5 mAh∙g−1 at 0.1 C and 94 mAh∙g−1 at 5 C). The capacity fading of the DMOP electrode over 900 cycles at 1 C was ≈10%. Based on the XPS spectra and DFT calculations, the stronger binding effect of Zn2+ was demonstrated for the composite with PEDOT.
The electrodeposition technique has been applied not only to MnO2 cathodes but also to oxygen-deficient ZnMn2O4 (ZMO) to fabricate flexible zinc-ion batteries [94]. It was suggested that the introduction of oxygen vacancies could increase the electrostatic repulsion of Zn2+ within the lattice and thus should facilitate the diffusion of Zn2+ ions from ZMO. PEDOT was deposited on the surface of the grains as a protective layer. The cathode material was prepared by a three-step process: electrodeposition of ZnMn2O4 on the carbon cloth, annealing in the reducing H2 atmosphere to extract oxygen anions, and electrodeposition of PEDOT on the composite surface. The structure and morphology of the resulting composite material were investigated by transmission and scanning electron microscopy. The dense material layer was observed after ZMO deposition, and it was maintained after the second deposition process. Based on the XPS results, the enrichment of oxygen vacancies after the reduction by hydrogen was identified. The ratio of manganese valence states Mn2+:Mn3+ was 1:3, so the ratio of oxygen vacancies was 8.6%. Electrochemical tests were performed in 1 M ZnSO4 aqueous solution and in gel electrolyte based on polyvinyl alcohol. The oxygen-deficient ZMO@PEDOT composite delivered the specific capacity of 221 mAh∙g−1 at the low current density (0.5 mA∙cm−2), while for ZMO without the polymer layer, it was only 174 mAh∙g−1, and at the high current density (10 mA∙cm−2), the capacity of ZMO@PEDOT composite was 62.5 mAh∙g−1. It was shown that the main effect on the electrochemical performance was due to the structural disorder caused by oxygen vacancies, while PEDOT had no electroactivity in this composite. The capacity decay for the ZMO@PEDOT electrode material over 300 cycles at I = 8 mA∙cm−2 was only 6.2%. The anion-deficient structure was also studied by DFT calculations, and it was found that oxygen-deficient ZnMn2O4 has lower energy values, leading to facilitated Zn2+ diffusion kinetics, lower Zn vacancy formation energy, and lower energy barrier of Zn mobility. With the presence of conducting protective PEDOT layer, the remarkable performance is observed, even for flexible battery with lower capacitive properties: 207 mAh∙g−1 at 0.5 mA∙cm−2 and 20% of capacity decay over 300 cycles at 8 mA∙cm−2.
Chemical oxidation of EDOT monomer in the presence of KMnO4 is an alternative route to prepare MnO2/PEDOT composites, which was applied in [95], where the layered-type MnO2/PEDOT composite was prepared by the precipitation method at a low temperature. The MnO2 nanowires were coated by PEDOT, with some voids between the nanowires. The XRD diffraction patterns confirmed the layered structure of MnO2, and the structure of PEDOT was evaluated by FTIR and Raman spectra. XPS and EDX analyses confirmed the material composition and the uniform distribution of elements in the composite material. The electrochemical properties were investigated in a 2 M ZnSO4/0.2 M MnSO4 solution, and it was shown that the PEDOT layer on the MnO2 surface prevents manganese dissolution, resulting in superior performance. At the low current density of 0.2 A∙g−1, the specific capacity was 242 mAh∙g−1, and at I = 3.0 A∙g−1, it was 121 mAh∙g−1. The cycling performance of the MnO2@PEDOT composite at current densities of 1.0 and 2.0 A∙g−1 was more stable due to the reduced manganese dissolution/deposition process, and the capacity retention over 1000 cycles at 2.0 A∙g−1 was 85%. It was also observed that the charge-transfer resistance of the MnO2@PEDOT electrode was much lower than that of the undoped material, which may be related to the enhanced conductivity of the MnO2/PEDOT composite.
PEDOT-intercalated MnO2 (or PEDOT-MnO2) was obtained in [96] by the reaction of KMnO4 and MnSO4 in the presence of 3,4-ethylenedioxythiophene monomer. In this case, KMnO4 oxidized both MnSO4 and EDOT monomer. The morphology of the MnO2 samples was hydrangea-like, with a tendency to form separated layers. The XRD diffraction peaks of the samples were indexed to a layered-type MnO2 without the crystal lattice changes after PEDOT intercalation. However, HR-TEM measurements showed that the interlayer distance increased from 0.68 nm (for pristine MnO2) to 0.73 nm (for PEDOT-MnO2), indicating that PEDOT was inserted into the lattice. The specific surface area of the composite grains was relatively low (72.5 m2∙g−1) due to the introduction of PEDOT, which reduced the grain size. The rate capability and cycling stability were evaluated in an aqueous solution of 2 M ZnSO4/0.2 M MnSO4. The specific capacity of this composite material decreased from 300 mAh∙g−1 at 0.2 A∙g−1 to 122 mAh∙g−1 at 2.0 A∙g−1, but it was fully recovered when low currents were applied after high currents. During continuous cycling at 2 A∙g−1, no capacity decay was observed for the first 20 cycles, then insignificant capacity decay and gradual increase were observed for 1200 cycles, and then negligible capacity decay was observed for the last 300 cycles, with excellent coulombic efficiency. The intercalation of PEDOT expands the interlayer spacing of MnO2, which improves the diffusion kinetics of the material, resulting in a more stable and enhanced electrochemical performance.
The simplest way to modify electrodes with PEDOT is to drop-cast PEDOT:PSS onto the material surface or to add PEDOT:PSS or as-synthesized PEDOT to the electrode slurry [97,98,99]. The first method was applied in [98] to develop binder-free MnO2-PEDOT:PSS electrode material on the vertical graphene. Vertical graphene was obtained on the carbon fiber surface by plasma-enhanced chemical vapor deposition, then MnO2 was obtained hydrothermally on the graphene substrate, and finally PEDOT:PSS dispersion was cast on the material surface, followed by vacuum drying. As observed from HR-TEM images and XRD patterns, the initial graphene substrate had a highly porous structure, and after MnO2 synthesis, the pores were occupied by α-MnO2 particles, resulting in a 3D conducting structure coated with PEDOT:PSS. The specific capacities of the composite in a 1 M ZnSO4/0.1 M MnSO4 aqueous electrolyte were 367 mAh∙g−1 at 0.5 A∙g−1 and 148 mAh∙g−1 at 6.0 A∙g−1. The cycling performance of the electrode was good, with a capacity retention of 68% over 1000 cycles at I = 5.0 A∙g−1. These functional properties were related to the electrode composition: the carbon fiber current collector acted as an efficient conductor, the vertical graphene enhanced the charge transfer kinetics of MnO2, and the PEDOT:PSS coating suppressed the dissolution and damage of MnO2 during repeated electrical and mechanical cycling.
Figure 11. (a) Scheme of the CMOP composite cathode synthesis [91]. (b) Cycling stability of CMO and CMOP cathodes (Zn anodes were replaced by fresh ones at the 1001st cycle) [91]. (c) Cycling performance and Coulombic efficiency of the quasi-solid-state Zn–CMOP battery [91]. (d) SEM of PEDOT powder [97]. (e) C-rate capabilities of δ-MnO2-based electrodes. (f) Cycling stability of δ-MnO2-based electrodes at 0.3 A·g−1 [97].
Figure 11. (a) Scheme of the CMOP composite cathode synthesis [91]. (b) Cycling stability of CMO and CMOP cathodes (Zn anodes were replaced by fresh ones at the 1001st cycle) [91]. (c) Cycling performance and Coulombic efficiency of the quasi-solid-state Zn–CMOP battery [91]. (d) SEM of PEDOT powder [97]. (e) C-rate capabilities of δ-MnO2-based electrodes. (f) Cycling stability of δ-MnO2-based electrodes at 0.3 A·g−1 [97].
Energies 16 03221 g011
The preparation of electrode slurry with the addition of chemically polymerized PEDOT or PEDOT:PSS aqueous dispersion was performed by our research group [97]. PEDOT powder (Figure 11d) was obtained via the polymerization of EDOT in the presence of FeCl3, and PEDOT:PSS was used as the coating for hydrothermally synthesized MnO2 powder. Both PEDOT-containing electrode materials had a smoother and denser surface than the pristine MnO2-based cathode. The crystalline phase determined from XRD patterns corresponded to the layered-type δ-MnO2. The electrochemical tests of δ-MnO2/PEDOT and δ-MnO2/PEDOT:PSS in a 2 M ZnSO4/0.1 M MnSO4 aqueous electrolyte are shown in Figure 11e,f. It is clearly seen that MnO2/PEDOT:PSS electrodes had the best electrochemical performance (298 mAh∙g−1 at I = 0.1 A∙g−1), with a gradual capacity increase at low current densities, due to MnOx electrodeposition on the cathode surface. The cycling stability evaluated at 0.3 A∙g−1 was excellent, with a capacity retention of 99% per maximum value (278 mAh∙g−1). The improvement of the electrochemical properties of MnO2 cathodes with conducting polymers can be explained by the increase of both electronic and ionic conductivity of the electrode material due to more conductive media between electroactive grains. In addition, the PEDOT:PSS layer on the electrode surface supports the mechanical integrity and prevents the manganese dissolution during charge/discharge processes.
In conclusion, the introduction of a conducting polymer has a great impact on improving the electrochemical performance of MnO2-based cathodes for AZIBs. The above results on the electrochemical properties of the polymer-modified materials are shown in Table 2. The conducting polymer coating improves the mechanical integrity of the composite electrodes, and they are able to further enhance conductivity and buffer shrinkage/expansion during charge/discharge. Protecting the MnO2 grains from direct contact with the aqueous electrolyte also plays a role in suppressing interfacial side reactions and preventing cathode dissolution.

4. On the Charge–Discharge Mechanism in Rechargeable Zn//MnO2 Batteries

Since the first appearance of rechargeable Zn/MnO2 batteries [8,100], one of the most important issues of these cells has been the mechanism of electrochemical reaction in mildly acidic aqueous electrolyte. The presence of water molecules and Mn2+ ions, along with the presence of Zn2+, in the electrolyte solutions complicates a better understanding of the reactions that occurs.
The basic charge–discharge mechanism of the MnO2 and the transfer (such as intercalation) of Zn2+ ions into the manganese oxide structure are still controversial. So far, four energy storage mechanisms have been proposed (Figure 12) [101,102]:
(i)
Reversible (de)intercalation of Zn2+ ions,
(ii)
Reversible (de)intercalation of H+,
(iii)
Co-(de)intercalation of Zn2+ and H+,
(iv)
Electrolytic deposition/dissolution of MnO2.
In addition, it is worth mentioning that the zinc hydroxide sulfate Zn4(OH)6SO4∙nH2O (ZHS) phase has been found in many works as a by-product formed on the cathode surface during discharge. This chemical precipitation of ZHS occurs due to the local increase in pH near the electrode surface upon the insertion of H+ into MnO2. The ZHS plays a dual role, (a) serving as a passivation layer to inhibit deep dissolution of MnO2 during discharge and (b) promoting the electrochemical deposition kinetics of active MnO2 during charging.
In this section, recent observations on the mechanisms of the charge–discharge processes in pure MnO2 and metal-ion-doped and polymer-modified MnO2 cathode materials are discussed.
Reversible insertion/extraction of Zn2+ ions has been proposed for zinc-ion batteries, as well as for conventional metal-ion batteries, such as lithium-ion or sodium-ion batteries, where reversible intercalation and deintercalation of Li+ or Na+ is the only way to store charge. For example, it was shown by ex situ XRD for K-pre-intercalated MnO2 that the main diffraction peak of the (0,0,3) plane formed reversibly during charging and disappeared during discharging, while the peak corresponding to the (0,0,6) plane shifted to the low-angle region during discharging, which can be associated with reversible Zn2+ intercalation into the δ-MnO2 lattice [28]. Similar results were obtained by EDX method and ex situ XRD and XPS measurements for a ternary Fe/α-MnO2@PPy composite, where the new phase of ZnMn2O4 was formed or disappeared during the electrochemical process [84]. For Al-doped MnO2 cathode material, this mechanism was also observed and confirmed by in situ Raman spectroscopy (the phase transition is shown in Figure 13a) [24]. Nevertheless, the mechanism involving only Zn2+ insertion [16,24,25,28,84] is now considered not plausible due to many novel experimental results, as shown below.
The only insertion/extraction mechanism of H+ has also been established in a number of works [20,103,104,105,106]. Due to the intercalation of H+ ions during the electrochemical reaction, the pH value of the electrolyte solution has a decisive influence on the mechanism of the processes and the resulting electrochemical properties. In different cycles, the reduction and oxidation peaks shifted to more positive potentials with the decreasing of the pH. The pH value influenced the formation of ZHS, Mn2+ oxidation, and Mn3+ disproportionation processes. A detailed study of the pH changes during the cycling process [20,103] showed that the local pH changed from 4.3 to 5.2 because increasing the pH up to 5.6 led to an oxygen evolution reaction (OER), which may be a competitive process at high voltage values. Several strategies could be used to prevent OER, such as abruptly cutting the potential at E = 1.7 V, increasing the acidity of the electrolyte, or using the electrodes with high mass loading. In operando pH measurements in different electrolytes (0.9 mM H2SO4, 0.5 M MnSO4, 2 M ZnSO4, and mixed Zn-Mn electrolytes) showed a strong dependence of the pH on the electrolyte composition. In 0.9 mM H2SO4, the pH increased sharply up to 8.0, driven by MnO2 dissolution during the discharge process, and then returned to the initial value. For 2 M ZnSO4, insignificant pH changes were observed after the first discharge, while for MnSO4 and mixed electrolyte, the pH value fluctuated. In the absence of ZnSO4 in the electrolyte solution, the shape of the CV curve changed dramatically, and only one reduction peak was observed at E = 1.10 V (in 0.9 mM H2SO4) or E = 1.17 V (in 0.5 M MnSO4) vs. Zn/Zn2+ [103]. In summary, the following reaction mechanisms could be proposed by studying the pH behavior: H+ (de)intercalation with a minor influence on the overall charge/discharge capacity, Mn2+ deposition/MnO2 dissolution at the positive electrode surface as the main source of the capacity, ZHS precipitation/dissolution as a consequence of local pH decreases/increases, and OER/HER processes (Figure 13b). Zn2+ intercalation was not discussed in this version of the mechanism [20].
Figure 13. (a) Scheme of phase transitions in Al-doped MnO2 material with Zn2+ ions reversible intercalation [24]. (b) Summary of the chemical processes in the AZIB with a Zn/ZnSO4+MnSO4/MnO2 cell: (1) Mn2+ deposition/MnO2 dissolution, (2) H+ (de-)intercalation, (3) OER/ORR, (4) ZHS precipitation/dissolution, (5) zinc plating/stripping, and (6) HER [103]. (c) Schematic illustration of the reaction mechanism of α-MnO2 in AZIBs with the dissolution/deposition mechanism [107]. (d) Joint non-diffusion controlled Zn2+ intercalation and H+ conversion reaction mechanism in δ-MnO2 [108].
Figure 13. (a) Scheme of phase transitions in Al-doped MnO2 material with Zn2+ ions reversible intercalation [24]. (b) Summary of the chemical processes in the AZIB with a Zn/ZnSO4+MnSO4/MnO2 cell: (1) Mn2+ deposition/MnO2 dissolution, (2) H+ (de-)intercalation, (3) OER/ORR, (4) ZHS precipitation/dissolution, (5) zinc plating/stripping, and (6) HER [103]. (c) Schematic illustration of the reaction mechanism of α-MnO2 in AZIBs with the dissolution/deposition mechanism [107]. (d) Joint non-diffusion controlled Zn2+ intercalation and H+ conversion reaction mechanism in δ-MnO2 [108].
Energies 16 03221 g013
Another mechanism (iii) of electrochemical reaction discussed in the literature is the co-insertion of Zn2+ and H+ due to the small size of H+ ions. A typical scheme of this mechanism is shown in Figure 13d. As was shown by ex situ XRD, two main products formed at the end of the discharge were zinc hydroxide sulfate Zn4(OH)6SO4∙nH2O (ZHS), which dissolves during the subsequent charging process, and MnOOH [25,29,43,48,96,108,109,110]. These discharge products could be associated with the co-insertion of protons and zinc ions into the lattice with partial reduction of the MnO2 host and local increase of the pH near the electrode surface. It has been shown by ex situ XPS that the valence state of manganese in the fully discharged state is close to Mn3+, and the presence of Zn2+ ions extracted during charging with Mn3+-Mn4+ redox transition has also been confirmed [48]. Reversible deposition of ZHS with significant lattice expansion due to Zn2+ insertion has been described for layered MnO2 doped with alkali metal ions [23,36].
Thus, for the first discharge process in pure ZnSO4 solution, only one peak pair is observed from the cyclic voltammograms, while in mixed 2 M ZnSO4/0.5 M MnSO4, two peak pairs were detected. In addition, the reversibility of the electrochemical reaction was better in the mixed electrolyte [111]. Ex situ XRD measurements of the electrodes during the discharge process up to 1.4 V showed no significant changes compared to the initial patterns, while after decreasing the voltage to 1.2 and 1.0 V, new phases were observed and attributed to Zn4(OH)6SO4∙5H2O, α-MnOOH, and Mn2O3 with the presence of zinc in the Mn-containing phase, so that the insertion of Zn2+ and H+ was detected, and H+ was intercalated at lower voltages than that of Zn2+, which contradicts the other reports [66,79,112,113,114], where the first voltage plateau/peak corresponded to H+ insertion and the second to Zn2+ intercalation. H+ intercalation is accompanied by the conversion of MnO2 to MnOOH or Mn2O3 due to the reduction of Mn to MnO2. During the charging process, Zn4(OH)6SO4∙5H2O is dissolved, and ZnMn3O7∙3H2O is formed. A detailed analysis showed that Zn4(OH)6SO4∙5H2O reacted with Mn2+ ions to form the ZnMn3O7∙3H2O phase, while MnOOH transformed into MnO2 during the charging process. Thus, it was shown that three ions (H+, Zn2+, and Mn2+) participate in the electrochemical reaction, and after continuous cycling, many forms of Zn-containing manganese oxides and Zn4(OH)6SO4∙5H2O coexist in the discharged state, and the formation of these phases, except for Zn4(OH)6SO4∙5H2O, is irreversible [111]. Thus, a complete electrochemical mechanism could be presented as a two-step intercalation and conversion mechanism of transformation of MnO2 to MnOOH with a further transition to Mn3O4 and deposition of ZHS [108,109]. Furthermore, from the analysis of cyclic voltammograms obtained in the scan rate range 0.1–1.0 mV s−1, it was found that Zn2+ intercalation is a non-diffusion controlled process, while intercalation of H+ led to structural transformations (detected on the ex situ XRD patterns), which is the reason for the conversion process of MnO2 [108].
The initial MnO2 crystal lattice affects the electrochemical performance of the electrode materials in mildly acidic mixed electrolyte. In particular, β-MnO2 and δ-MnO2 showed longer capacity growth during cycling at 0.1 A∙g−1 than other used polymorphs (α-, γ-, ε-, λ-, and R-MnO2) [115]. The structural evolution of β-MnO2 at different current densities (0.1, 0.3, 0.5, and 1.0 A∙g−1) was studied by ex situ XRD. It was shown that the initial crystal phase changed gradually during 50–100 cycles depending on the applied current: the fastest evolution was observed at 0.1 A∙g−1 and the slowest at 1.0 A∙g−1. According to the appearance of the nanosheets (which could also be associated with layered-type ZHS), the initial product of these phase transitions was birnessite. The final product at the end of the long-term cycling was mixed ZnMn2O4∙Mn3O4 oxide, or (Zn,Mn)Mn2O4, which is an electrochemically inactive phase. These observations were verified for other MnO2 polymorphs by using acetic acid to buffer the electrolyte acidity, and all results confirmed the formation of inactive spinel [116]. In this process, dual Zn2+/H+ co-insertion occurs with accompanying distortion of Mn3+, followed by its dissolution. The main cause of cathode degradation is MnO2 dissolution. An XRD analysis of ramsdellite-type MnO2 showed no metastable phase as the birnessite was formed, and 95% of the initial capacity was retained. Ex situ XRD patterns of R-MnO2 showed the same diffraction peaks after the 1st, 100th, and 300th cycles, indicating that Mn2+ dissolution/deposition had no effect on this type of material [115]. In addition, an extremely low degree of dissolution of ramsdellite was observed. As for other MnO2 polymorphs, the new phase of ZHS was observed by ex situ XRD during the discharge process of R-MnO2.
In several reports, no or weak reversible intercalation of Zn2+ occurred, while irreversible insertion of Zn2+ with structural transformations of MnO2 cathodes was observed [104,105]. Such works are classified as a (iv) type of mechanism, electrolytic deposition/dissolution of MnO2 (Figure 13c). The initial β-MnO2 phase was destroyed during the discharge in 2 M ZnSO4 solution, and ZHS layers (with different amounts of crystallization water) formed on the electrode surface. During the first charge, no peaks of β-MnO2 were observed in the diffraction patterns. After five charge/discharge cycles, three valence states (Mn2+, Mn3+, and Mn4+) were observed in the XPS spectra of manganese in the ratio of 41.7%, 36.5%, and 21.8%, respectively. In addition, the Zn2+ signals disappeared after acid treatment, indicating that the Zn2+ intercalation mechanism was not implemented [104]. The presence of Mn3+ in the resulting compound allowed the authors [104] to conclude that HMnO2 was formed during H+ insertion without any changes in the crystalline structure. The formation of ZHS at E = 1.35 V was demonstrated by the chelation of Zn by a sodium trimetaphosphate additive and the absence of the corresponding peak on the CV curve. Thus, at low number of cycles, no zinc was inserted into the MnO2 lattice. After 100 cycles, 0.26% of Zn was present after acidic treatment of the sample, and the Zn–O–Mn bond was observed in the O 1 s spectra. A detailed ex situ XRD analysis showed that the highly crystalline material became amorphous, and several peaks consistent with the ZnMn2O4 phase appeared, but they formed during the charging of the cell rather than during discharging due to alkalinization of the electrolyte solution after five cycles. It should be noted that ZnMn2O4 is not an electroactive compound here, and its presence hinders the electroactivity of the initial cathode materials. The TEM characterization of the cathode after the first discharge cycle showed that ZHS formed on the cathode surface, and after the first charge, MnO2 with an excess of Zn and the valence state of Mn close to +3.3 (denoted as Zn0.33MnO2) were formed [105]. The formation of this Zn-containing phase is not due to the intercalation of the Zn2+ ions into MnO2 but the consequence of several oxidation processes with the trapping of Zn2+ ions, as was also observed from XRD patterns [97]. The formation of ZHS occurred even in the absence of the cathode material on the substrate, but in the presence of Mn2+, the Mn2+ oxidation potential decreased from +1.9 V to +1.55 V. The studies by ex situ XRD in operando Raman spectroscopy and SEM showed two parallel reactions: a slow decrease of the intensity of the peaks in the Raman spectra of MnO2; and, at E < 1.3 V, a decrease of the peak intensity and the appearance of ZHS. During charging, almost the complete dissolution of ZHS and formation of the Zn-containing phase were observed. Thus, ZHS is an agent that hinders the overall electrochemical reaction. ZHS can also contain a significant amount of the phase of Mn-doped ZHS [117], while this compound is not formed by intercalation, but by conversion, since there is no phase change to the layered one (detected in the XRD data), and the manganese dissolves at low potentials. In addition, the pH has a crucial effect on the mechanism, as pH changes occur simultaneously with the dissolution and precipitation of MnO2 [118]. In addition, the constant pH value throughout the volume of the electrolyte affects the concentration of Zn2+ complexes, which are an intermediate in the formation of ZHS, as evidenced by the significant increase in the molar mass of transferred particles calculated from EQCM measurements at low potentials (1.3–1.0 V) [105,119].
The same observations can be applied to the dissolution/deposition mechanism of MnO2 transformations [106,107,120]. In this case, the deposition of ZHS is associated with almost complete dissolution of the initial cathode material due to intercalation of H+ and subsequent conversion reaction with MnO2. A nanoscale TEM analysis of a Zn//α-MnO2 battery showed that the dissolution of Mn2+ releases 4 OH, leading to the formation of the ZHS coating on the initial cathode material. During the charging process, the deposition of MnO2 and the trapping of Zn2+ ions from the electrolyte solution occurred, resulting in the formation of inactive Zn-containing manganese oxide, as described in [120]. The dissolution of manganese was also observed by the gradual disappearance of MnOOH during the continuous cycling of the electrode with the single crystalline phase, which was indexed as ZHS [48].
For comparison, in the absence of water in the composition of the electrolyte for a nonaqueous rechargeable Zn metal battery in organic electrolytes, the mechanism changes dramatically. Moreover, δ-MnO2 tested in Zn(TFSI)2–acetonitrile electrolyte solution showed a different shape of charge/discharge profiles with an average cell potential of ≈1.37 V [121]. It was shown that the highest specific capacity at low current (123 mAh∙g−1 at 0.04 C) corresponds to approximately 0.2 mol of Zn2+ ions inserted per mole of MnO2. During continuous cycling, the capacity first increased with the increasing electrode/electrolyte interface due to cathode wetting, and then the dissolution of Mn2+ and K+ led to a sharp capacity fade. The reversible insertion and extraction of Zn2+ ions during cycling were demonstrated by ex situ XRD measurements of the electrodes: an additional peak of ZnO and the shift of the (111) peak were observed. The valence state of manganese in pristine MnO2 from XANES data was +3.6, and during cycling, it changed to +3.2 for the discharged and +3.8 for the charged states. Thus, in nonaqueous cells, the mechanism of the electrochemical reaction was described as reversible structural transformations between layered-type δ-MnO2 and spinel-type Zn-containing phase ZnMn2O4, without the influence of protons on the properties and mechanism of cathode operation.
In summary, we can conclude that the insertion of H+ with small amounts of Zn2+ (iii) and the electrolytic dissolution/deposition mechanism of MnO2 cathode (iv) are not contradictory but complementary ways of electrochemical reaction in aqueous media. Many reports confirmed that the trapping of Zn2+ and the formation of electrochemically inactive Zn-containing manganese oxide led to capacity fading, while the formation of complex Zn4(OH)6SO4∙nH2O salt during discharge is the pH-controlled process caused by Mn3+ distortion and H+ intercalation during the discharge process.
The structural and interfacial design of MnO2 cathodes, metal ion doping, and introduction of conducting polymers lead to the implementation of mechanism (iii), the co-(de)intercalation of Zn2+ and H+. It should be noted that the structural transformations and the formation of ZnxMnO2 after the first discharge are confirmed by ex situ XRD studies. Ex situ XRD studies of PEDOT-coated MnO2 showed highly reversible formation of MnOOH and ZnMn2O4 during discharge, consistent with the dual insertion mechanism [95]. At the same time, there is often no clear evidence for the co-intercalation of H+ ions, but only indirect evidence, namely the formation of ZHS and a pH change of the electrolyte.

5. Summary and Outlook

The recent studies have shown that the doping of MnO2 with metal ions and the synthesis of MnO2-conducting polymer composites are the effective ways to improve the electrochemical performance of Zn//MnO2 rechargeable batteries.
In the case of doped MnO2 materials, the stabilization of the host lattice and the expansion of the interlayer spaces facilitate the diffusion of intercalated particles, thus significantly improving the rate performance. The presence of novel chemical bonds (such as Al-O) reduces the structural transformations of MnO2 during cycling and increases cycling stability. Conducting polymers significantly improve the ionic and electronic conductivity of the composite materials, which also results in superior cell performance. The use of conducting polymers is more advantageous due to the absence of toxic or heavy metals and the high number of synthetic routes. On the other hand, doping makes it possible to manipulate the crystal lattice and create materials with defects that significantly improve their own conductivity. Excellent and comparable effects in electrochemical performance have been achieved by using both strategies.
These two strategies can be combined to fabricate the composites of metal-ion-doped MnO2 with the conducting polymer, as shown by the example of Fe/α-MnO2@PPy and PEDOT@Co-MnO2 composites, demonstrating the improved performance [84,92]. In both cases, the synergistic effect of the application of conducting polymer coating and dopants increased the rate and cycling performance due to the improved conductivity of the ternary composites due to the increased interlayer spacing, while the conducting polymer stabilized the surface and suppressed manganese dissolution. Thus, the combination of these two approaches to improve the properties of MnO2-based cathodes is interesting for further investigation in the field of aqueous zinc batteries.
As a detailed analysis of the cited works shows, the conclusions on the mechanism of charge–discharge processes in manganese oxide-based materials are still contradictory and require further in-depth mechanistic analysis.
We believe that this review will shed more light on the novel strategies toward the practical application of Zn//MnO2 batteries, which could replace lead-acid and lithium-ion batteries in stationary energy storage. In addition, the understanding of the electrochemical reaction mechanism could be useful to develop other ways to improve the stability and cycle life of Zn//MnO2 batteries.

Author Contributions

Conceptualization, M.A.K. and V.V.K.; writing—original draft preparation, M.A.K. and F.S.V.; writing—review and editing, E.G.T., V.V.K., and S.N.E.; visualization, E.G.T. and M.A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bai, Y.; Muralidharan, N.; Sun, Y.-K.; Passerini, S.; Stanley Whittingham, M.; Belharouak, I. Energy and Environmental Aspects in Recycling Lithium-Ion Batteries: Concept of Battery Identity Global Passport. Mater. Today 2020, 41, 304–315. [Google Scholar] [CrossRef]
  2. Baum, Z.J.; Bird, R.E.; Yu, X.; Ma, J. Lithium-Ion Battery Recycling–Overview of Techniques and Trends. ACS Energy Lett. 2022, 7, 712–719. [Google Scholar] [CrossRef]
  3. Tang, B.; Shan, L.; Liang, S.; Zhou, J. Issues and Opportunities Facing Aqueous Zinc-Ion Batteries. Energy Environ. Sci. 2019, 12, 3288–3304. [Google Scholar] [CrossRef]
  4. Shi, Y.; Chen, Y.; Shi, L.; Wang, K.; Wang, B.; Li, L.; Ma, Y.; Li, Y.; Sun, Z.; Ali, W.; et al. An Overview and Future Perspectives of Rechargeable Zinc Batteries. Small 2020, 16, 2000730. [Google Scholar] [CrossRef]
  5. Zhang, T.; Tang, Y.; Guo, S.; Cao, X.; Pan, A.; Fang, G.; Zhou, J.; Liang, S. Fundamentals and Perspectives in Developing Zinc-Ion Battery Electrolytes: A Comprehensive Review. Energy Environ. Sci. 2020, 13, 4625–4665. [Google Scholar] [CrossRef]
  6. Zhang, X.; Wang, L.; Fu, H. Recent Advances in Rechargeable Zn-Based Batteries. J. Power Sources 2021, 493, 229677. [Google Scholar] [CrossRef]
  7. Borchers, N.; Clark, S.; Horstmann, B.; Jayasayee, K.; Juel, M.; Stevens, P. Innovative Zinc-Based Batteries. J. Power Sources 2021, 484, 229309. [Google Scholar] [CrossRef]
  8. Xu, C.; Li, B.; Du, H.; Kang, F. Energetic Zinc Ion Chemistry: The Rechargeable Zinc Ion Battery. Angew. Chemie-Int. Ed. 2012, 51, 933–935. [Google Scholar] [CrossRef]
  9. Chen, L.; An, Q.; Mai, L. Recent Advances and Prospects of Cathode Materials for Rechargeable Aqueous Zinc-Ion Batteries. Adv. Mater. Interfaces 2019, 6, 1900387. [Google Scholar] [CrossRef]
  10. Liu, X.; Yi, J.; Wu, K.; Jiang, Y.; Liu, Y.; Zhao, B.; Li, W.; Zhang, J. Rechargeable Zn–MnO2 Batteries: Advances, Challenges and Perspectives. Nanotechnology 2020, 31, 122001. [Google Scholar] [CrossRef] [PubMed]
  11. Bensalah, N.; De Luna, Y. Recent Progress in Layered Manganese and Vanadium Oxide Cathodes for Zn-Ion Batteries. Energy Technol. 2021, 9, 2100011. [Google Scholar] [CrossRef]
  12. Liang, R.; Fu, J.; Deng, Y.P.; Pei, Y.; Zhang, M.; Yu, A.; Chen, Z. Parasitic Electrodeposition in Zn-MnO2 Batteries and Its Suppression for Prolonged Cyclability. Energy Storage Mater. 2021, 36, 478–484. [Google Scholar] [CrossRef]
  13. Liu, Z.; Qin, L.; Lu, B.; Wu, X.; Liang, S.; Zhou, J. Issues and Opportunities Facing Aqueous Mn2+/MnO2-based Batteries. ChemSusChem 2022, 15, e202200348. [Google Scholar] [CrossRef] [PubMed]
  14. Yadav, P.; Kumari, N.; Rai, A.K. A Review on Solutions to Overcome the Structural Transformation of Manganese Dioxide-Based Cathodes for Aqueous Rechargeable Zinc Ion Batteries. J. Power Sources 2023, 555, 232385. [Google Scholar] [CrossRef]
  15. Zhang, Z.; Li, W.; Shen, Y.; Wang, R.; Li, H.; Zhou, M.; Wang, W.; Wang, K.; Jiang, K. Issues and Opportunities of Manganese-Based Materials for Enhanced Zn-Ion Storage Performances. J. Energy Storage 2022, 45, 103729. [Google Scholar] [CrossRef]
  16. Soundharrajan, V.; Sambandam, B.; Kim, S.S.; Islam, S.; Jo, J.; Kim, S.S.; Mathew, V.; Sun, Y.K.; Kim, J. The Dominant Role of Mn2+ Additive on the Electrochemical Reaction in ZnMn2O4 Cathode for Aqueous Zinc-Ion Batteries. Energy Storage Mater. 2020, 28, 407–417. [Google Scholar] [CrossRef]
  17. Qiu, C.; Zhu, X.; Xue, L.; Ni, M.; Zhao, Y.; Liu, B.; Xia, H. The Function of Mn2+ Additive in Aqueous Electrolyte for Zn/δ-MnO2 Battery. Electrochim. Acta 2020, 351, 136445. [Google Scholar] [CrossRef]
  18. Perez-Antolin, D.; Sáez-Bernal, I.; Colina, A.; Ventosa, E. Float-Charging Protocol in Rechargeable Zn–MnO2 Batteries: Unraveling the Key Role of Mn2+ Additives in Preventing Spontaneous pH Changes. Electrochem. Commun. 2022, 138, 107271. [Google Scholar] [CrossRef]
  19. Lv, H.; Song, Y.; Qin, Z.; Zhang, M.; Yang, D.; Pan, Q.; Wang, Z.; Mu, X.; Meng, J.; Sun, X.; et al. Disproportionation Enabling Reversible MnO2/Mn2+ Transformation in a Mild Aqueous Zn-MnO2 Hybrid Battery. Chem. Eng. J. 2022, 430, 133064. [Google Scholar] [CrossRef]
  20. Bischoff, C.F.; Fitz, O.S.; Burns, J.; Bauer, M.; Gentischer, H.; Birke, K.P.; Henning, H.-M.; Biro, D. Revealing the Local pH Value Changes of Acidic Aqueous Zinc Ion Batteries with a Manganese Dioxide Electrode during Cycling. J. Electrochem. Soc. 2020, 167, 020545. [Google Scholar] [CrossRef]
  21. Yong, B.; Ma, D.; Wang, Y.; Mi, H.; He, C.; Zhang, P. Understanding the Design Principles of Advanced Aqueous Zinc-Ion Battery Cathodes: From Transport Kinetics to Structural Engineering, and Future Perspectives. Adv. Energy Mater. 2020, 10, 2002354. [Google Scholar] [CrossRef]
  22. Zhang, B.; Chen, J.; Sun, W.; Shao, Y.; Zhang, L.; Zhao, K. Challenges and Perspectives for Doping Strategy for Manganese-Based Zinc-Ion Battery Cathode. Energies 2022, 15, 4698. [Google Scholar] [CrossRef]
  23. Xie, Q.; Cheng, G.; Xue, T.; Huang, L.; Chen, S.; Sun, Y.; Sun, M.; Wang, H.; Yu, L. Alkali Ions Pre-Intercalation of δ-MnO2 Nanosheets for High-Capacity and Stable Zn-Ion Battery. Mater. Today Energy 2022, 24, 100934. [Google Scholar] [CrossRef]
  24. Chen, C.; Shi, M.; Zhao, Y.; Yang, C.; Zhao, L.; Yan, C. Al-Intercalated MnO2 Cathode with Reversible Phase Transition for Aqueous Zn-Ion Batteries. Chem. Eng. J. 2021, 422, 130375. [Google Scholar] [CrossRef]
  25. Alfaruqi, M.H.; Islam, S.; Mathew, V.; Song, J.; Kim, S.; Tung, D.P.; Jo, J.; Kim, S.; Baboo, J.P.; Xiu, Z.; et al. Ambient Redox Synthesis of Vanadium-Doped Manganese Dioxide Nanoparticles and Their Enhanced Zinc Storage Properties. Appl. Surf. Sci. 2017, 404, 435–442. [Google Scholar] [CrossRef]
  26. Kataoka, F.; Ishida, T.; Nagita, K.; Kumbhar, V.; Yamabuki, K.; Nakayama, M. Cobalt-Doped Layered MnO2 Thin Film Electrochemically Grown on Nitrogen-Doped Carbon Cloth for Aqueous Zinc-Ion Batteries. ACS Appl. Energy Mater. 2020, 3, 4720–4726. [Google Scholar] [CrossRef]
  27. Ko, W.Y.; Lubis, A.L.; Wang, H.Y.; Wu, T.C.; Lin, S.T.; Lin, K.J. Facile Construction of Zn-Doped Mn3O4 −MnO2 Vertical Nanosheets for Aqueous Zinc-Ion Battery Cathodes. ChemElectroChem 2022, 9, e202200750. [Google Scholar] [CrossRef]
  28. Li, X.; Qu, J.; Xu, J.; Zhang, S.; Wang, X.; Wang, X.; Dai, S. K-Preintercalated MnO2 Nanosheets as Cathode for High-Performance Zn-Ion Batteries. J. Electroanal. Chem. 2021, 895, 115529. [Google Scholar] [CrossRef]
  29. Lin, M.; Shao, F.; Tang, Y.; Lin, H.; Xu, Y.; Jiao, Y.; Chen, J. Layered Co Doped MnO2 with Abundant Oxygen Defects to Boost Aqueous Zinc-Ion Storage. J. Colloid Interface Sci. 2022, 611, 662–669. [Google Scholar] [CrossRef]
  30. Ni, Z.; Liang, X.; Zhao, L.; Zhao, H.; Ge, B.; Li, W. Tin Doping Manganese Dioxide Cathode Materials with the Improved Stability for Aqueous Zinc-Ion Batteries. Mater. Chem. Phys. 2022, 287, 126238. [Google Scholar] [CrossRef]
  31. Chen, Q.; Jin, J.; Kou, Z.; Liao, C.; Liu, Z.; Zhou, L.; Wang, J.; Mai, L. Zn2+ Pre-Intercalation Stabilizes the Tunnel Structure of MnO2 Nanowires and Enables Zinc-Ion Hybrid Supercapacitor of Battery-Level Energy Density. Small 2020, 16, 2000091. [Google Scholar] [CrossRef] [PubMed]
  32. Ni, Z.; Liang, X.; Zhao, L.; Zhao, H.; Ge, B.; Li, W. MnO2 Cathode Materials with High Reversible Stability through Li Intercalating for Aqueous Zinc Ion Battery. Solid State Ionics 2022, 386, 116049. [Google Scholar] [CrossRef]
  33. Wu, Y.; Wang, M.; Tao, Y.; Zhang, K.; Cai, M.; Ding, Y.; Liu, X.; Hayat, T.; Alsaedi, A.; Dai, S. Electrochemically Derived Graphene-Like Carbon Film as a Superb Substrate for High-Performance Aqueous Zn-Ion Batteries. Adv. Funct. Mater. 2020, 30, 1907120. [Google Scholar] [CrossRef]
  34. Lin, M.X.; Shao, F.; Weng, S.; Xiong, S.; Liu, S.; Jiang, S.; Xu, Y.; Jiao, Y.; Chen, J. Boosted Charge Transfer in Oxygen Vacancy-Rich K+ Birnessite MnO2 for Water Oxidation and Zinc-Ion Batteries. Electrochim. Acta 2021, 378, 138147. [Google Scholar] [CrossRef]
  35. Lv, W.; Meng, J.; Li, Y.; Yang, W.; Tian, Y.; Lyu, X.; Duan, C.; Ma, X.; Wu, Y. Inexpensive and Eco-Friendly Nanostructured Birnessite-Type δ-MnO2: A Design Strategy from Oxygen Defect Engineering and K+ Pre-Intercalation. Nano Energy 2022, 98, 107274. [Google Scholar] [CrossRef]
  36. Wang, D.; Zhang, S.; Li, C.; Chen, X.; Wang, W.; Han, Y.; Lin, H.; Shi, Z.; Feng, S. Engineering the Interplanar Spacing of K-Birnessite for Ultra-Long Cycle Zn-Ion Battery through “Hydrothermal Potassium Insertion” Strategy. Chem. Eng. J. 2022, 435, 134754. [Google Scholar] [CrossRef]
  37. Liu, G.; Huang, H.; Bi, R.; Xiao, X.; Ma, T.; Zhang, L. K+ Pre-Intercalated Manganese Dioxide with Enhanced Zn2+ Diffusion for High Rate and Durable Aqueous Zinc-Ion Batteries. J. Mater. Chem. A 2019, 7, 20806–20812. [Google Scholar] [CrossRef]
  38. Tan, J.; Feng, T.; Hu, S.; Liang, Y.; Zhang, S.; Xu, Z.; Zhou, H.; Wu, M. In Situ Synthesis of a Self-Supported MnO2-Based Cathode for High-Performance Zinc-Ion Batteries by K+ Pre-Intercalation. Appl. Surf. Sci. 2022, 604, 154578. [Google Scholar] [CrossRef]
  39. Xu, T.H.; Liou, S.; Hou, F.L.; Li, Y.Y. Potassium-Doped Hydrated Manganese Dioxide Nanowires-Carbon Nanotubes on Graphene for High-Performance Rechargeable Zinc-Ion Batteries. J. Alloys Compd. 2022, 913, 165278. [Google Scholar] [CrossRef]
  40. Sun, T.; Nian, Q.; Zheng, S.; Shi, J.; Tao, Z. Layered Ca0.28MnO2·0.5H2O as a High Performance Cathode for Aqueous Zinc-Ion Battery. Small 2020, 16, 2000597. [Google Scholar] [CrossRef]
  41. Wang, J.; Wang, J.-G.; Liu, H.; Wei, C.; Kang, F. Zinc Ion Stabilized MnO2 Nanospheres for High Capacity and Long Lifespan Aqueous Zinc-Ion Batteries. J. Mater. Chem. A 2019, 7, 13727–13735. [Google Scholar] [CrossRef]
  42. Shi, M.; Zhu, H.; Chen, C.; Jiang, J.; Zhao, L.; Yan, C. Synergistically Coupling of Graphene Quantum Dots with Zn-Intercalated MnO2 Cathode for High-Performance Aqueous Zn-Ion Batteries. Int. J. Miner. Metall. Mater. 2023, 30, 25–32. [Google Scholar] [CrossRef]
  43. Li, Y.; Li, X.; Duan, H.; Xie, S.; Dai, R.; Rong, J.; Kang, F.; Dong, L. Aerogel-Structured MnO2 Cathode Assembled by Defect-Rich Ultrathin Nanosheets for Zinc-Ion Batteries. Chem. Eng. J. 2022, 441, 136008. [Google Scholar] [CrossRef]
  44. Li, H.; Huang, Z.; Chen, B.; Jiang, Y.; Li, C.; Xiao, W.; Yan, X. A High-Performance MnO2 Cathode Doped with Group VIII Metal for Aqueous Zn-Ion Batteries: In-Situ X-Ray Diffraction Study on Zn2+ Storage Mechanism. J. Power Sources 2022, 527, 231198. [Google Scholar] [CrossRef]
  45. Zhong, Y.; Xu, X.; Veder, J.-P.; Shao, Z. Self-Recovery Chemistry and Cobalt-Catalyzed Electrochemical Deposition of Cathode for Boosting Performance of Aqueous Zinc-Ion Batteries. iScience 2020, 23, 100943. [Google Scholar] [CrossRef] [Green Version]
  46. Yang, S.; Zhang, L.; Luo, M.; Cui, Y.; Wang, J.; Zhao, D.; Yang, C.; Wang, X.; Cao, B. Synergistic Combination of a Co-Doped σ-MnO2 Cathode with an Electrolyte Additive for a High-Performance Aqueous Zinc-Ion Battery. ChemPhysMater 2022, 2, 77–82. [Google Scholar] [CrossRef]
  47. Zhao, Q.; Song, A.; Zhao, W.; Qin, R.; Ding, S.; Chen, X.; Song, Y.; Yang, L.; Lin, H.; Li, S.; et al. Boosting the Energy Density of Aqueous Batteries via Facile Grotthuss Proton Transport. Angew. Chem. 2021, 133, 4215–4220. [Google Scholar] [CrossRef]
  48. Zhang, R.; Liang, P.; Yang, H.; Min, H.; Niu, M.; Jin, S.; Jiang, Y.; Pan, Z.; Yan, J.; Shen, X.; et al. Manipulating Intercalation-Extraction Mechanisms in Structurally Modulated δ-MnO2 Nanowires for High-Performance Aqueous Zinc-Ion Batteries. Chem. Eng. J. 2022, 433, 133687. [Google Scholar] [CrossRef]
  49. Long, F.; Xiang, Y.; Yang, S.; Li, Y.; Du, H.; Liu, Y.; Wu, X.; Wu, X. Layered Manganese Dioxide Nanoflowers with Cu2+ and Bi3+ Intercalation as High-Performance Cathode for Aqueous Zinc-Ion Battery. J. Colloid Interface Sci. 2022, 616, 101–109. [Google Scholar] [CrossRef]
  50. Liao, Y.; Yang, C.; Xu, Q.; Zhao, W.; Zhao, J.; Wang, K.; Chen, H.C. Ag-Doping Effect on MnO2 Cathodes for Flexible Quasi-Solid-State Zinc-Ion Batteries. Batteries 2022, 8, 267. [Google Scholar] [CrossRef]
  51. Zheng, Z.; Yang, G.; Yao, J.; Li, J.; Zheng, J.; Wu, Z.; Gan, Y.; Wang, C.; Lv, L.; Wan, H.; et al. High-Valence Molybdenum Promoted Proton Migration and Inhibited Dissolution for Long-Life Aqueous Zn-MnO2 Batteries. Appl. Surf. Sci. 2022, 592, 153335. [Google Scholar] [CrossRef]
  52. Wang, Z.; Han, K.; Wan, Q.; Fang, Y.; Qu, X.; Li, P. Mo-Pre-Intercalated MnO2 Cathode with Highly Stable Layered Structure and Expanded Interlayer Spacing for Aqueous Zn-Ion Batteries. ACS Appl. Mater. Interfaces 2023, 15, 859–869. [Google Scholar] [CrossRef]
  53. Zhang, H.; Liu, Q.; Wang, J.; Chen, K.; Xue, D.; Liu, J.; Lu, X. Boosting the Zn-Ion Storage Capability of Birnessite Manganese Oxide Nanoflorets by La3+ Intercalation. J. Mater. Chem. A 2019, 7, 22079–22083. [Google Scholar] [CrossRef]
  54. Wang, J.; Sun, X.; Zhao, H.; Xu, L.; Xia, J.; Luo, M.; Yang, Y.; Du, Y. Superior-Performance Aqueous Zinc Ion Battery Based on Structural Transformation of MnO2 by Rare Earth Doping. J. Phys. Chem. C 2019, 123, 22735–22741. [Google Scholar] [CrossRef]
  55. Song, Y.; Li, J.; Qiao, R.; Dai, X.; Jing, W.; Song, J.; Chen, Y.; Guo, S.; Sun, J.; Tan, Q.; et al. Binder-Free Flexible Zinc-Ion Batteries: One-Step Potentiostatic Electrodeposition Strategy Derived Ce Doped-MnO2 Cathode. Chem. Eng. J. 2022, 431, 133387. [Google Scholar] [CrossRef]
  56. Xu, J.; Hu, X.; Alam, M.A.; Muhammad, G.; Lv, Y.; Wang, M.; Zhu, C.; Xiong, W. Al-Doped α-MnO2 coated by Lignin for High-Performance Rechargeable Aqueous Zinc-Ion Batteries. RSC Adv. 2021, 11, 35280–35286. [Google Scholar] [CrossRef]
  57. Zhou, S.; Wu, X.; Du, H.; He, Z.; Wu, X.; Wu, X. Dual Metal Ions and Water Molecular Pre-Intercalated δ-MnO2 Spherical Microflowers for Aqueous Zinc Ion Batteries. J. Colloid Interface Sci. 2022, 623, 456–466. [Google Scholar] [CrossRef] [PubMed]
  58. Li, Y.; Zhang, H.; Tian, T.; Weng, Q.; Zan, L.; Zhao, S.; Liu, T.; Tang, Z.; Tang, H. Boosting High-Rate Zn-Ion Storage Capability of α-MnO2 through Tri-Ion Co-Intercalation. J. Alloys Compd. 2023, 939, 168813. [Google Scholar] [CrossRef]
  59. Ma, Y.; Xu, M.; Liu, R.; Xiao, H.; Liu, Y.; Wang, X.; Huang, Y.; Yuan, G. Molecular Tailoring of MnO2 by Bismuth Doping to Achieve Aqueous Zinc-Ion Battery with Capacitor-Level Durability. Energy Storage Mater. 2022, 48, 212–222. [Google Scholar] [CrossRef]
  60. Wang, D.; Wang, L.; Liang, G.; Li, H.; Liu, Z.; Tang, Z.; Liang, J.; Zhi, C. A Superior δ-MnO2 Cathode and a Self-Healing Zn-δ-MnO2 Battery. ACS Nano 2019, 13, 10643–10652. [Google Scholar] [CrossRef]
  61. Kim, J.J.H.; Lee, J.; You, J.; Park, M.S.; Al Hossain, M.S.; Yamauchi, Y.; Kim, J.J.H. Conductive Polymers for Next-Generation Energy Storage Systems: Recent Progress and New Functions. Mater. Horizons 2016, 3, 517–535. [Google Scholar] [CrossRef]
  62. Balint, R.; Cassidy, N.J.; Cartmell, S.H. Conductive Polymers: Towards a Smart Biomaterial for Tissue Engineering. Acta Biomater. 2014, 10, 2341–2353. [Google Scholar] [CrossRef]
  63. Zhang, F.; Wang, C.; Pan, J.; Tian, F.; Zeng, S.; Yang, J.; Qian, Y. Polypyrrole-Controlled Plating/Stripping for Advanced Zinc Metal Anodes. Mater. Today Energy 2020, 17, 100443. [Google Scholar] [CrossRef]
  64. Ciric-Marjanovic, G. Recent Advances in Polyaniline Research: Polymerization Mechanisms, Structural Aspects, Properties and Applications. Synth. Met. 2013, 177, 1–47. [Google Scholar] [CrossRef]
  65. Huang, J.; Tu, J.; Lv, Y.; Liu, Y.; Huang, H.; Li, L.; Yao, J. Achieving Mesoporous MnO2@polyaniline Nanohybrids via a Gas/Liquid Interfacial Reaction between Aniline and KMnO4 Aqueous Solution towards Zn-MnO2 Battery. Synth. Met. 2020, 266, 116438. [Google Scholar] [CrossRef]
  66. Huang, J.; Wang, Z.; Hou, M.; Dong, X.; Liu, Y.; Wang, Y.; Xia, Y. Polyaniline-Intercalated Manganese Dioxide Nanolayers as a High-Performance Cathode Material for an Aqueous Zinc-Ion Battery. Nat. Commun. 2018, 9, 2906. [Google Scholar] [CrossRef] [Green Version]
  67. Mao, J.; Wu, F.-F.; Shi, W.-H.; Liu, W.-X.; Xu, X.-L.; Cai, G.-F.; Li, Y.-W.; Cao, X.-H. Preparation of Polyaniline-Coated Composite Aerogel of MnO2 and Reduced Graphene Oxide for High-Performance Zinc-Ion Battery. Chinese J. Polym. Sci. 2020, 38, 514–521. [Google Scholar] [CrossRef]
  68. Ruan, P.; Xu, X.; Gao, X.X.; Feng, J.; Yu, L.; Cai, Y.; Gao, X.X.; Shi, W.; Wu, F.; Liu, W.; et al. Achieving Long-Cycle-Life Zn-Ion Batteries through Interfacial Engineering of MnO2-Polyaniline Hybrid Networks. Sustain. Mater. Technol. 2021, 28, e00254. [Google Scholar] [CrossRef]
  69. Li, N.; Hou, Z.; Liang, S.; Cao, Y.; Liu, H.; Hua, W.; Wei, C.; Kang, F.; Wang, J.G. Highly Flexible MnO2@polyaniline Core-Shell Nanowire Film toward Substantially Expedited Zinc Energy Storage. Chem. Eng. J. 2023, 452, 139408. [Google Scholar] [CrossRef]
  70. Li, Y.; Liu, Y.; Chen, J.; Zheng, Q.; Huo, Y.; Xie, F.; Lin, D. Polyaniline Intercalation Induced Great Enhancement of Electrochemical Properties in Ammonium Vanadate Nanosheets as an Advanced Cathode for High-Performance Aqueous Zinc-Ion Batteries. Chem. Eng. J. 2022, 448, 137681. [Google Scholar] [CrossRef]
  71. Zhang, Y.; Xu, L.; Jiang, H.; Liu, Y.; Meng, C. Polyaniline-Expanded the Interlayer Spacing of Hydrated Vanadium Pentoxide by the Interface-Intercalation for Aqueous Rechargeable Zn-Ion Batteries. J. Colloid Interface Sci. 2021, 603, 641–650. [Google Scholar] [CrossRef] [PubMed]
  72. Li, R.; Xing, F.; Li, T.; Zhang, H.; Yan, J.; Zheng, Q.; Li, X. Intercalated Polyaniline in V2O5 as a Unique Vanadium Oxide Bronze Cathode for Highly Stable Aqueous Zinc Ion Battery. Energy Storage Mater. 2021, 38, 590–598. [Google Scholar] [CrossRef]
  73. Hao, L.; Yu, D. Progress of Conductive Polypyrrole Nanocomposites. Synth. Met. 2022, 290, 117138. [Google Scholar] [CrossRef]
  74. Kausar, A. Overview on Conducting Polymer in Energy Storage and Energy Conversion System. J. Macromol. Sci. Part A 2017, 54, 640–653. [Google Scholar] [CrossRef]
  75. Chavan, U.D.; Prajith, P.; Kandasubramanian, B. Polypyrrole Based Cathode Material for Battery Application. Chem. Eng. J. Adv. 2022, 12, 100416. [Google Scholar] [CrossRef]
  76. Ruan, P.; Xu, X.; Feng, J.; Yu, L.; Gao, X.; Shi, W.; Wu, F.; Liu, W.; Zang, X.; Ma, F.; et al. Boosting Zinc Storage Performance via Conductive Materials. Mater. Res. Bull. 2021, 133, 111077. [Google Scholar] [CrossRef]
  77. Morávková, Z.; Taboubi, O.; Minisy, I.M.; Bober, P. The Evolution of the Molecular Structure of Polypyrrole during Chemical Polymerization. Synth. Met. 2021, 271, 116608. [Google Scholar] [CrossRef]
  78. Huang, J.; Tang, X.; Liu, K.; Fang, G.; He, Z.; Li, Z. Interfacial Chemical Binding and Improved Kinetics Assisting Stable Aqueous Zn–MnO2 Batteries. Mater. Today Energy 2020, 17, 100475. [Google Scholar] [CrossRef]
  79. Liao, X.; Pan, C.; Pan, Y.; Yin, C. Synthesis of Three-Dimensional β-MnO2/PPy Composite for High-Performance Cathode in Zinc-Ion Batteries. J. Alloys Compd. 2021, 888, 161619. [Google Scholar] [CrossRef]
  80. Shen, X.; Wang, X.; Yu, N.; Yang, W.; Zhou, Y.; Shi, Y.; Wang, Y.; Dong, L.; Di, J.; Li, Q. A Polypyrrole-Coated MnO2/Carbon Nanotube Film Cathode for Rechargeable Aqueous Zn-Ion Batteries. Acta Phys. Chim. Sin. 2020, 38, 2006059. [Google Scholar] [CrossRef]
  81. Zhang, Y.; Xu, G.; Liu, X.; Wei, X.; Cao, J.; Yang, L. Scalable In Situ Reactive Assembly of Polypyrrole-Coated MnO2 Nanowire and Carbon Nanotube Composite as Freestanding Cathodes for High Performance Aqueous Zn-Ion Batteries. ChemElectroChem 2020, 7, 2762–2770. [Google Scholar] [CrossRef]
  82. Niu, T.; Li, J.; Qi, Y.; Huang, X.; Ren, Y. Preparation and Electrochemical Properties of α-MnO2/RGO-PPy Composite as Cathode Material for Zinc-Ion Battery. J. Mater. Sci. 2021, 56, 16582–16590. [Google Scholar] [CrossRef]
  83. Huang, A.; Zhou, W.; Wang, A.; Chen, M.; Chen, J.; Tian, Q.; Xu, J. Self-Initiated Coating of Polypyrrole on MnO2/Mn2O3 Nanocomposite for High-Performance Aqueous Zinc-Ion Batteries. Appl. Surf. Sci. 2021, 545, 149041. [Google Scholar] [CrossRef]
  84. Xu, J.W.; Gao, Q.L.; Xia, Y.M.; Lin, X.S.; Liu, W.L.; Ren, M.M.; Kong, F.G.; Wang, S.J.; Lin, C. High-Performance Reversible Aqueous Zinc-Ion Battery Based on Iron-Doped Alpha-Manganese Dioxide Coated by Polypyrrole. J. Colloid Interface Sci. 2021, 598, 419–429. [Google Scholar] [CrossRef]
  85. Wang, Z.; Ruan, Z.; Liu, Z.; Wang, Y.; Tang, Z.; Li, H.; Zhu, M.; Hung, T.F.; Liu, J.; Shi, Z.; et al. A Flexible Rechargeable Zinc-Ion Wire-Shaped Battery with Shape Memory Function. J. Mater. Chem. A 2018, 6, 8549–8557. [Google Scholar] [CrossRef]
  86. Sun, K.; Zhang, S.; Li, P.; Xia, Y.; Zhang, X.; Du, D.; Isikgor, F.H.; Ouyang, J. Review on Application of PEDOTs and PEDOT:PSS in Energy Conversion and Storage Devices. J. Mater. Sci. Mater. Electron. 2015, 26, 4438–4462. [Google Scholar] [CrossRef]
  87. Chen, H.W.; Li, C. PEDOT: Fundamentals and Its Nanocomposites for Energy Storage. Chin. J. Polym. Sci. (Engl. Ed.) 2020, 38, 435–448. [Google Scholar] [CrossRef]
  88. Heywang, G.; Jonas, F. Poly(Alkylenedioxythiophene)s—New, Very Stable Conducting Polymers. Adv. Mater. 1992, 4, 116–118. [Google Scholar] [CrossRef]
  89. Petsagkourakis, I.; Kim, N.; Tybrandt, K.; Zozoulenko, I.; Crispin, X. Poly(3,4-ethylenedioxythiophene): Chemical Synthesis, Transport Properties, and Thermoelectric Devices. Adv. Electron. Mater. 2019, 5, 1800918. [Google Scholar] [CrossRef]
  90. Zeng, Y.; Zhang, X.; Meng, Y.; Yu, M.; Yi, J.; Wu, Y.; Lu, X.; Tong, Y. Achieving Ultrahigh Energy Density and Long Durability in a Flexible Rechargeable Quasi-Solid-State Zn-MnO2 Battery. Adv. Mater. 2017, 29, 1700274. [Google Scholar] [CrossRef]
  91. Zhang, X.; Wu, S.; Deng, S.; Wu, W.; Zeng, Y.; Xia, X.; Pan, G.; Tong, Y.; Lu, X. 3D CNTs Networks Enable MnO2 Cathodes with High Capacity and Superior Rate Capability for Flexible Rechargeable Zn–MnO2 Batteries. Small Methods 2019, 3, 1900525. [Google Scholar] [CrossRef]
  92. Shi, X.; Liu, X.; Wang, E.; Cao, X.; Yu, Y.; Cheng, X.; Lu, X. Boosting the Zn Ion Storage Ability of Amorphous MnO2 via Surface Engineering and Valence Modulation. Carbon Neutraliz. 2023, 2, 28–36. [Google Scholar] [CrossRef]
  93. Fu, N.; Zhao, Q.; Xu, Y.; Wang, H.; Hu, J.; Wu, Y.; Yang, L.; Wu, X.; Zeng, X. Exploiting the Synergistic Effect of Multiphase MnO2 Stabilized by an Integrated Conducting Network for Aqueous Zinc-Ion Batteries. Mater. Chem. Front. 2022, 6, 1956–1963. [Google Scholar] [CrossRef]
  94. Zhang, H.; Wang, J.; Liu, Q.; He, W.; Lai, Z.; Zhang, X.; Yu, M.; Tong, Y.; Lu, X. Extracting Oxygen Anions from ZnMn2O4: Robust Cathode for Flexible All-Solid-State Zn-Ion Batteries. Energy Storage Mater. 2019, 21, 154–161. [Google Scholar] [CrossRef]
  95. Wang, L.; Wang, X.; Song, B.; Wang, Z.; Zhang, L.; Lu, Q. Facile in Situ Synthesis of PEDOT Conductor Interface at the Surface of MnO2 Cathodes for Enhanced Aqueous Zinc-Ion Batteries. Surfaces and Interfaces 2022, 33, 102222. [Google Scholar] [CrossRef]
  96. Chen, H.; Ma, W.; Guo, J.; Xiong, J.; Hou, F.; Si, W.; Sang, Z.; Yang, D. PEDOT-Intercalated MnO2 Layers as a High-Performance Cathode Material for Aqueous Zn-Ion Batteries. J. Alloys Compd. 2023, 932, 167688. [Google Scholar] [CrossRef]
  97. Kamenskii, M.A.; Volkov, F.S.; Eliseeva, S.N.; Holze, R.; Kondratiev, V.V. Comparative Study of PEDOT- and PEDOT:PSS Modified δ-MnO2 Cathodes for Aqueous Zinc Batteries with Enhanced Properties. J. Electrochem. Soc. 2023, 170, 010505. [Google Scholar] [CrossRef]
  98. Chen, J.; Liang, J.; Zhou, Y.; Sha, Z.; Lim, S.; Huang, F.; Han, Z.; Brown, S.A.; Cao, L.; Wang, D.; et al. A Vertical Graphene Enhanced Zn–MnO2 Flexible Battery towards Wearable Electronic Devices. J. Mater. Chem. A 2021, 9, 575–584. [Google Scholar] [CrossRef]
  99. Kamenskii, M.A.; Eliseeva, S.N.; Kondratiev, V.V. The Electrochemical Performance of δ-MnO2 Cathode Material for Aqueous Zinc-Ion Batteries: The Role of Current Collector. ECS Trans. 2021, 105, 135–142. [Google Scholar] [CrossRef]
  100. Yamamoto, T.; Shoji, T. Rechargeable Zn∣ZnSO4∣MnO2-Type Cells. Inorg. Chim. Acta 1986, 117, L27–L28. [Google Scholar] [CrossRef]
  101. Xue, T.; Fan, H.J. From Aqueous Zn-Ion Battery to Zn-MnO2 Flow Battery: A Brief Story. J. Energy Chem. 2021, 54, 194–201. [Google Scholar] [CrossRef]
  102. Zhao, Y.; Zhu, Y.; Zhang, X. Challenges and Perspectives for Manganese-Based Oxides for Advanced Aqueous Zinc-Ion Batteries. InfoMat 2020, 2, 237–260. [Google Scholar] [CrossRef]
  103. Fitz, O.; Bischoff, C.; Bauer, M.; Gentischer, H.; Birke, K.P.; Henning, H.M.; Biro, D. Electrolyte Study with in Operando pH Tracking Providing Insight into the Reaction Mechanism of Aqueous Acidic Zn//MnO2 Batteries. ChemElectroChem 2021, 8, 3553–3566. [Google Scholar] [CrossRef]
  104. Li, L.; Hoang, T.K.A.; Zhi, J.; Han, M.; Li, S.; Chen, P. Functioning Mechanism of the Secondary Aqueous Zn-β-MnO2 Battery. ACS Appl. Mater. Interfaces 2020, 12, 12834–12846. [Google Scholar] [CrossRef]
  105. Aguilar, I.; Lemaire, P.; Ayouni, N.; Bendadesse, E.; Morozov, A.V.; Sel, O.; Balland, V.; Limoges, B.; Abakumov, A.M.; Raymundo-Piñero, E.; et al. Identifying Interfacial Mechanisms Limitations within Aqueous Zn-MnO2 Batteries and Means to Cure Them with Additives. Energy Storage Mater. 2022, 53, 238–253. [Google Scholar] [CrossRef]
  106. Wu, Y.; Zhi, J.; Han, M.; Liu, Z.; Shi, Q.; Liu, Y.; Chen, P. Regulating Proton Distribution by Ion Exchange Resin to Achieve Long Lifespan Aqueous Zn-MnO2 Battery. Energy Storage Mater. 2022, 51, 599–609. [Google Scholar] [CrossRef]
  107. Guo, X.; Zhou, J.; Bai, C.; Li, X.; Fang, G.; Liang, S. Zn/MnO2 Battery Chemistry with Dissolution-Deposition Mechanism. Mater. Today Energy 2020, 16, 100396. [Google Scholar] [CrossRef]
  108. Jin, Y.; Zou, L.; Liu, L.; Engelhard, M.H.; Patel, R.L.; Nie, Z.; Han, K.S.; Shao, Y.; Wang, C.; Zhu, J.; et al. Joint Charge Storage for High-Rate Aqueous Zinc–Manganese Dioxide Batteries. Adv. Mater. 2019, 31, 1900567. [Google Scholar] [CrossRef] [PubMed]
  109. Li, Y.; Wang, S.; Salvador, J.R.; Wu, J.; Liu, B.; Yang, W.; Yang, J.; Zhang, W.; Liu, J.; Yang, J. Reaction Mechanisms for Long-Life Rechargeable Zn/MnO2 Batteries. Chem. Mater. 2019, 31, 2036–2047. [Google Scholar] [CrossRef] [Green Version]
  110. Zhou, T.; Zhu, L.; Xie, L.; Han, Q.; Yang, X.; Chen, L.; Wang, G.; Cao, X. Cathode Materials for Aqueous Zinc-Ion Batteries: A Mini Review. J. Colloid Interface Sci. 2022, 605, 828–850. [Google Scholar] [CrossRef] [PubMed]
  111. Huang, Y.; Mou, J.; Liu, W.; Wang, X.; Dong, L.; Kang, F.; Xu, C. Novel Insights into Energy Storage Mechanism of Aqueous Rechargeable Zn/MnO2 Batteries with Participation of Mn2+. Nano-Micro Lett. 2019, 11, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Guo, C.; Liu, H.; Li, J.; Hou, Z.; Liang, J.; Zhou, J.; Zhu, Y.; Qian, Y. Ultrathin δ-MnO2 Nanosheets as Cathode for Aqueous Rechargeable Zinc Ion Battery. Electrochim. Acta 2019, 304, 370–377. [Google Scholar] [CrossRef]
  113. Liu, D.-S.; Mai, Y.; Chen, S.; Liu, S.; Ang, E.H.; Ye, M.; Yang, Y.; Zhang, Y.; Geng, H.; Li, C.C. A 1D–3D Interconnected δ-MnO2 Nanowires Network as High-Performance and High Energy Efficiency Cathode Material for Aqueous Zinc-Ion Batteries. Electrochim. Acta 2021, 370, 137740. [Google Scholar] [CrossRef]
  114. Zhou, W.; Wang, A.; Huang, A.; Chen, M.; Tian, Q.; Chen, J.; Xu, X. Hybridizing δ-Type MnO2 With Lignin-Derived Porous Carbon as a Stable Cathode Material for Aqueous Zn–MnO2 Batteries. Front. Energy Res. 2020, 8, 182. [Google Scholar] [CrossRef]
  115. Liao, Y.; Chen, H.C.; Yang, C.; Liu, R.; Peng, Z.; Cao, H.; Wang, K. Unveiling Performance Evolution Mechanisms of MnO2 Polymorphs for Durable Aqueous Zinc-Ion Batteries. Energy Storage Mater. 2022, 44, 508–516. [Google Scholar] [CrossRef]
  116. Siamionau, U.; Aniskevich, Y.; Mazanik, A.; Kokits, O.; Ragoisha, G.; Jo, J.H.; Myung, S.T.; Streltsov, E. Rechargeable Zinc-Ion Batteries with Manganese Dioxide Cathode: How Critical Is Choice of Manganese Dioxide Polymorphs in Aqueous Solutions? J. Power Sources 2022, 523, 231023. [Google Scholar] [CrossRef]
  117. Stoševski, I.; Bonakdarpour, A.; Fang, B.; Lo, P.; Wilkinson, D.P. Formation of MnxZny(OH)ZSO4⋅5H2O–Not Intercalation of Zn–Is the Basis of the Neutral MnO2/Zn Battery First Discharge Reaction. Electrochim. Acta 2021, 390, 138852. [Google Scholar] [CrossRef]
  118. Liu, Z.; Yang, Y.; Liang, S.; Lu, B.; Zhou, J. pH-Buffer Contained Electrolyte for Self-Adjusted Cathode-Free Zn–MnO2 Batteries with Coexistence of Dual Mechanisms. Small Struct. 2021, 2, 2100119. [Google Scholar] [CrossRef]
  119. Efremova, A.O.; Volkov, A.I.; Tolstopyatova, E.G.; Kondratiev, V.V. EQCM Study of Intercalation Processes into Electrodeposited MnO2 Electrode in Aqueous Zinc-Ion Battery Electrolyte. J. Alloys Compd. 2022, 892, 162142. [Google Scholar] [CrossRef]
  120. Kim, S.J.; Wu, D.; Sadique, N.; Quilty, C.D.; Wu, L.; Marschilok, A.C.; Takeuchi, K.J.; Takeuchi, E.S.; Zhu, Y. Unraveling the Dissolution-Mediated Reaction Mechanism of α-MnO2 Cathodes for Aqueous Zn-Ion Batteries. Small 2020, 16, 2005406. [Google Scholar] [CrossRef]
  121. Han, S.D.; Kim, S.; Li, D.; Petkov, V.; Yoo, H.D.; Phillips, P.J.; Wang, H.; Kim, J.J.; More, K.L.; Key, B.; et al. Mechanism of Zn Insertion into Nanostructured δ-MnO2: A Nonaqueous Rechargeable Zn Metal Battery. Chem. Mater. 2017, 29, 4874–4884. [Google Scholar] [CrossRef]
Figure 1. (a) Thermodynamic operating window of the zinc-ion battery with respect to the pH value. (b) Zoomed area and pH-E plot of the cycling path of the battery cell with the 2 M ZnSO4/0.1 M MnSO4 electrolyte (positive orientation, shaded area should be avoided for cycling) [20].
Figure 1. (a) Thermodynamic operating window of the zinc-ion battery with respect to the pH value. (b) Zoomed area and pH-E plot of the cycling path of the battery cell with the 2 M ZnSO4/0.1 M MnSO4 electrolyte (positive orientation, shaded area should be avoided for cycling) [20].
Energies 16 03221 g001
Figure 2. (a) XRD patterns and (b) cyclic performance at current density 0.3 A∙g–1 of X-δ-MnO2 (X = Li, Na, K) [23]. (c) Cycling performance of MnO2 and Li-MnO2 electrodes at a current density of 0.1 A g–1 [32]. (d) EIS spectra of MnO2 and Li-MnO2 at open-circuit voltage [32]. (e) Rate performance of Li-MnO2 electrode at current densities from 0.1 to 2 A g–1 [32]. (f) Long-term cycling performance of MnO2 and Li-MnO2 electrodes at a high current density of 1 A g–1 [32].
Figure 2. (a) XRD patterns and (b) cyclic performance at current density 0.3 A∙g–1 of X-δ-MnO2 (X = Li, Na, K) [23]. (c) Cycling performance of MnO2 and Li-MnO2 electrodes at a current density of 0.1 A g–1 [32]. (d) EIS spectra of MnO2 and Li-MnO2 at open-circuit voltage [32]. (e) Rate performance of Li-MnO2 electrode at current densities from 0.1 to 2 A g–1 [32]. (f) Long-term cycling performance of MnO2 and Li-MnO2 electrodes at a high current density of 1 A g–1 [32].
Energies 16 03221 g002
Figure 3. (a) The synthesis scheme of the GCF from the raw graphite paper and the photographs of the samples through the synthesis route. (b) The surface SEM of the Na:MnO2/GCF electrode. (c) XRD patterns of the Na:MnO2/GCF electrode. (d) Na 1s core-level XPS spectra of the Na:MnO2/GCF electrode [33].
Figure 3. (a) The synthesis scheme of the GCF from the raw graphite paper and the photographs of the samples through the synthesis route. (b) The surface SEM of the Na:MnO2/GCF electrode. (c) XRD patterns of the Na:MnO2/GCF electrode. (d) Na 1s core-level XPS spectra of the Na:MnO2/GCF electrode [33].
Energies 16 03221 g003
Figure 4. (a) Scheme of the preparation process of LGP and synthesis of LGP@KxMnO2. (b,c) The cross-sectional SEM images of LGP@K0.15MnO2. (d) Rate performance and (e) cycling stability and coulombic efficiency at 1000 mA/g of K-doped MnO2 grown on the layered graphite paper and carbon cloth [38].
Figure 4. (a) Scheme of the preparation process of LGP and synthesis of LGP@KxMnO2. (b,c) The cross-sectional SEM images of LGP@K0.15MnO2. (d) Rate performance and (e) cycling stability and coulombic efficiency at 1000 mA/g of K-doped MnO2 grown on the layered graphite paper and carbon cloth [38].
Energies 16 03221 g004
Figure 5. (a) Scheme of the synthesis route for GQDs@ZnxMnO2 composite [42]. (b) SEM image of GQDs@ZnxMnO2 [42]. (c) Long-term cycling of GQDs@ZnxMnO2 electrode [42]. (d) Comparison of the cyclic voltammograms for pristine MnO2 and GQDs@ZnxMnO2 cathodes [42]. (e) Schematic illustration of the structure of fabricated Zn-doped mixed Mn3O4-MnO2 nanosheets on the AgCNT- modified Ni foam [27]. (f) Galvanostatic charging/discharging profiles of Zn-doped Mn3O4-MnO2 vertical nanosheets [27].
Figure 5. (a) Scheme of the synthesis route for GQDs@ZnxMnO2 composite [42]. (b) SEM image of GQDs@ZnxMnO2 [42]. (c) Long-term cycling of GQDs@ZnxMnO2 electrode [42]. (d) Comparison of the cyclic voltammograms for pristine MnO2 and GQDs@ZnxMnO2 cathodes [42]. (e) Schematic illustration of the structure of fabricated Zn-doped mixed Mn3O4-MnO2 nanosheets on the AgCNT- modified Ni foam [27]. (f) Galvanostatic charging/discharging profiles of Zn-doped Mn3O4-MnO2 vertical nanosheets [27].
Energies 16 03221 g005
Figure 6. (a) XRD spectra of pristine MnO2 and FMO [44]. (b) SEM image of FMO [44]; (c) HRTEM image of FMO [44]. (d) The mechanism of the catalysis effect in Co-modified δ-MnO2 [45]. (e) Rate performance of plasma-treated Co-MnO2-X (X = treatment time) at current densities 1–5 A∙g−1 [29]. (f) Cycling performance of Co-MnO2-0 and Co-MnO2-2 at 3 A∙g−1 [29].
Figure 6. (a) XRD spectra of pristine MnO2 and FMO [44]. (b) SEM image of FMO [44]; (c) HRTEM image of FMO [44]. (d) The mechanism of the catalysis effect in Co-modified δ-MnO2 [45]. (e) Rate performance of plasma-treated Co-MnO2-X (X = treatment time) at current densities 1–5 A∙g−1 [29]. (f) Cycling performance of Co-MnO2-0 and Co-MnO2-2 at 3 A∙g−1 [29].
Energies 16 03221 g006
Figure 9. (a) Scheme of the reaction between aniline and potassium permanganate and the expanded structure of PANI-intercalated MnO2 nanolayers [66]. (b) Typical galvanostatic charge/discharge curves of PANI-intercalated MnO2 cathode at 0.05 A∙g−1 [66]. (c) XRD patterns and (d) C-rate capabilities of MnO2, MnO2/rGO, and MnO2/rGO/PANI composites [67]. (e) Rate performance and (f) Galvanostatic discharge/charge curves of MnO2/CC, PANI-MnO2/CC and PANI [68]. (g) Nyquist plots for MnO2/CC, PANI-MnO2/CC, and PANI [68].
Figure 9. (a) Scheme of the reaction between aniline and potassium permanganate and the expanded structure of PANI-intercalated MnO2 nanolayers [66]. (b) Typical galvanostatic charge/discharge curves of PANI-intercalated MnO2 cathode at 0.05 A∙g−1 [66]. (c) XRD patterns and (d) C-rate capabilities of MnO2, MnO2/rGO, and MnO2/rGO/PANI composites [67]. (e) Rate performance and (f) Galvanostatic discharge/charge curves of MnO2/CC, PANI-MnO2/CC and PANI [68]. (g) Nyquist plots for MnO2/CC, PANI-MnO2/CC, and PANI [68].
Energies 16 03221 g009
Figure 12. A general classification of working mechanisms of MnO2-based AZIBs: (a) Zn2+ insertion/extraction, (b) Zn2+/H+ (co)insertion/extraction, (c) hybrid process, and (d) electrolytic deposition/dissolution of MnO2 [101].
Figure 12. A general classification of working mechanisms of MnO2-based AZIBs: (a) Zn2+ insertion/extraction, (b) Zn2+/H+ (co)insertion/extraction, (c) hybrid process, and (d) electrolytic deposition/dissolution of MnO2 [101].
Energies 16 03221 g012
Table 1. Electrochemical performance of selected MnO2-based materials doped with metal ions as cathodes in AZIBs.
Table 1. Electrochemical performance of selected MnO2-based materials doped with metal ions as cathodes in AZIBs.
MaterialSynthesis MethodMorphologyElectrolyteSpecific Capacity, mAh g−1 (Current Density, A·g−1)Capacity Retention, (Number of Cycles and Current, A·g−1)Ref.
K-δ-MnO2precipitationtwo-dimensional
nanosheets
2 M ZnSO4 +
0.1 M MnSO4
270.5 (0.1);
220.1 (1);
95.1 (3)
104.6%
(100, 0.3)
[23]
Li0.023Mn0.87O2two-stage hydrothermal method
(140/110 °C)
nanorods2 M ZnSO4 +
0.1 M MnSO4
184 (0.1);
0.54 (2)
89%
(1000, 1)
100%
(100, 0.1)
[32]
Na-δ-MnO2redox reactionnanoplates2 M ZnSO4 +
0.2 M MnSO4
278 (0.308)
161 (1.85)
103 (6.16)
~100%
(2000, 2.46/3.08)
[60]
Na:MnO2/GCFelectrochemical deposition on graphene-like carbon foamnanosheets2 M ZnSO4 +
0.1 M MnSO4
381.8 (0.1)
258.5 (1)
94.8 (3)
~80%
(100, 0.1)
~75%
(1000, 1)
[33]
K0.19MnO2·0.56H2Oone-pot hydrothermal method
(180 °C)
nanosheets3 M Zn(CF3SO3)2107 (1)87.5%
(2000, 10)
[28]
K-δ-MnO2-Vone-pot hydrothermal method
(160 °C)
layered structure2 M ZnSO4 +
0.1 M MnSO4 +
0.1 M K2SO4
288.8 (0.1)
85.7 (1)
91.9%
(1500, 1)
89.4%
(500, 0.6)
[35]
K0.29MnO2 0.67H2Oone-pot hydrothermal method
(180 °C)
nanosheets2.5 M ZnSO4 +
0.2 M MnSO4
300 (0.2)
219 (1)
136 (3)
92 %
(500, 0.2)
[36]
α-K0.19MnO2self-sacrificial template methodnanotubes3 M Zn(CF3SO3)2 + 0.2 M Mn(CF3SO3)2 +
3 M K(CF3SO3)
270 (0.308)
222.8 (0.616)
200 (1.54)
98.5%
(50, 0.308)
92%
(200, 0.616)
90%
(400, 1.54)
[37]
LGP@K0.15MnO2in-situ hydrothermal synthesis (150 °C)nanosheets2 M ZnSO4 +
0.1 M MnSO4
402.6 (0.05)
196.1 (0.8)
116.1 (2)
92.5%
(100, 0.2)
83.3%
(1000, 1)
[38]
KMO-CNT/graphenepolyol
reduction method
nanowires2 M ZnSO4 +
0.4 M MnSO4
373.1 (0.1)
213.6 (1)
108.8 (3)
82.5%
(350, 0.5)
77%
(1000, 3)
[39]
Ca0.28MnO2·0.5H2Oone-step hydrothermal methodinterconnected nanoflakes1 M ZnSO4 +
0.1 M MnSO4
298 (0.175)
277.7 (0.35)
124.5 (3.5)
85%
(1000, 4)
92%
(5000, 3.5)
[40]
Zn-δ-MnO2redox reactionflower-like
nanospheres
2 M ZnSO4 +
0.1 M MnSO4
275 (0.3)
121 (3)
100%
(100, 0.3/0.6)
100%
(500, 1)
[41]
GQDs·ZnxMnO2redox reactionnanoflowers1 M ZnSO4403.6 (0.3)
211.5 (4)
88.1%
(500, 1)
[42]
Zn-doped Mn3O4-MnO2-NSselectrochemical
deposition
vertical nanosheets2 M ZnSO4 +
0.1 M MnSO4
562.1 (0.3)
272.7 (6)
69.4%
(200, 3)
[27]
V-doped δ-MnO2redox reactionnanoparticles1 M ZnSO4266 (0.066)
150 (0.266)
67 (1.064)
52.4%
(100, 0.066)
[25]
V-doped δ-MnO2modified
coprecipitation
nanosheets with aerogel-like
morphology
2 M ZnSO4194 (0.2)
74 (2)
71%
(100, 0.3)
52%
(600, 3)
[43]
Fe-doped δ-MnO2one-step hydrothermal process (120 °C)nanoflowers2 M ZnSO4 +
0.1 M MnSO4
390 (0.1)
320 (1)
160 (3)
86.3%
(200, 1)
[44]
Co/Zn-doped δ-MnO2 on N-doped CCelectrochemical
deposition
film on the carbon nanowires2 M ZnSO4 +
0.07 M MnSO4
280 (1.2)
30 (10.5)
~100%
(600, 1.2)
[26]
Co-doped δ-MnO2molten-salt synthesis processnanosheets2 M ZnSO4 +
0.2 M MnSO4
500 (0.1)
125 (5)
63%
(5000, 2)
100%
(0.3, 100)
[45]
Co-doped α-MnO2
on CC
one-step hydrothermal process (120 °C) + plasma treatmentnanowires2 M ZnSO4 +
0.1 M MnSO4
511 (0.5)
337 (1)
100 (5)
98%
(1000, 3)
[29]
Co-doped σ-MnO2one-step
electrodeposition
nanosheets2 M ZnSO4 +
0.2 M MnSO4 +
0.02 CoAc
313.8 (0.5)91.8%
(1000, 1)
[46]
Ni-doped α-MnO2
(Ni0.052K0.119Mn0.948O2
0.208H2O)
one-step hydrothermal process (120 °C)nanorods3 M ZnSO4 +
0.2 M MnSO4
303 (0.015)71.4%
(2000, 1.232)
[47]
Cu-doped δ–MnO2 on CCone-step
hydrothermal process (160 °C)
nanowires2 M ZnSO4 +
0.2 M MnSO4
398.2 (0.1)
224.9 (1)
124.9 (5)
90.1%
(700, 5)
[48]
Cu0.06MnO2·1.7H2O
(δ–MnO2)
one-step
hydrothermal process (180 °C)
nanoflowers2 M ZnSO4 +
0.1 M MnSO4
493.3 (0.1)
350 (0.5)
125.8 (5)
80%
(150, 0.5)
[49]
Bi0.09MnO2·1.5H2O
(δ–MnO2)
one-step
hydrothermal process (180 °C)
nanoflowers2 M ZnSO4 +
0.1 M MnSO4
175.5 (0.1)
65 (2)
96%
(1100, 1)
72.3 %
(500, 0.5)
[49]
Ag-doped α-MnO2one-step
hydrothermal process (120 °C)
nanowires2 M ZnSO4 +
0.1 M MnSO4
315 (0.05)
177 (0.5)
85 (2)
94.4%
(500, 0.5)
[50]
Mo-doped α-MnO2one-step
hydrothermal process (120 °C)
nanorods2 M ZnSO4 +
0.2 M MnSO4
222.8 (0.1);
65.8 (5)
82.6%
(1000, 2)
[51]
Mo-doped δ-MnO2one-step
hydrothermal method (120 °C)
flower-like
nanospheres
2 M ZnSO4 +
0.1 M MnSO4
327 (0.2)
207 (1)
107 (3)
76.8%
(1000, 1)
[52]
La3+-inserted δ -MnO2redox reactionnanoflorets1 M ZnSO4 +
0.4 M MnSO4
278.5 (0.1)71%
(200, 0.2)
[53]
Ce-doped α- MnO2one-step
hydrothermal method (140 °C)
nanorod-like structure2 M ZnSO4 +
0.1 M MnSO4
134 (1.54)74%
(100, 1.54)
[54]
Ce-MnO2@CCone-step
electrodeposition
porous lamellar structuresPAM/2 M ZnSO4 +
0.1 M MnSO4
292 (0.1)
212 (0.5)
106 (2)
64%
(450, 0.1)
[55]
Al-intercalated α-MnO2one-step
hydrothermal method (140 °C)
nanorodsPVA: 1 M ZnSO4 (1:4)333.6 (1)
198.6 (4)
94.5%
(2000, 2)
[24]
Al-Doped α-MnO2 coated by Ligninone-step
hydrothermal process (200 °C)
1D nanorod
structures
2 M ZnSO4 +
0.2 M MnSO4
~420 (0.1)
180 (1.5)
66.7%
(3000, 1.5)
[56]
Al3+ pre-intercalated
K0.27MnO2·0.54H2O
(δ-MnO2)
modified
hydrothermal method
(160 °C)
spherical
microflowers
2 M ZnSO4 +
0.1 M MnSO4
323.7 (0.1)
250 (0.5)
191.7 (2)
99%
(300, 0.5)
[57]
α-MnO2@KCoAlco-precipitation methodirregular lumpy particles with agglomeration2 M ZnSO4 +
0.05 M MnSO4
524 (0.5)
431 (1)
221 (5)
~66.4%
(100, 0.5)
[58]
Bi-doped α-MnO2redox process followed by annealingnanoparticles2 M ZnSO4 +
0.2 M MnSO4
363 (0.1)
286 (0.6)
197 (1)
93%
(10,000, 1)
[59]
Sn-doped α-MnO2hydrothermal process (180 °C) with further calcinationnanorods2 M ZnSO4 +
0.1 M MnSO4
210 (0.1)
106 (1)
80 %
(500, 1)
[30]
Table 2. Electrochemical performance of selected MnO2-based composites with conducting polymers as cathodes in AZIBs.
Table 2. Electrochemical performance of selected MnO2-based composites with conducting polymers as cathodes in AZIBs.
MaterialSynthesis MethodMorphologyElectrolyteSpecific Capacity, mAh g−1 (Current Density, A·g−1)Capacity Retention, (Number of Cycles and Current, A·g−1)Ref.
δ-MnO2@polyanilinegas/liquid interface reactionmesoporous
nanohybrids
2 M ZnSO4 +
0.2 M MnSO4
313 (0.1)
145 (1)
88 (3)
~100%
(500, 0.5)
[65]
Polyaniline-
intercalated
δ-MnO2
one-step inorganic/
organic interface
reaction
nanolayers with spongiform structure2 M ZnSO4 +
0.1 M MnSO4
298 (0.05)
280 (0.2)
110 (3)
90%
(200, 0.2)
40%
(5000, 2)
[66]
Polyaniline-coated
β-MnO2/rGO
MnO2 ball-milling + hydrothermal process with rGO (160 °C) +
in situ polymerization
aerogel-supported2 M ZnSO4241.1 (0.1)
111.7 (1)
82.7%
(600, 1)
[67]
PANI-δ-MnO2/CChydrothermal method (150 °C) + in situ polymerizationnanosheets2 M ZnSO4 +
0.1 M MnSO4
286 (0.5)
233 (2)
177 (4)
96.9%
(9000, 4)
[68]
α-MnO2@PANIhydrothermal process (160 °C) + in situ interfacial polymerizationcore-shell2 M ZnSO4 +
0.1 M MnSO4
342 (0.2)
100 (3)
82%
(2000, 2)
[69]
α-MnO2/PPyhydrothermal process (160 °C) + in situ polymerizationnanorods2 M ZnSO4 +
0.1 M MnSO4
256 (0.1)
104 (1)
100%
(500, 1)
100%
(50, 0.1)
[78]
β-MnO2/PPyone-step
hydrothermal process (120 °C)
micro-spherical structure of nanowires and clusters of nanorods2 M ZnSO4 +
0.1 M MnSO4
215.4 (0.1)
214.1 (0.2)
171.5 (0.5)
69.9 (1.5)
100%
(160, 0.2)
[79]
CNT/α-MnO2-PPyin situ reactive self-
assembly and following vacuum filtration
core-shell
structure and
rod-shaped
morphology
2 M ZnSO4 +
0.1 M MnSO4
253.9 (0.3)
83.3 (2)
87.4%
(1000, 1)
75.5%
(200, 0.3)
[81]
α-MnO2/rGO-PPyhydrothermal process (140 °C) + in situ polymerizationnanowires wrapped by PPy3 M Zn(CF3SO3)2438.3 (0.1)
248.8 (0.5)
~85.9%
(100, 0.5)
[82]
Mn2O3/α-MnO2@PPymolten salt method + self-initiated
polymerization
nanobelts and
nanoparticles
2 M ZnSO4 +
0.2 M MnSO4
289.9 (0.2)
252.6 (1)
199.8 (3)
~100%
(1000, 3)
96.7%
(1000, 1)
[83]
Fe-doped α-MnO2 coated by PPychemical precipitation
method + in situ polymerization
nanoparticles2 M ZnSO4 +
0.1 M MnSO4
270 (0.1)
164 (0.4)
73 (1)
99.6%
(100, 0.1)
[84]
α-MnO2/PPy@SSelectrodepositionnanocrystallites1 M ZnSO4 +
0.1 M MnSO4
143.2 (0.308)
102.2 (0.924)
86.8 (1.54)
74.2%
(850, 1.54)
[85]
MnO2@PEDOTelectrodepositionnanosheets2 M ZnCl2 +
0.4 M MnSO4
366.6 (0.74)
143 (7.43)
83.7%
(300, 1.11)
[90]
PEDOT@Co-MnO2low-temperature
hydrothermal process + electrochemical polymerization
nanoflakes2 M ZnSO4298.9 (1)92.3%
(1000, 5.0)
[92]
δ-MnO2/α-MnO2
/PEDOT
decomposition (δ -MnO2) + hydrothermal process (150 °C, α-MnO2) + electrodeposition (PEDOT)nanowires of
δ-MnO2 and nanoflakes of α-MnO2
2 M ZnSO4 +
0.1 M MnSO4
360.5 (0.031)
174.5 (0.308)
94 (1.54)
78%
(860, 0.308)
[93]
δ-MnO2@PEDOTredox reactionnanowires2 M ZnSO4 +
0.2 M MnSO4
242 (0.2)
133 (1)
120.7 (2)
85.1%
(1000, 2)
[95]
VG-α-MnO2 coated with PEDOT:PSShydrothermal process (150 °C)MnO2 nano-
particles on VG nanosheets with 3D porous structure
1 M ZnSO4 +
0.1 M MnSO4
367.4 (0.5)
280.5 (1)
148.2 (6)
73.7%
(1000, 5)
[98]
K0.46Mn2O4·1.55 H2O (δ-MnO2)/PEDOT:PSShydrothermal method (160 °C) +mechanical mixing with
PEDOT:PSS
nanoflowers2 M ZnSO4 +
0.1 M MnSO4
380 (0.3)
243 (1)
40 (5)
100%
(120, 0.3)
[97]
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

Kamenskii, M.A.; Volkov, F.S.; Eliseeva, S.N.; Tolstopyatova, E.G.; Kondratiev, V.V. Enhancement of Electrochemical Performance of Aqueous Zinc Ion Batteries by Structural and Interfacial Design of MnO2 Cathodes: The Metal Ion Doping and Introduction of Conducting Polymers. Energies 2023, 16, 3221. https://doi.org/10.3390/en16073221

AMA Style

Kamenskii MA, Volkov FS, Eliseeva SN, Tolstopyatova EG, Kondratiev VV. Enhancement of Electrochemical Performance of Aqueous Zinc Ion Batteries by Structural and Interfacial Design of MnO2 Cathodes: The Metal Ion Doping and Introduction of Conducting Polymers. Energies. 2023; 16(7):3221. https://doi.org/10.3390/en16073221

Chicago/Turabian Style

Kamenskii, Mikhail A., Filipp S. Volkov, Svetlana N. Eliseeva, Elena G. Tolstopyatova, and Veniamin V. Kondratiev. 2023. "Enhancement of Electrochemical Performance of Aqueous Zinc Ion Batteries by Structural and Interfacial Design of MnO2 Cathodes: The Metal Ion Doping and Introduction of Conducting Polymers" Energies 16, no. 7: 3221. https://doi.org/10.3390/en16073221

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