Home / Publications / Hard carbon as anode material for metal-ion batteries

Hard carbon as anode material for metal-ion batteries

Share
Cite this
GOST
Cite
GOST copy
Elena N. Abramova et al. Hard carbon as anode material for metal-ion batteries // Russian Chemical Reviews. 2024. Vol. 93. No. 2. RCR5100
GOST all authors (up to 50) copy
Zoya V. Bobyleva, Oleg A. Drozhzhin, Artem M. Abakumov, Evgeny V. Antipov, Elena N. Abramova Hard carbon as anode material for metal-ion batteries // Russian Chemical Reviews. 2024. Vol. 93. No. 2. RCR5100
 | 
RIS
Cite
RIS copy
TY - GENERIC
DO - 10.59761/RCR5100
UR - https://rcr.colab.ws/publications/10.59761/RCR5100
TI - Hard carbon as anode material for metal-ion batteries
T2 - Russian Chemical Reviews
PB - Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii
AU - Bobyleva, Zoya V.
AU - Drozhzhin, Oleg A.
AU - Abakumov, Artem M.
AU - Antipov, Evgeny V.
AU - Abramova, Elena N.
PY - 2024
SP - RCR5100
IS - 2
VL - 93
ER -
 | 
BibTeX
Cite
BibTeX copy
@misc{2024_Abramova,
author = {Zoya V. Bobyleva and Oleg A. Drozhzhin and Artem M. Abakumov and Evgeny V. Antipov and Elena N. Abramova},
title = {Hard carbon as anode material for metal-ion batteries},
month = {mar},
year = {2024}
}
 | 
MLA
Cite
MLA copy
Abramova, Elena N., et al. “Hard carbon as anode material for metal-ion batteries.” Russian Chemical Reviews, vol. 93, no. 2, Mar. 2024, p. RCR5100. https://rcr.colab.ws/publications/10.59761/RCR5100.
Publication views
56

Keywords

carbon-based anode materials
electrochemical performance
hard carbon
materials research techniques
metal-ion batteries

Abstract

The development of large-scale energy storage systems based on the mature technology of lithium-ion batteries is hampered by the high cost of lithium. Therefore, analogous technologies based on other alkali metals (sodium and potassium) are being developed. Among various types of negative electrode (anode) materials for such batteries, carbon materials are most promising; in particular, hard carbon is of most interest. The present review addresses the current state of research on the structure, composition and properties of this type of material and gives analysis of methods of its preparation and investigation. Description of the microstructure of hard carbon is a highly ambiguous and challenging problem; therefore, the review pays special attention to various microstructural models. In addition, the methods of synthesis are systematized and the results of studies of the physicochemical properties of hard carbon are analyzed. The correlations between the preparation method, characteristics and electrochemical properties in metal-ion batteries are identified. A large array of results of electrochemical studies is considered, the views on the mechanisms of electrochemical interactions of Na+ and K+ cations with hard carbon are systematized, and the currently existing contradictions in various models of the interaction mechanisms are shown.

Bibliography — 247 references.

1. Introduction

The continuous growth of energy consumption has led to the development of energy storage technologies. The most effective of them are based on the use of batteries as electrochemical power sources. Widely used lead-acid batteries suffer from a number of disadvantages, the most important of which are low specific capacity (25 – 40 W h kg–1)[1], short lifetime, high toxicity of lead.

Lithium-ion (Li-ion) batteries have none of these drawbacks, therefore they now dominate the market for portable electronics and electric transport[2]. However, the use of Li-ion batteries for many large-scale applications (stationary energy storage, power plant buffer systems, electric passenger transport, etc.) is hampered by the high cost of lithium and by the fact that its natural sources are localized at a few places in the world. Replacing lithium with its alkali metal analogues (sodium and potassium) appears to be a promising alternative due to the abundance of Na and K in the Earth’s crust (2.5 and 1.7 wt.% respectively)[3] and their low cost compared to lithium. In addition, lithium forms alloys with aluminum during electrochemical charging, forcing battery manufacturers to use more expensive and heavier copper foil as the current collector for the negative electrode (anode). Sodium and potassium are free of this shortcoming, so aluminum foil can be used both for positive electrodes (cathodes) and for anodes, which is an additional factor in reducing the cost of sodium and potassium ion batteries (SIBs and PIBs) due to the lower cost of aluminum. It is also worth noting the fundamental similarity of LIB, NIB and PIB technologies, which makes it possible to apply the technological solutions developed for LIB to SIB and PIB, including those for the production of electrodes, cells, batteries, monitoring and control systems.

In addition, the electrode potentials of the Na/Na+ and K/K+ redox pairs in the propylene carbonate-based electrolyte are close to the electrode potential of the Li/Li+ pair (higher by 0.23 V and lower by 0.09 V respectively)[4]. This allows for high operating voltage and energy density of SIBs and PIBs and makes their characteristics potentially comparable to those of LIBs. Despite the fact that sodium and potassium ions have larger weights and sizes relatively to lithium ions (the ionic radii of Na+ and K+ are 1.02 and 1.38 Å, respectively, the ionic radius of Li is 0.76 Å), the sizes of solvated sodium and potassium ions in propylene carbonate are 4.6 and 3.6 Å, respectively, which is smaller than the Stokes radius of lithium ions (4.8 Å)[5]. Therefore, electrolytes for Na-ion and K-ion electrochemical systems may have higher ionic conductivity, which will reduce the overall resistance of SIB and PIB cells[5].

Cathode materials for these electrochemical systems include polyanionic compounds (e.g., phosphates, fluoride phosphates, etc.), oxide materials and Prussian blue analogues. The specific capacity of SIB and PIB cathode materials is lower than similar metrics of LIB materials, which is mainly due to higher values of ionic radii and atomic weights of potassium and sodium.

Among the various anode materials for metal-ion batteries, carbonaseous materials have attracted the most interest[6-10] due to their low cost, high electronic conductivity and ability for reversible electrochemical interaction with sodium and potassium ions. However, graphite, which is widely used as an anode material in LIBs, has an extremely low specific capacity (~35 mA h g–1) in SIBs and a significant volume change of the material (up to 61%) in PIBs, which makes it technologically unpromising in both cases[11, 12]. In this context, the attention of the researchers and developers of SIBs and PIBs is drawn to other carbon materials, primarily hard carbon. However, it should be noted that the main research is focused on the use of hard carbon as an anode material in sodium-ion electrochemical systems. The reasons for this seem to be the high performance characteristics of hard carbon in SIBs (specific capacity, cycling stability, Coulombic efficiency), as well as a higher maturity of the technology itself compared to PIBs, which prospects are not rather clear.

The first review devoted to the use of hard carbon as an anode material for SIB appeared in 2015[3]. In this paper, the authors used results from the study of their own samples by gas adsorption-desorption, X-ray diffraction and Raman spectroscopy to describe the properties of the material. However, there was no comparison of these results with those of other authors. The main focus of this paper was the electrochemical properties of hard carbon in sodium-ion electrochemical systems.

Mittal et al.[13] gave general characteristics of hard carbon and noted that the morphology of the material depends on the synthesis conditions. The main issue of this paper is the mechanism of electrochemical interaction of hard carbon with sodium cations and a review of methods for studying the material. Another review[14] also highlights the electrochemical properties of hard carbon in sodium-ion systems. It classifies models of the interaction of the material with Na+, considers practical issues of the synthesis of hard carbon and the preparation of sodium-ion batteries with improved electrochemical properties.

A comprehensive analysis of the microstructure of hard carbon is given in another review paper[15]. It should be noted that reviews devoted to hard carbon in SIBs require additions summarizing the results of characterization of its composition and structure using modern research methods. The application of this material to other electrochemical systems, such as PIBs and LIBs, is also relevant for generalization and analysis.

The subject of the present review is therefore hard carbon as a material for negative electrodes of metal-ion batteries. It has several objectives: 1) to reflect the state-of-the-art understanding of the microstructure of hard carbon and the mechanisms of its interaction with alkali metal cations; 2) to systematize the methods used in materials research and the information obtained from them; 3) to find out the correlations between the methods of synthesis of hard carbon, its microstructure and electrochemical properties; 4) to analyze the results of the application of this material in three electrochemical systems — LIBs, SIBs and PIBs.

2. The definition of hard carbon and 3 models of its microstructures

Among all the variety of anode materials, a special role in the development of SIBs and PIBs is played by non-graphitizable carbon (or non-graphitic carbon, also termed as ‘hard carbon’).

The term ‘hard carbon’ refers to carbon materials that are not graphitized by heat treatment at high temperature, up to 3000 °C, in an inert environment. Samples of hard carbon have a disordered microstructure which differs from that of graphitizable carbon (termed as ‘soft carbon’)[16, 17]. If graphitizable carbon gradually forms the graphite structure during high-temperature pyrolysis, the ‘final’ form of non-graphitizable carbon is glassy carbon[18]. This is why many studies of the macro- and microstructure of hard carbon are closely intertwined with studies of glassy carbon[19, 20].

The first models of the structure of hard carbon were proposed in the 1950s. Among them, it is worth mentioning the packet and fringe model (other names include fringed micellar, crystallite) (Figure 1a), proposed by V.I.Kasatochkin (see [21] and references therein). Another model, the ‘house of cards’ one, was proposed by R.Franklin[22, 23] in 1951 (Figure 1b). According to these models, hard carbon is a set of randomly arranged crystallites (domains) consisting of several graphitic atomic planes (4 – 6 layers) with a lateral size of about 40 Å. The domains are in turn connected by a ‘fringe’ of linear carbon chains.

Figure 1
Hard carbon microstructure models: (a) packet and fringe Kasatochkin’s model[21], (b) ‘house of cards’ Franklin’s model[22, 24] (c) Jenkins’ model[25] and (d) Harris’ model[26].

It was later suggested (Jenkins’ model, 1972, Figure 1c)[25], that hard carbon consists of intertwined graphitic ribbons — randomly oriented and interconnected twisted microfibrils with a higher concentration of voids and hence lower material density than graphite (~1.5 vs 2.3 g cm–3).

A new concept of the microstructure of hard carbon is based on the model of P.Harris (1997) developed as a result of the active study of fullerene-like structures[26-28]. According to this model, hard carbon consists of fragments of bent graphene-like layers containing five-, six- and seven-membered rings (Figure 1d).

At present, machine learning methods have greatly simplified the modelling of the microstructure of amorphous hard carbon. Recent studies[29, 30] have shown that the density of hard carbon is inversely proportional to the diameter of the micropores. It was found[29] that five- and seven-membered rings are inherent in the microstructure of amorphous carbon, but that five-membered motifs are more common than seven-membered or other, rarer, types of defects in the graphene-like sheet, for example Stone-Wales defects (which arise due to rotations of C – C bonds through an angle of ∼90° in the hexagonal structure, bringing two pentagonal and two heptagonal rings into the structure).

Hard carbon has defects, mainly in the form of dangling bonds at the edges of graphite-like domains and vacancies in graphene-like layers. In addition, these materials often contain heteroatoms such as O, N, S (the latter found in the case of nitrogen- or sulfur-containing precursors used for synthesis). The ratio of carbon atoms with different types of hybridization, the presence of heteroatoms, as well as material parameters such as porosity, defectiveness, etc. are determined by the choice of precursor and synthesis conditions. The microstructure of hard carbon, including the mutual arrangement of atoms and groups of atoms, layers of carbon atoms, hybridization of C atoms, as well as micropores of the material resulting from the peculiarities of the mutual arrangement of layers of C atoms, defects in the material, remains a relevant and fundamentally important issue for the understanding of the mechanism of intercalation/deintercalation of alkali metal ions and the optimization of the microstructure of the material for its practical application.

3. Synthesis of hard carbon

Hard carbon materials are obtained by high temperature annealing (pyrolysis) of organic raw materials in an inert atmosphere. The synthetic route can be divided into several stages: pretreatment of the precursor, high-temperature annealing of the pretreated product or pure precursor, and sometimes post-processing of the annealed product[31].

The precursors most commonly used to produce hard carbon can be broadly categorized into carbohydrates, including cellulose[32-37], and synthetic polymers such as PAN (polyacrylonitrile), PET (polyethylene terephthalate), PFA (perfluoroalkoxide polymers), phenolic resins and pitches[38, 39], lignin[40-46]. Since hard carbon is a very promising material for large-scale production, a number of environmentally friendly technologies are being developed to produce it from a variety of low-cost materials, including biomass[47-69]. For example, Lakienko et al.[70] have recently proposed a technology for the production of hard carbon from Sosnowskyi hogweed, an invasive and widespread plant. The key step in the proposed synthesis is acid washing prior to carbonization, which helps to improve the Coulombic efficiency of the first charge-discharge cycle up to 87% (initial Coulombic efficiency, ICE).

The choice of precursor is determined by its availability in the region, the cost and the yield of hard carbon. Irisarri et al.[71] noted a high yield (50%) of product derived from phenol formaldehyde resin compared to that from lignin and microcrystalline cellulose. A high product yield (40 – 67%) is also typical for PAN-derived hard carbon[72]. Górka et al.[73] considered biomass as a cheap precursor for the hard carbon, but reported low yields of hard carbon derived from biomass and various sugars (up to 10%) except for lignin (up to 50%). Abramova et al.[74] reported a higher yield of hard carbon (10 – 28%) when polytetrafluoroethylene (PTFE) was added to sucrose.

The production of hard carbon is quite similar to the technological process of producing activated carbon, since the key step of the synthesis is also high-temperature annealing[75]. The main difference is that the production of hard carbon anode materials does not require the activation step, as the creation of mesoporosity contributes to the deterioration of the electrochemical properties.

The first step of the synthesis is usually a preliminary heat treatment of the precursor in air at temperatures above 100 °C. The main objectives of the pretreatment are dehydration of the feedstock and formation of the desired morphology of the precursor. Thermal pretreatment of sugars is called caramelization. It is a well known process in the food industry. Pretreatment is also carried out under hydrothermal, solvothermal and microwave hydrothermal conditions[76-79]. Varying the conditions of this process makes it possible to obtain micro- or nanoparticles of microspherical shape. Both caramelization and hydro- and solvothermal carbonization are complex processes. During caramelization, the crystalline structure of the sugars is destroyed to give amorphous glasses with different C : H : O ratios[80]. In hydro- and solvothermal carbonization, sugars are hydrolyzed to form monosaccharides, which subsequently undergo dehydration and condensation and produce both liquid and solid carbon-rich products[80]. The temperature of the process is chosen according to the composition of the precursors[56, 66, 81].Pretreatment is not a mandatory step in the synthesis of hard carbon, but pre-dehydration followed by high-temperature carbonization results in a material with a lower specific surface area, which usually improves the — ICE[36]. Bobyleva et al.[82] showed that the conditions of thermal pretreatment of glucose in a narrow temperature range (200 ± 40 °C) have a decisive influence on the morphology and specific surface area of hard carbon. It has been found that reducing the specific surface area, the degree of defectiveness and the oxygen content on the surface can improve the ICE of hard carbon up to 89% in sodium-ion half-cells.

Pretreatment is also carried out to remove impurities such as potassium, calcium, magnesium and silicon, which are present in biomass[73]. Washing with water or solutions of acids and alkalis removes impurities and has a positive effect on the electrochemical properties of the biomass-derived hard carbons[65, 83-85].

The key step in the synthesis of hard carbon is carbonization — high temperature annealing at temperatures between 900 and 2600 °C in an inert atmosphere. During this stage, heteroatoms and remaining functional groups are removed from the precursor or a pretreated carbonaceous product[86]. As the temperature increases, the ratio of carbon to oxygen and other heteroatoms increases significantly. For the glucose-derived materials subjected to hydrothermal pretreatment, the molar ratio of carbon to oxygen, as determined by X-ray photoelectron spectroscopy, changes from 15.9 at 1000 °C to 74.9 at 1900 °C[87].

Carbonization temperatures up to 1000 °C do not allow removing of a significant portion of heteroatoms, which affects the electrochemical performance of the material, as sodium and potassium ions are able to interact with heteroatomic functional groups[72]. Such interaction is mostly irreversible, so higher carbonization temperatures are used for the synthesis of hard carbon, as this helps to improve the electrochemical properties.

A number of studies confirmed the dependence between the electrochemical properties, capacity and Coulombic efficiency of the anode material and the carbonization temperature for various precursors (see, e.g., Ref. [86]). This dependence is not direct — for example, the maximum capacity for hard carbon in SIBs can be reached as annealing temperatures varies in the range of 1300 – 1500 °C. With further increase in temperature, the capacity begins to decrease (Figure 2). The nature of this process and possible reasons for the decrease in capacity are discussed below.

Figure 2
(a) Galvanostatic charge/discharge curves of hard carbons obtained via hydrothermal carbonization followed by annealing (HG samples) or via direct annealing (DG samples) of glucose at different annealing temperatures (from 1000 to 1500 °С) for the same samples: (b) difference in slope and plateau capacities; (с) cyclic performance at different current densities[79].

It should be noted that although the annealing temperature of 1300 – 1500 °C makes it possible to obtain materials with good electrochemical performance, attempts are being made to reduce it to 1000 – 1100 °C in order to reduce the energy costs of material production and the corresponding technological challenges[88]. Problems associated with the presence of heteroatoms and the high specific surface area of such materials are solved by more complex pretreatment of the precursor, as well as the use of dehydrating agents that help to remove water therefrom.

The post-processing of hard carbon includes grinding in a ball mill, removal of mechanical impurities from the material and additional drying to remove adsorbed water from the carbon surface prior to preparation of the electrode slurry.

There are other less common methods of synthesis of hard carbon, such as

1) plasma-chemical synthesis, which involves deposition of nano-sized carbon on a carrier from a gaseous medium containing hydrocarbons (e.g. xylene[61]) and plasma generated by a vacuum-arc discharge;
2) laser ablation of carbon fibres[89];
3) template synthesis, widely used to prepare carbonaceous materials with micropores. Zeolites[90, 91], silica[92], magnesia[93] and zinc oxide[94-96] can be utilized as the template. After carbonizing a mixture of the template and precursor, the former is washed out using alkali or acid solutions. To reduce the specific surface area, the carbonaceous material obtained after washing is subjected to repeated carbonization[93].

As a rule, the presence of heteroatoms, impurities degrade the electrochemical performance of hard carbon due to irreversible interaction with sodium/potassium cations. However, there is an area of research aimed to make this interaction reversible, which will allow to improve the electrochemical capacity of doped hard carbon materials. The main dopants are nitrogen[97-99], sulfur[100, 101], phosphorus[102] and boron[103, 104]. To prepare doped hard carbons, special precursors are used which already contain the necessary heteroatoms. For example, to obtain nitrogen-doped carbon materials, polymer precursors are used such as PAN[72], polyaniline (PANI)[105] and polypyrrol (PP)[106, 107]. The doping process can also be carried out through carbonization of the precursor in a mixture with a source of heteroatoms, or by annealing in an inert atmosphere with the addition of gases containing the required heteroatoms[108]. Also of interest is the co-doping of hard carbon material with multiple heteroatoms to create a larger number of probable interaction sites and boost the discharge capacity[109-112].

4. Methods for studying the composition and structure of hard carbon

A number of modern analytical methods are used to study hard carbon. The main methods for studying the chemical composition of hard carbon include CHNS/O elemental analysis, energy-dispersive X-ray spectroscopy (EDS) and X-ray photoelectron spectroscopy (XPS), particles morphology and open pores are studied using scanning electron microscopy (SEM) and gas adsorption/desorption, respectively. The microstructure of the material under consideration is studied by means of transmission electron microscopy (TEM), X-ray diffraction (XRD), Raman spectroscopy, pair distribution function (or pair distance distribution function (PDF)) based on the analysis of X-ray total scattering, synchrotron and neutron scattering data, small-angle X-ray scattering (SAXS), atomic force microscopy (AFM), scanning tunnelling microscopy (STM), electron paramagnetic resonance (EPR) and others techniques.

When studying hard carbon by various methods, the data obtained are usually compared with those of graphite (Figure 3). Figure 3 shows the results of studies on these materials using different techniques.

Figure 3
(а) Schemes of graphite and hard carbon structures (graphene-like sheets are indicated by black bars); (b) X-ray diffraction pattern of hard carbon and position of X-ray reflections of graphite; (c) Raman spectra; (d) small-angle X-ray scattering spectra[15, 86, 113].

4.1. Chemical composition

Elemental analysis is widely used to study the chemical composition of hard carbon. It helps to determine the chemical composition of both the final material and the products of the processing of precursors at different stages of the hard carbon production. Titirici et al.[114] showed that the precursors obtained via the hydrothermal treatment of various raw materials (sugars, dehydrated hydrocarbon derivatives, etc.) contain the elements C, O, H. Examination of hard carbon after the high-temperature carbonization shows that it consists predominantly of carbon, irrespective of the starting materials, and that as the annealing temperature increases, the carbon atom ratio increases while the ratio of other atoms decreases[59, 115]. As the annealing temperature is further increased, the ratio of O atoms becomes comparable to the measurement error.

XPS is another method of studying the chemical composition, which provides data on the chemical composition of the surface of hard carbon samples[116-122]. [ type="anchor" figureId="977" ]] shows a typical XPS spectrum of hard carbon.

Figure 4
XPS spectra of hard carbon samples obtained from corncobs at different carbonization temperatures: (a) wide spectrum[116], (b) fitted C1s XPS spectra of a sample carbonized at 1000 °C[116]; (c) fitted O1s spectra of Ganoderma lucidum residue-derived samples carbonized at 800, 900 and 1000 °C[118].

The XPS spectra indicate that the surface of hard carbon contains carbon as the main component, corresponding to the C1s line. In various studies, this peak is characterized by binding energies in the range of 280 to 292 eV.

Moreover, oxygen is often found on the surface of hard carbon. Oxygen-containing functional groups are detected in the energy range of the C1s peak as well as at binding energies of around 532 eV, where the O1s spectral line is identified. Among the oxygen-containing groups, (C – O), (C=O), (O – C=O), (COOH), (C – OH) and the H2O molecule have been identified in various studies[11, 32, 33, 56, 58, 65, 66, 69, 72, 117-122].

In addition to carbon and oxygen, nitrogen-containing groups (C – N), (C=N), (N – C=O) have also been detected on the surface of hard carbon by the XPS technique[56, 65, 117, 118, 121]. In addition to the above atoms, small amounts of P and S were detected in the Ganoderma lucidum residue-derived material[118].

The studies on the composition of hard carbon samples carbonized at different temperatures show that the content of heteroatoms decreases with increasing carbonization temperature, in particular the content of oxygen groups (C=O) and (C – O). At the same time, the (C=O) peak decreases faster than that of the (C – O) group (Figure 4c)[118]. This may influence the mechanism of energy storage in the material (as noted by the authors of paper[118], a reversible reaction — C=O + Na+ + e- ↔ –C – O – Na is possible).

4.2. Morphology of materials

SEM is the main method used to study the morphology of hard carbon. Comparison of SEM images of samples obtained from different precursors by different processing techniques[16, 32, 33, 42, 53, 65, 66, 69, 81, 117, 118, 56-59, 122-130] shows a great diversity in the morphology of the materials. The particles of samples derived from carbohydrates by caramelizing the precursor in air at 180 °C before carbonization (Figure 5a)[82], have an asymmetric shape that depends on the milling conditions of the sample[82, 131, 132]. Hard carbon samples obtained from glucose or sucrose by hydrothermal pretreatment at 180 °C are characterized by spherical particle shapes (Figure 5b)[35, 74, 132]. It is worth noting that it is extremely difficult to favour one or the other type of morphology, as both types of materials can exhibit quite attractive electrochemical performance.

Figure 5
SEM microphotographs of hard carbon samples derived from: (a) glucose via caramelization (the image was taken from authors archive to Ref. [77]); (b) glucose via hydrothermal carbonization (the image was taken from authors archive to Ref. [74]); (c) cellulose (the image was taken from authors archive to Ref. [132]); (d) Sosnowskyi hogweed (the image was taken from authors archive to Ref. [70])

Analysis of the morphology of hard carbon samples derived from different biomasses indicates that it inherits the morphology of the starting materials (Figure 5c,d)[70, 132].

4.3. Microstructure of hard carbon

High resolution TEM and electron diffraction, Raman spectroscopy, XRD and PDF methods are used to study the microstructure of hard carbon. The key challenges solved by these methods are to determine the influence of synthesis conditions on the microstructure of hard carbon (due to the special scientific and practical interest, the study of defects is considered in a separate Section).

For example, Zhang et al.[72] showed that PAN-derived hard carbon has a highly disordered microstructure at a carbonization temperature below 1000 °C. As the carbonization temperature is increased up to 2000 °C, the formation of randomly oriented domains consisting of bent sheets of carbon atoms is observed. Further increase in the carbonization temperature result in a significant rise in the domain length. Damodar et al.[127] found that in hard carbon obtained from sepals of Palmyra palm fruit calyx, the appearance of domains of several graphene-like layers is observed already at the carbonization temperature above 700 °C and there is a tendency to increase the number of such domains with increasing carbonization temperature. Similar changes, from virtually amorphous microstructure to the formation of nano-sized graphite-like domains, are also observed in the natural cotton biomass-derived hard carbon[126] with increasing carbonization temperature from 1300 to 1600 °C. Simone et al.[133] also observed an increase in ordering of the samples of cellulose-derived hard carbon with increasing carbonization temperature. TEM clearly shows an increase in the degree of graphitization of the material with increasing carbonization temperature (Figure 6)[58, 60, 81, 118, 128, 133, 134]. This trend is also confirmed by electron diffraction: a number of publications[66, 75, 114, 117, 120, 124, 52-54] report a decrease in the width of the diffraction rings with increasing carbonization temperature. It indicates an increase in structural ordering. At the same time, the temperature of graphite-like domains genesis varies for different precursors and depends on their pretreatment.

Figure 6
TEM images of a sample of hard carbon derived from mangosteen shell carbonized at 800 °С (a), 1300 °С (b), 1500 °С (c), 1600 °С (d).

It is worth noting that during carbonization of hard carbon derived from various precursors (e.g. coffee waste[129], resorcinol-formaldehyde resin[135]) at temperatures above 2000 °C, a tendency towards further ordering of the microstructure is observed. The increase in temperature from 2000 to 3000 °C causes parallel sheets of carbon atoms to first bend and then straighten during carbonization up to 2500 °C, but graphite formation does not occur.

In addition to visualizing the material microstructure, a number of studies[60, 65, 66, 69, 125, 126] have used high-resolution TEM to determine the distance between the layers of carbon atoms (d ) in graphite-like domains, which typically varies from 0.37 (Ref. [60]) to 0.44 nm (Ref. [65]) depending on the carbonization temperature of the samples. It should be noted that these values are always larger than d002 of graphite (0.335 nm).

Similar results are obtained with PXRD. X-Ray diffraction patterns of hard carbon show two strongly broadened reflections at 2θ ~ 23° (22 – 24°) and 43° (43 – 44°) (λCuKα = 1.54 Å), which are ascribed to the corresponding (002) and (100) crystallographic planes of graphite (see Figure 3b)[86]. The average distance between the graphene-like layers varies from 0.347 (Ref. [136]) to 0.443 nm (Ref. [137]) depending on the type of precursor and carbonization temperature. The most common values are 0.37 – 0.38 nm.

It is noteworthy that some papers[40, 66, 138] reported a ‘bell-shaped’ dependence of the interlayer distance on the annealing temperature: an increase was observed with an increase in temperature from ~800 to 1250 °C[40] and a decrease with a further increase in temperature, which may indicate the existence of an ‘intermediate’, more disordered state.

Li et al.[34] and Xiao et al.[35] studied the influence of other carbonization parameters (in addition to the carbonization temperature) on hard carbon structural properties. More ordered structures were observed at slower heating rate[35]. Additional microwave treatment of hard carbon after its annealing leds to an increase in the length of carbon domains to almost 15 nm, while without additional microwave treatment this parameter did not exceed 5 nm[34].

Since the strong broadening of reflexes in X-ray powder diffraction patterns provides only a rough estimation of the structure of hard carbon, the pair distribution function method is used for a more in-depth study. The atomic structure of hard carbon, its defects and the mechanisms of interaction of metal ions with the material are studied using synchrotron X-ray and neutron total scattering data[34, 81, 115, 139-143]. The relevance of this method for hard carbon is due to the nanometer dimensions of graphitic (graphite-like) domains (crystallites), which violates the assumption of the presence of a long-range order in the crystal lattice, and to local defects in the atomic structure. The pair distribution function describes the atomic structure around a selected atom by the probability of finding another atom at a given distance from it. The graph of the pair distribution function G(r) contains peaks at a distance r from the conditionally central atom, which are caused by the corresponding atoms in the material structure (Figure 7).

Figure 7
(a) Graph of the pair distribution function for hard carbon[142], (b) schematic image based on paired distribution function data, showing the nearest neighborhood of the central atom in the graphene layer.

The position of the maxima determines the interatomic distance. The intensity of each peak reflects the relative ‘contribution’ of atoms in the corresponding interatomic distance. The width of the peaks is related to the factors that can cause changes in the interatomic distances, primarily thermal fluctuations of the atoms[115]. The distances r determined in hard carbon using PDF correspond to different distances of the hexagons of graphene sheets[115, 139-141]. In this case, as noted by the authors of papers[104, 136], the peak at a distance of 3.35 – 3.45 Å, corresponding to carbon atoms in adjacent graphite sheets, is practically absent regardless of the annealing temperature (it should be noted that its intensity is quite low in the case of graphite)[115, 139]. At higher annealing temperatures, the peaks in the G(r) graph become narrower, indicating a greater ordering of the atomic structure.

Kubota et al.[131] showed that the maxima on the G(r) plot at values of r greater than 3 Å for hard carbon differ in position and intensity from graphite. This may be due to the non-parallel arrangement of individual layers in hard carbon, defects at their edges, as well as the curvature of the layers, which in turn may be associated with a some of 5- and 7-membered carbon rings and other factors[115, 139-142].

The degree of ordering, or graphitization of hard carbon is estimated from Raman spectroscopy data[33, 56, 61, 117, 119, 121, 122, 130, 144].There are several approaches to estimating the degree of graphitization. They are based on the study of D and G bands, which are characteristic of carbon materials with different microstructure and degree of ordering. The G band is observed at a frequency of ~1580 – 1600 cm–1 and is caused by in-plane vibrations of sp2 carbon atoms in six-membered carbon rings (the E2g mode of the irreducible representation of the D6h group)[145, 146]. At ~1350 – 1360 cm–1, the D band (D1) is observed, which is not characteristic of ideal graphitic lattice[147]. Its nature is the subject of long scientific debate. The D band has been associated with defects at the edges and inside graphitic planes (Ag1 mode)[146], with symmetric stretching vibrations of six-membered aromatic rings[129], or with double-resonant Raman scattering on carbon layers[148]. Raman bands and vibrational modes for carbon black and graphite are discussed in detail in a study[146]. The characteristic Raman spectrum of hard carbon is shown in Figure 8.

Figure 8
Raman spectrum of hard carbon and Gaussian approximation of peaks (the Figure was taken from authors archive to Ref. [74])

Approaches assessing the degree of graphitization of a material are based on estimating the D1 : G ratio or the integral D3 : G ratio[136]. The D3 peak at 1500 cm–1 is observed for sp3 carbon atoms in amorphous states[146], so the second approach proposes to estimate the degree of graphitization by the ratio of sp2- and sp3-hybridized atoms. The first approach is the most common. The ratio is calculated from the peak heights or integral intensities signed as IG and ID (or AG and AD). The integral intensity ratio is calculated from the ratio of the areas under peaks G and D1 (IG/ID1). Gomez-Martin et al.114 proposed to estimate the degree of graphitization by the calculation method (IG/(ID1 + IG)), which allows to obtain a value in the range from 0 to 1. The values of the ratio of the integral intensities D1/G for different samples of hard carbon are close and usually greater than 1. In addition, the width of the D and G peaks at half height decreases with increasing annealing temperature of the hard carbon samples, also indicating an increase in the ordering of the materials[58, 126].

In addition to the presence of graphite-like domains, micropores were found in the microstructure of the samples (as visualized by TEM[92, 123, 128]), formed as a result of disordered stacking of nanoscale domains. A more detailed investigation of the porous structure of hard carbon is carried out using gas adsorption/desorption methods to study open pores and the SAXS method to study pores isolated from an external environment. Due to the small scattering angles, SAXS has a low resolution for single atoms and a high resolution for larger particles. The scattering coefficient of a particle is determined by the total electron density of the atoms within it, so the presence of regions of significantly different electron density in the sample (e.g., closed micropores) will be reflected in the SAXS diffraction profile[149]. According to model representations, closed pores in hard carbon are spaces between graphite-like domains or separate graphene-like layers in the bulk. An example of the SAXS curve for hard carbon is shown in Figure 3d. The part described as Q–4 on the plot of the reflection intensity vs the wave vector is associated with the reflection of X-rays from the surface of graphite-like particles[113]. The pores of hard carbon account for the appearance of convex regions on the curve (for a detailed model to interprete the SAXS curves for disordered carbon materials see[113]).

Several approaches have been proposed to determine the pore size of hard carbon[64, 131, 133, 150]. For example, Simone et al.[133] used Guinier’s dependence (1) to analyze the scattering intensity in the convex region:

\[ \begin{equation} I=N \cdot V^2 \cdot \exp \left(\frac{-Q^2 \cdot R g^2}{3}\right) \end{equation} \]
(1)

where \( N \) is a pore number, \( V \) is the pore volume and \( Rg \) is radius of gyration.

Pore diameter (\( D_0 \)) can be estimated using Eq. (2), a factor of 3/5 is introduced to account for the non-spherical oval-shaped[133] pores (see Figure 3a):

\[ \begin{equation} R g=\sqrt{\frac{3}{5}} \cdot\left(\frac{D_0}{2}\right) \end{equation} \]
(2)

Pore sizes of hard carbon derived from different precursors vary from ~0.6 to ~4.5 nm. Based on the SAXS data, it was found that the pore size increases with increasing annealing temperature[64, 128, 131, 133]. At the same time, when the annealing temperature is increased from 700 to 1300 °C, the pycnometric density of the samples of hard carbon increases from 1.727 to 2.105 g cm–3, whereas further increase of the temperature leads to a decrease of this parameter to 1.392 g cm–3 at 2000 °C (for comparison, the pycnometric density of graphite is 2.26 g cm–3). Based on the analysis of the SAXS data and the determination of the skeletal (or effective) density (which is determined by the total density of the skeleton and the closed pores) by helium pycnometry, it is assumed that this is due to an increase in the volume of closed pores with a simultaneous decrease in their number. This is explained by the formation of larger pores due to the merging of small ones[128]. It is also likely that closed pores can be formed due to the collapse of open pores as the annealing temperature of the material increases[128].

The open pore portion in hard carbon is mainly determined by the gas adsorption/desorption method[29, 35, 36, 40, 64, 82, 133, 144, 57-59, 151-155]. Nitrogen at 77.4 K is used to measure the specific surface area of the material (the ratio of the total surface area of the sample to its mass) and pore sizes larger than 0.7 nm[41, 128], CO2 is used to measure pores smaller than 0.7 nm and, to account for their contribution to the specific surface area value, measurements are made at 273 K[156]. Measurements using N2 and CO2 can be complementary[41, 72, 115]. The reason for the differences in specific surface area values for the same sample of hard carbon when different gases are adsorbed thereon is related to the different sizes of their molecules[115], and also to the fact that measurements with CO2 at higher pressures (up to 10 kPa) allow the contribution of pores smaller than 0.7 nm, which are inaccessible to N2, to be taken into account due to the better diffusion of carbon dioxide[41, 152].

There are at least 5 types of adsorption/desorption isotherms for gases that do not undergo chemisorption on the surface of the sample (Figure 9). Type I isotherms are the feature of samples with micropores and relatively small surface areas, while types II and III are typical for macroporous and non-porous objects. Type IV and V isotherms are observed for mesoporous samples[157]. There is also the type VI isotherms, obtained for non-porous materials[158]. Type IV and V isotherms may have a hysteresis loop, which is interpreted when analyzing pore characteristics, such as their shape. There are a number of theoretical approaches and models for the analysis and calculation of material parameters which are discussed in detail in studies[158-161]. For hard carbon, different isotherms are obtained depending on the precursors, primary processing methods and annealing temperatures. For example, cork-derived samples are characterized by a combination of I and II type isotherms[128]. Materials from peanut shells and phenol formaldehyde resins have a combined I/IV type adsorption isotherm[40, 63]. Type IV is characteristic of hard carbon from sodium polyacrylate[53] and some plant precursors, e.g., Ganoderma lucidum residue[118]. Treatment of starting biomasses with hydrochloric acid can change the adsorption isotherm, e.g., from type I to type II for argan samples, and from type III to type IV for typha[117] and peanut shell[121].

Figure 9
Gas adsorption/ desorption isotherms on the adsorbate (see text for explanation)[159].

The Brunauer, Emmett and Teller (BET) equation is used in most works to describe the isotherms and calculate the specific surface area of the samples[3, 35, 36, 65, 115, 117, 123, 130, 155, 162, 40-42, 57-59]. The specific surface area (SSABET) of hard carbons varies considerably depending on the precursor and can range from ~1 m2 g–1 (Ref. [133]) to ~2500 m2 g–1 (Ref. [63]) for different starting materials. The specific surface area of hard carbons derived from sugars, natural and synthetic polymers varies from several tens[35, 162] to several hundreds of m2 g–1 (Ref. [63]).

According to several studies[59, 64, 123, 126, 128], increase in carbonization temperature from 800 °C (Refs [64, 123, 128]) to 1600 °C (Refs [126, 128]), leads to a decrease in the specific surface area of the samples, which may be related to the collapse of open pores[128]. For example, for hard carbon derived from phenol formaldehyde resin, SSABET decreases from 350 to 101 m2 g–1 with increasing carbonization temperature from 800 to 1500 °C[40]. It was shown[115, 133] that at higher carbonization temperatures the specific surface area ceases to decrease significantly and can even slightly increase. In addition to the carbonization temperature, the argon flow rate over the samples during their high temperature carbonization has a significant effect on the specific surface area[3]: as the rate of Ar supply increases from 200 to 600 and 1000 cm3 min–1 respectively, a decrease in SSABET of hard carbon from sucrose is observed from 670 to 120 and 11 m2 g–1 respectively. This may be due to the faster removal of gaseous carbonization products, which can react with the carbon material, activating the surface and thus increasing the surface area of the samples. A high heating rate results in an increase in the specific surface area of hard carbon samples[35], while a low heating rate of the samples contributes to a more effective desorption of gaseous products from the surface of the material during carbonization, thus reducing pore formation.

4.4. Study of defects

The main types of defects in hard carbon that can significantly affect its electrochemical performance are dangling bonds at the edges of graphite-like domains, vacancies within graphene-like layers, heteroatoms, atomic structure defects (due to the presence of sp3 atoms in the material), etc.

The dangling bonds were studied by EPR spectroscopy[131]. Localized paramagnetic centers were attributed to dangling σ-bonds of carbon atoms at the edges of graphite-like domains and at the surface of open pores. The spectra showed a broadening of the peaks, which was explained by the presence of sp3 carbon atoms in hard carbon, leading to a broadening of the energy gap between the valence and conduction bands. Using the EPR method, it was found that localized paramagnetic centres are able to interact with atmospheric oxygen, giving rise to a difference between the EPR spectra recorded in air and in vacuo.

Datta et al.[163] calculated the energy of vacancy formation in a graphene-like layer using the DFT method and showed that the formation of divacancies, i.e. two vacancies located at the adjacent points of the lattice, is energetically the most advantageous (the energy of monovacancy and divacancy formation is the same and is 8 eV, hence the removal of an atom in the first case requires 8 and 4 eV in the second case) (Figure 10a). When divacancies are formed in the material, the atomic lattice remains virtually unchanged. According to calculations, in addition to divacancies, the formation of Stone-Wales defects is the most energetically favourable. They arise from rotations of C – C bonds by an angle of ∼90° in the hexagonal structure (Figure 10b), resulting in two pentagonal and two heptagonal rings in the structure (their formation energy is 5 eV). At the same time, the dangling bonds are unstable in graphene-like layers.

Figure 10
Schematic representation of defects in hard carbon: divacancies (a)[164] and Stone-Wales defects (b)[165].

The content of defects in the layer of sp2 atoms can also be determined by evaluating the integral areas of the peaks in the PDF plot[139].

Gomez-Martin et al.[115] proposed to estimate the degree of deformation of hard carbon β by Raman spectroscopy as 1 – α, where α is the degree of graphitization of the material. It decreases from 0.78 to 0.66 with increasing annealing temperature of hard carbon samples from 800 to 2000 °C 114 and from 0.76 to 0.65 with decreasing heating rate of samples from 5 to 0.5 deg min–1 (for annealing at 1300 °C)[35].

Defects in hard carbon can also be studied by measuring the active surface area, which is calculated from the chemisorption of oxygen. The amount of oxygen-containing complexes is determined by mass spectrometry from the amount of CO and CO2 gases released during outgassing at 950 °C[3, 166]. The active surface area depends on defects, such as vacancies and dangling bonds, as well as the presence of heteroatoms, adsorbed atoms and groups of atoms, etc. Zhang et al.[72] found that by increasing the annealing temperature from 950 to 1250 °C and decreasing the number of defects, the active surface area of hard carbon decreases from 23 to 1.1 m2 g–1, at higher annealing temperatures further reduction of the active surface area is insignificant.

The defects in the layers and heteroatoms, which determine the value of the active surface area, can influence the ICE of hard carbon and the process of formation of passivation layers due to electrolyte decomposition. It was shown that as the active surface area decreases with increasing annealing temperature, there is initially a small growth in the ICE for anodes in sodium-ion half-cells from ~50 – 60 to ~70 – 75%, and then its decrease[72]. The mechanism of these changes requires further in-depth investigation. This may be due to the influence of other parameters of the material that are also determined by the annealing temperature, such as a reduction in the volume of open pores, the number of which affects the availability of the material surface for electrolyte decomposition. Defects and heteroatoms can also affect the electrochemical capacity of the material. For example, (C=O) groups can provide a reaction pathway for the reversible intercalation of Na, K and Li ions to form (C – O Na/K/Li)[167].

5. Electrochemical performance of hard carbon

Most of applied and fundamental studies and reviews are devoted to the electrochemical properties of hard carbon in SIBs, and to a much lesser extent in PIBs. As for LIBs, the practical interest to this material in this system is not very high today, mainly due to the more attractive properties of graphite, although it is often noted that hard carbon can be operated in LIBs at higher current densities and low temperatures.

The diversity of precursors and synthesis conditions mentioned above hampers the comparison of the electrochemical properties of hard carbons. Bommier et al.[31] gave a statistical analysis of the electrochemical characteristics for different carbon materials as a function of material type (graphite, graphitizable carbon, various samples of non-graphitizable carbon, etc.), charge-discharge current density and other parameters of the cycling process and suggested that there is a limit to the electrochemical capacity that can be achieved for hard carbon in SIBs. Theoretical calculations have shown that the theoretical capacity of hard carbon is in the range of 300 – 400 mA h g–1 for SIBs[168]. In practice, capacity values of around 300 mA h g–1 and higher have been achieved for samples of hard carbon derived from various precursors such as sucrose[75, 131, 169], cellulose[36, 170], phenol formaldehyde resin[43, 135], biomass products[171]. Higher capacities (above 400 mA h g–1) have also been reported, while the materials with such electrochemical properties are characterized by a high microporosity. The electrochemical performance of various samples of hard carbon in SIBs are given in Table 1.

Table 1
\[ \]
Comparison of synthesis techniques from different precursors and electrochemical performance of hard carbon samples as anode material for SIBs. Refs. [36, 38, 39, 43, 70, 78, 82, 87, 93, 94, 135, 154, 172-174]
(1)

The main factors affecting the electrochemical performance of hard carbon-based materials and the mechanisms underlying the electrochemical reactions of this material with alkali metal cations are discussed below.

5.1. Coulombic efficiency of hard carbon

Coulombic efficiency of the first charge-discharge cycle of anode material is a critical parameter that determines the mass ratio of the main components (cathode, anode and electrolyte) in metal-ion batteries. It is governed by intensity of irreversible processes, primarily, the formation of a passivation layer (or solid electrolyte interphase, SEI), responsible for the exclusion of a part of ions from the cycling process. SEI is formed through electrochemical decomposition of the electrolyte to produce a complex mixture of organic and inorganic compounds on the surface of the anode material. The composition, thickness, conductivity and chemical stability of the passivation layer depend on the electrolyte composition. This layer protects the electrolyte and the anode material from degradation and therefore influences the cycling stability.

The low ICE significantly complicates battery production: manufacturers are forced to use an excess of cathode weight in relation to the anode weight[175]. Another method of compensating irreversible capacity is the use of the so-called ‘sacrificial’ salt, a special additive to the cathode material that replenishes the lack of alkali metal cations after the first charge[176, 177]. The carbon electrode can also be subjected to a preliminary electrochemical cycle in a half-cell with a metal counter electrode.For sodium-ion systems, the chemical ‘presodiation’ of anodes by contacting them with sodium metal or its salts[178-180] has been reported,with their subsequent utilization to assemble full cells[181]. However, the application of the last two methods in the battery production requires the addressing of serious technological and engineering challenges due to the complexity of the operations performed on an industrial scale.

Chemical ‘prepotassiation’ is also known for potassium-ion systems. Prepotassiation is achieved by electrochemical pre-cycling of anode materials in a potassium-ion half-cell and further assembly of full cells with the treated electrode. Patent[182] proposes a method of treating finished electrodes with a mixture of naphthalene and potassium dissolved in an organic solvent. However, these approaches have not been scaled-up.

The main factors affecting the Coulombic efficiency of hard carbon in metal-ion batteries are:
1) electrode composition (which typically includes binder and various conductive additives, such as carbon black);
2) electrolyte composition;
3) material properties (specific surface area, defectiveness, presence and the type of heteroatoms).

5.1.1. Coulombic efficiency of hard carbon in SIBs

For hard carbon electrodes in sodium-ion electrochemical systems, the ICE can vary over a wide range. A number of articles have reported an increase in the ICE for hard carbon electrodes in sodium-ion cells when using materials with a specific surface area of less than 10 m2 g–1[35, 183].

The influence of the anode composition in SIBs on the ICE has been studied in a number of works[135, 136, 172, 184-186]. Bommier and Ji[187] showed that the fabrication of hard carbon electrodes without the addition of acetylene black helps to avoid decomposition of the electrolyte on the extended surface of this conductive additive, resulting in anodes having higher ICEs. The formation and optimization of a conductive carbon coating on hard carbon particles have been reported, with such composites showing ICEs above 70%[185, 186]. This is attributed to the fact that a special conductive coating reduces the specific surface area of the material (from 95 to 70 m2 g–1 for the coated sample)[185]. High Coulombic efficiencies are also achieved in self-supported ‘monolithic’ electrodes prepared in the form of a polymer binder-free finished carbon film[172, 184]. The authors explain this observation by the absence of significant grinding of the components of the electrode mixture[135, 188], whereas the preparation of traditional electrode slurry may increase the specific surface of the material due to the fragmentation of its components. At the same time, Bobyleva et al.[82] showed that even after milling, hard carbon with a monolith-like morphology (a monolithic carbon rod is formed during synthesis, and requires additional milling to be powdered) maintains high Coulombic efficiencies of about 89%. Some authors have demonstrated a significant increase in ICEs (up to 99.5%) by using a graphite plate as a template for the synthesis of hard carbon[173, 174].

The composition of electrolytes, i.e., salts and solvents, determines the composition of the passivation layer formed on the surface of the hard carbon as the electrolyte decomposes.

Typical electrolytes for hard carbon anodes are solutions of salts in non-aqueous organic solvents, such as alkyl carbonates and ethers[187, 189, 190]. For sodium ion systems, the salts of choice are perchlorate (NaClO4), hexafluorophosphate (NaPF6) or sodium bis(trifluoro)sulfonylimide (NaN(CF3SO2)2, NaTFSI), sodium bis(trifluoro)sulfonylamide (NaN(SO2F)2) and others. Among the organic alkyl carbonates, the common solvents are ethylene carbonate (EC), propylene carbonate (PC) and their binary mixtures, as well as binary mixtures of EC with diethyl carbonate (DEC) and dimethyl carbonate (DMC), and other combinations. Among organic ethers, the most demanded are glymes — ethylene glycol dimethyl ethers with the general formula СН3О(С2Н4О)nСН3 (n = 1 – 4).

The electrochemical properties of hard carbon in electrolytes based on the two classes of the above solvents, namely, esters (EC, PC, DEC, DMC) and ethers (glymes), were studied in a recent works[191, 192]. The use of glymes improves the cyclic stability and increases the Coulombic efficiency, which is associated with the formation of a thinner and more stable passivation layer (referred to as ‘pseudo-SEI’). At the same time, it has been reported that the use of glymes for hard carbon is characterized by the phenomenon of ‘co-intercalation’, i.e. the intercalation of sodium ions surrounded by a solvate shell into the interlayer spacing[193].

A number of studies have reported a beneficial effect of the fluoroethylene carbonate (FEC) electrolyte additive on the resulting passivation layer, reducing the degradation of anode materials during cycling in both sodium-ion half cells and sodium-ion full cells. However, the negative impact of this additive on the electrochemical performance of hard carbon also has been noted[169, 194]. The observed negative effect may be related to the formation of a less electrically conductive passivation layer when FEC is added to the electrolyte.

One of the reasons why it is difficult to analyze and systematize data on electrolytes is the lack of a commercially available electrolyte that can be used as a benchmark. Hosaka et al.[195] noted that the differences in the quality of the salt, especially NaPF6, and differences in the purity of the solvent significantly affect the properties of the finished electrolyte. The use of NaClO4 is limited by its thermal instability, and application of NaTFSI is hampered by its ability to react with aluminium current collector[189].

In addition to the use of various additives in the electrolyte, its properties can be modified by varying the salt : solvent ratio. Thus, the use of medium (2.5 – 3 M) and high (>3 M) concentrated solutions instead of the standard 1 M salt solutions (NaPF6, NaFSA) improves the Coulombic efficiency of hard carbon anodes in SIBs, since it has been shown that increasing the concentration of the solutions contributes to the formation of a more stable passivation layer[196-198].

Among the less common salts and solvents, sodium tetraphenylborate was found to increase the Coulombiс efficiency to 95%[199]; also sodium bis(oxalato)borate in triethyl phosphate featuring high stability of the electrolyte to thermal decomposition[200, 201]. The ability to operate at high current densities (charging within 5 min) was demonstrated by a system with a solid electrolyte based on sodium carborane Na(CB9H10)0.7(CB11H12)0.3[202].

The electrolyte composition also has a significant impact on the cycling performance of hard carbon in full cells. For example, a full cell with Na3V2(PO4)2F3 cathode and hard carbon anode retains 80% of its original capacity after cycling for 12 months in an electrolyte based on 1.5 M NaPF6 in ethylene carbonate[203]. The authors noted that the use of electrolytes based on cyclic carbonates is preferable to maintain capacity during long-term cycling. In a follow-up study[204], the cycling performance was improved through the use of a number of additives in carbonate electrolytes (vinylene carbonate, 1,3-propanesultone, succinonitrile, sodium difluoro(oxalato) borate), which allowed to retain 89% of the original capacity after 60 cycles at 55 °C.

5.1.2. Coulombic efficiency of hard carbon in PIBs

Compared to the ICE of hard carbon anodes in SIBs, the Coulombic efficiency of PIBs is significantly lower.

Similar to SIBs, PIB electrolytes typically comprise organic solvents (alkyl carbonates or ethers) and potassium salts. The latter are potassium hexafluorophosphate (KPF6) or bis(trifluoro) sulfonylimide (KN(CF3SO2)2, KTFSI)[195].

The Coulombic efficiencies of negative electrodes in such electrolytes is usually 40 – 65%[205-210]. Katorova et al.[206] studied the influence of the concentration of KPF6 in diethylene glycol dimethyl ether (diglym) on the electrochemical performance of hard carbon anodes. It was shown that as the salt concentration increases from 1 to 2.5 M, the Coulombic efficiency increases from 70 to 77%. Dai et al.[207] compared two electrolyte solutions containing 1 M KPF6, the first based on dimethoxyethane and the second based on the EC – DEС mixture. In the first case, the Coulombic efficiency was 77.3%, in the second case it was less than 50%. In addition, the dimethoxyethane-based electrolyte had a higher capacity, stable long-term cycling performance and the ability to operate at high current densities. For example, in the dimethoxyethane-based electrolyte at a current density that allows charging or discharging in 3 minutes (denoted as 20C), the capacity of hard carbon was 112 mA h g-1, and in a mixture of ethylene carbonate and diethyl carbonate this value was 11.9 mA h g–1.

As for the salts for electrolytes, the Coulombic efficiency of hard carbon anodes in potassium-ion half-cells with dimethoxyethane electrolyte was 76.3% in the case of 1 M KPF6 and 44.8% in the case of 1 M KTFSI[208]. However, in alkyl carbonate solutions it is close for both salts, just over 50%. Thus, a higher Coulombic efficiency in ether-based electrolytes can be assumed as a general trend.

5.2. Capacity and operating potential of hard-carbon anodes

The capacity and operating voltage of anodes are usually estimated based on their galvanostatic measurements in half cells, where the counter electrode is an alkali metal electrode as an ‘infinite’ source of corresponding ions.

5.2.1. Capacity and operating potential of hard-carbon anodes in SIBs

A typical galvanostatic charge curve for hard carbon in a half cell with a sodium counter electrode can be divided into two distinct regions, a sloping region up to ~100 mV vs Na/Na+ and the subsequent more flat region (termed as a plateau, quasi-plateau or pseudo-plateau) from ~100 to 0 mV (Figure 11a). Capacity values for both regions are usually analyzed in terms of being affected by material properties such as interplanar distance d002, ordering, specific surface area, porosity, defectiveness and the presence and composition of heteroatoms (e.g., O, N).

Figure 11
Typical galvanostatic charge-discharge curves of hard carbon in an electrochemical half-cell vs sodium metal (a), potassium metal (b), lithium metal (c) (the Figure was created by the authors using original data from the studies[74, 82, 131]). The charge curve is indicated by the solid line, the discharge curve is indicated by the dashed line. The division of the charge and discharge curves into two colours helps to explain its division into two regions (slope and plateau) described in the text.

The interlayer distance is one of the key parameters in the study of the mechanisms of electrochemical reversible intercalation of metal ions into the microstructure of the material. For sodium-ion electrochemical systems, there is an increase in the plateau region and a decrease in the sloping region of the galvanostatic curves with a decrease in the distance d002 and an increase in the lateral dimensions of the graphite-like domains, observed when raising the annealing temperature of the samples, i.e. with a greater degree of graphitization of the material. As already mentioned, in sodium-ion half-cells the maximum capacity of hard-carbon anodes is observed for materials obtained by annealing at 1300 – 1500 °C. Increasing the annealing temperature above these values decreases the anode capacity. This can be due to an increase in the degree of graphitization and growth in the size of graphite-like domains (as mentioned above, reversible intercalation of sodium ions into graphite is difficult, unlike lithium or potassium ions).

An important task is to elucidate the relationship between the porosity and pore size of a material and its key electrochemical properties such as capacity and Coulombic efficiency. Among the main parameters of the material that determine its capacity, the size and number of closed micropores are singled out. Metal ions are considered to be able to reversibly fill micropores of hard carbon through diffusion between carbon layers. As the annealing temperature increases (with increasing pore size, as determined by SAXS) a boost in capacity is observed in sodium-ion half-cells[36, 64, 128, 133] with a concomitant increase in capacity in the plateau region of the galvanostatic charge-discharge curve[131].

For sodium-ion systems, a number of studies contained experimental data with the charge-discharge curves representing only a sloping region down to 0 V[8, 62, 63]. The tested samples of carbonaceous materials featured high specific surface area (above 1430 m2 g–1)[63]. In this case, it was assumed that sodium cations interact with defects and heteroatoms on the material surface rather than with its internal volume. The charge curve in such cases is similar to that of graphitizable carbon, for which the intercalation mechanism of the interaction of sodium ions with the material is being questioned by some authors[31, 186]. However, it should be noted that the possibility of graphitization of the material was rarely tested in the works, so it is not always possible to clearly understand which type of carbon, graphitizable or non-graphitizable, was studied by the authors.

Non-graphitizable carbonaceous materials have been shown to be capable of long-term cycling stability. A number of papers have reported retention of more than 93% of capacity for 1000 charge-discharge cycles at a current density of 0.2 A g–1 (see [209]) and 73% of initial capacity for 10 000 cycles at 2 A g–1 (see [210]).

5.2.2. Capacity and operating potential of hard-carbon anodes in PIBs

In potassium-ion systems, two regions are also distinguished on the galvanostatic charge-discharge curves, a sloping region above 0.4 V vs K/K+ and a plateau or quasi-plateau region below these values[120, 136, 167]. Other authors, e.g., Kubota et al.[131] identified three sloping regions (Figure 12) in the ranges of 0.002 – 0.2, 0.2 – 0.7, 0.7 – 2.0 V vs K+/K.

Figure 12
Galvanostatic curves of the hard carbon samples carbonized at different temperatures[131].

The influence of the synthesis conditions and the microstructure of hard carbon on the capacity ratio for these regions has been explored in a number of papers. For example, it was shown that raising the annealing temperature during the synthesis of hard carbon from 1100 to 1500 °C increases the specific capacity of the anode material in all regions of the charge-discharge curve[36, 43]. Kubota et al.[131] found that the specific capacity of anodes based on hard carbon samples carbonized at temperatures between 700 and 2000 °C does not change significantly over the whole range of potentials, but in the same time, there are changes in capacities of different regions of the galvanostatic curve. Thus, the capacity in the range of 0.7 – 2.0 V decreased drastically and the capacity in the range of 0.2 – 0.7 V increased. The capacity in the potential range of 0.002 – 0.2 V is significantly lower than in other potential ranges, and reaches its maximum of 47 mA h g-1 for the material carbonized at 1800 °C. The authors believe that the capacity of the samples carbonized at lower temperatures (700 – 900 °C) is mainly associated with the intercalation of potassium cations in the space between the defective graphene-like layers with a large d002 distance.

In a number of works devoted to PIBs, materials derived from different biomasses were studied[121, 144, 211]. For example, the capacity of ahard carbon anode material derived from corn husk was about 230 mA h g–1, while after 100 cycles it retained 89.1% of the original capacity and was ~205 mA h g–1[121]. The special feature of various biomasses is the presence of naturally occurring heteroatoms in the hard carbon; and such dopants have been specifically explored[109, 112, 167, 212]. The analysis of the publications devoted to carbon doping does not allow to speak about a boost in capacity or the ICE for such anode materials. Moreover, virtually no quasi-plateau region was observed in galvanostatic curves at low potentials[109, 212]. The predominance of pseudo-capacity interactions between hard carbon and potassium cations observed in these works is associated with an increase in the number of defects and interlayer space in hard carbon. On the other hand, it was noted[109, 112, 167, 212] that the hard carbon doping with heteroatoms improves the capacity maintenance during long-term cycling and an increase in the anode capacity at high current densities.

5.2.3. Capacity and operating potential of hard-carbon anodes in LIBs

Due to the larger interlayer distance compared to graphite, faster diffusion of lithium ions is expected in hard carbon, leading to better operation at high current densities[213]. In addition, the microstructure of hard carbon allows additional storage of lithium ions in micropores below 0 V vs Li/Li+ (see Figure 11c)[213-215]. Therefore, some studies consider hard carbon as a safer and higher capacity alternative to graphite for the new generation of lithium-ion batteries[216-219]. For example, the prototype of LIBs based on the LiNi0.6Mn0.2Co0.2O2 cathode and the hard carbon anode showed good cycling stability at current densities providing charge or discharge for 20 min (designated as 3C)[119]. At the same time, in the case of SIBs and PIBs, cycling stability is usually studied at current densities that provide charge or discharge in 1 h (designated as 1C). It was shown that increasing the current density above 1C in the SIB prototypes promotes the deposition of sodium metal[220].

In addition to studying the relationship between the ratio of sloping and plateau capacities with synthesis conditions and material properties, the elucidation of mechanisms of the electrochemical processes occurring during the charge-discharge cycle remains relevant. Given the complexity of the subject under study, these mechanisms are still the matter of intense debate.

5.3. Interaction models for hard carbon and alkali metal ions in electrochemical cells

Based on the available experimental data, the main reaction pathways for the interaction with alkali metal cations can be distinguished:

1) interaction with defects and heteroatoms;
2) intercalation of ions into the interlayer space;
3) filling of open and closed micropores by metal ions, where the formation of metal clusters can be observed.

5.3.1. Interaction models for SIBs

Stevens and Dahn[221] were the first to study hard carbon as an anode material for SIBs in 2000. The D-glucose-derived material demonstrated a capacity of 300 mA h g–1. The authors proposed a model of a two-step mechanism by which sodium cations intercalate into the microstructure of hard carbon. According to the authors, two processes occur during charge and discharge, each corresponding to its own part in the galvanostatic curve[218]. At the sloping region (above 100 mV), a shift of the peak detected by the in situ wide angle X-ray scattering. It was suggested that intercalation of metal ions into the interlayer space occurs at this site. According to the authors’ model, after intercalation, in the range from 100 to 0 mV vs Na/Na+, sodium metal deposites in the pores of the carbonaceous material. This model has been called ‘house of cards’ or ‘intercalation – filling’ (see Figure 13a); it is the model adopted by most authors exploring the electrochemical properties of hard carbon (see, e.g., Ref. [15]). A similar mechanism of interaction between hard carbon and alkali metal cations has been proposed by another scientific group[222]. In this study, hard carbon electrodes were charged to different degrees (a number of ‘points’ in the galvanostatic curve) in a half-cell vs sodium metal and studied by PXRD, SAXS and Raman spectroscopy. The combination of these methods showed that in the sloping region, the change in the position of the G line in the Raman spectra together with the shift of the peak (002) towards lower angles is associated with the intercalation of sodium ions into the interlayer space of hard carbon. The processes occurring in the plateau region were interpreted from the SAXS measurements, i.e. the drop in intensity in the region of 0.03 – 0.07 Å–1 was ascribed to the filling of the hard carbon nanopores with sodium ions.

Figure 13
Models of two-stage mechanism of electrochemical interaction of sodium ions with hard carbon anode: ‘intercalation–filling’ (a); ‘adsorption–intercalation’ (b); ‘adsorption–filling’ (c). The Figure was created by the authors using original data from the studies[74, 222-227].

With the increased attention paid to SIBs, studies of the mechanisms of interaction of hard carbon with alkali metal cations have recently received a new impetus. The ‘intercalation-filling’ mechanism has received a number of additions and critical remarks, in particular because of the curve region below 100 mV, which is mainly associated with the deposition of sodium metal clusters inside micropores. This issue was studied by 23Na NMR spectroscopy. In the studies[223, 224], the formation of metal clusters was not observed, although it was found that sodium ions occupy at least two different energy positions.

In 2012, Cao et al.[225] proposed a new model of a two-step ‘adsorption – intercalation’ mechanism (Figure 13b). According to the authors, based on cyclic voltammetry data, the sloping voltage curve corresponds to the reaction of sodium cations with the surface of graphite-like domains. At lower voltage, sodium ions are introduced into the interlayer space of hard carbon due to the large interlayer distance.

Subsequently, the ‘adsorption – intercalation’ model was confirmed by ex situ PXRD[226], which showed that the shift of the peak corresponding to (002) reflection of the graphite structure occurs in the range of potentials below 0.2 V, while the interlayer distance increases from 3.96 to 4.16 Å for the charge in this region. Bommier et al.[139] supplemented this model by an assumption of the third stage of the mechanism of interaction of sodium ions with hard carbon at potentials close to 0 V vs Na/Na+. The authors found that in the plateau region, the diffusion coefficient of sodium ions reached a minimum at a potential of 0.05 V, and its significant increase was observed with decreasing potential. Based on these findings, it was concluded that the deposition of sodium metal occurs in the range of 0.05 – 0.02 V, which corresponds to the third, more sloping voltage region.

Using on the above models, Zhang et al.[74] classified hard carbon materials into three groups according to the character of the galvanostatic curve in SIBs and the physical properties of the materials, which in turn depend on the annealing temperature. In the materials of the first group, obtained at temperatures below 950 °C, the energy storage mechanism is significantly influenced by the presence of oxygen- and nitrogen-containing functional groups, as well as by the strong defectiveness of the material. Sodium ions can interact both reversibly and irreversibly with functional groups of hard carbon obtained at temperatures below 950 °C. In materials of the second group, prepared at annealing temperatures of 950 to 1550 °C, the galvanostatic curve in sodium-ion systems is divided into two regions. According to the authors, the mechanism of storage, corresponding to the sloping region, is related to the interaction of sodium ions with defect sites presenting in graphite-like domains. Charge storage in the plateau region occurs via filling the micropores of the material with sodium ions. The third group of materials, synthesized at temperatures between 1550 and 2200 °C, do not show the plateau region. At the same time, according to the in situ XRD data, no shifts of the (002) reflection were detected, and therefore, the authors questioned the idea of intercalation of sodium ions into the interlayer spacing of the graphite-like domains. Based on these data, the third model of the two-step ‘adsorption – filling’ mechanism was proposed (Figure 13c). Li et al.[126] argued that this model is confirmed by the TEM results for the hard carbon material charged to 0 V. The authors have not found a change in the distance between graphene-like layers, but found a peak of sodium metal.

The fact that the sloping region of the charge-discharge curve is related to the processes of chemo- or adsorption of sodium ions onto the surface of the carbon particles, is supported by the disappearance of the plateau region and the retention of only the sloping region of the galvanostatic curve with an increase in the specific surface area and the formation of a large number of micropores open to the electrolyte[75, 228].

Samples of hard carbon doped with boron, phosphorus and sulfur atoms also demonstrated the change in electrochemical performance (elongation of different charge/discharge voltage regions)[103]. The authors note that doping hard carbon with these heteroatoms increases the storage capacity of the sloping region. At the same time, the highest capacity is observed for B-doped carbon. Such an influence of heteroatoms on the capacity is due to the fact that their presence can increase the graphite interlayerdistance (in the case of S and P) and create more defects in the microstructure (N-, O- and B-doping)[229].

The electrochemical processes in the plateau region at the voltage range from 0.1 to 0 V in sodium-ion systems are also much discussed. It is the most important for SIBs because its length determines the energy storage capacity of the battery. The nature of the processes taking place at the anode at potentials close to 0 V has been studied by various methods, including NMR spectroscopy.

Grey’s group 140 confirmed the hypothesis of metal cluster formation using operando 23Na NMR measurements. Two sodium signals, ‘ionic’ at –40 ppm and ‘metallic’ at 760 ppm, were detected in the NMR spectra of the electrode in the charged state. During anode charging, the signal from the sodium involved in the charging process is shifted towards metallic sodium. The authors also hypothesized a high activity of charged electrodes in air, which could explain why the signal of sodium metal was not detected in previous ex situ NMR studies[230]. The importance of carrying out operando experiment was pointed out: in the ex situ experiment, i.e. after disassembly of the electrochemical cell, the hard carbon electrode degrades due to the high reactivity of the newly formed clusters. Morita et al.[150, 231] used ex situ NMR to find out a correlation between the parameters of material synthesis and the nature of the Na signal in fully charged electrodes and concluded that metal clusters detected by NMR are formed in materials with large micropore radii of up to 1.95 nm, as determined from the SAXS curves, in contrast to materials with small pores (radius is 1.42 nm). The authors found that the pore size estimated from the SAXS data did not exhibit an evident correlation with the signal position of metal sodium clusters, which was determined by NMR. However, hard carbon samples obtained at higher carbonization temperatures (above 1600 °C) show a stronger signal shift towards metallic sodium. For the fully charged electrode from the material prepared at 2000 °C, the sodium signal is almost identical to that of metallic sodium.

Alvin et al.[136] proposed that the interaction of sodium ions with surface defects and functional groups is associated with a sloping voltage region up to 0.2 V, and the second part of the sloping region from 0.2 to 0.1 V appears due to adsorption of cations on the surface of the graphene-like layer. The intercalation of sodium ions into the interlayer spacing corresponds to the plateau voltage region. The authors also identified the fourth stage of the interaction process, which involves secondary filling of hard-to-reach closed pores at potentials close to 0 V.

The interaction of sodium ions with hard carbon has also been studied using small- and wide-angle X-ray scattering (SAXS and WAXS). In one of their first papers, Stevens and Dahn[218] used in situ SAXS measurements. In 2019, a group led by Yamada[232] used detailed analysis of ex situ WAXS data acquired during plateau charging of hard carbons with different degrees of ordering to reveal the appearance of a new peak (at Q ≈ 2.0 – 2.1 Å–1, with d ≈ 3.7 Å) close to the peak corresponding to the reflection of the 002 structure of graphite (at 2θ = 23 – 25°, Q ≈ 1.6 – 1.8 Å–1). Based on theoretical calculations, models of metallic sodium clusters have been proposed for this new peak. Another research group 183 attributes the appearance of a new peak to the formation of sodium intercalates. In addition to the XRD measurements, the formation of metal clusters in the plateau region was investigated by Raman spectroscopy[233-235]. It was shown that during charging of hard carbon, the G-band signal is not shifted in the plateau region, as would be expected in the case of intercalation. At the same time, a shift from 1600 to 1560 cm–1 is observed in the sloping region above 100 mV[234] as indicating the probable formation of intercalates.

It should be noted that the interpretation of PXRD and Raman spectroscopy data for disordered materials is somewhat difficult. For example, the peak corresponding to the (002) reflection of the graphite structure in hard carbon is significantly broadened, and the typical Raman spectrum of amorphous carbon show other bands in addition to the D1 and G bands[236].

The charge-discharge process for hard carbon anodes in sodium-ion cells has also been explored by PDF[115, 131, 139, 140]. Gomez-Martin et al.[115] found that the capacity of the sloping region of the electrochemical charge-discharge curve decreases monotonically with increasing annealing temperature, which is in line with a decrease in defect concentration and an increase in the size of graphite-like domains. In turn, the plateau region capacity reaches its maximum at the annealing temperature of 1400 °С. The maximum interplanar distance in graphite-like domains determined by PDF corresponds to the same temperature. The correlation of the interplanar distances and defect concentrations with the capacity of each section of the galvanostatic curve led the authors to the conclusion that the interaction followed the ‘adsorption-intercalation’ mechanism. On the other hand, Kubota et al.[131] considered the mechanism of interaction of sodium ions with hard carbon according to the ‘intercalation-adsorption’ model, and they also found that as the annealing temperature is increased up to 2000 °C, the capacity continues to rise monotonically in the plateau region. Interestingly, this trend was not observed when interacting with lithium and potassium ions. For these systems, the authors proposed a three-step model, in which metal ions first interact with defects on the surface of graphite-like domains, followed by intercalation into the interlayer spacing, and near 0 V begin to fill micropores to form metal clusters.

Bobyleva et al.[217] confirmed a three-stage model similar to that presented in a study[139] by exploring the pseudocapacitive (surface-controlled) properties of materials with different textural characteristics using linear voltammetry. The authors concluded that the sloping voltage region corresponds to pseudocapacitive processes and the plateau region corresponds to intercalation. It was also suggested that there was a third stage corresponding to the filling of closed pores of hard carbon with sodium ions.

In 2020, an extended version of the mechanism of interaction of sodium ions with hard carbon was presented. Based on the ex situ 23Na NMR measurements, small-angle neutron scattering and PDF in combination with theoretical calculations, it suggests that the sloping region corresponds to several processes at once, including not only interactions with the surface and defects (all types of the above defects), but also intercalation into the interlayer spacing[87].

An important publication in the study of the mechanism of sodium ion insertion was the result of the collaborative survey of the research groups of Allan and Grey[142], who used a combination of operando 23Na NMR and operando PDF and showed that the size of the metal clusters does not exceed 13 – 15 Å regardless of the pores radii in the material. The authors also noted that the intercalation of sodium ions in the interlayer spacing can occur due to the presence of five- and seven-membered rings in the microstructure of hard carbon.

Surta et al.[237] modelled the microstructure of hard carbon samples using PDF results from neutron scattering. The authors put forward a fundamentally new hypothesis that sodium ions interact mainly with specific defects in the microstructure of hard carbon. One of these defects is thought to be the most curved regions of the graphene-like layer.

5.3.2. Interaction models for PIBs

Studying the interaction between potassium ions and hard carbon by cyclic voltammetry (CV), galvanostatic intermittent titration technique, ex situ and in situ Raman spectroscopy[212, 238, 239], resulted in a conclusion about different interaction ways between potassium cations and hard carbon in the sloping and plateau regions. The sloping region was attributed to a pseudo-capacitive processes caused by the interaction of K+ ions with defects, including heteroatoms. The plateau region was assigned to intercalation. Based on the calculation of the density of states, Alvin et al.[238] proposed a model for the intercalation of K+ ions between carbon layers, according to which potassium cations tend to form ordered structures. In this case, the potassium atoms are located directly under the carbon atoms in the intercalates. Such intercalates are unstable. Kubota et al.[131] also assumes the formation of ordered intercalates of graphite with potassium cations.

Chen et al.[120] suggested a slightly different mechanism of electrochemical interaction of K+ ions with hard carbon. It is based on an in-depth analysis of the CV measurements, the shift of the X-ray diffraction peak at 2θ ~ 23° towards smaller angles during anode charging and back again during discharging, the reversible appearance of the K 2p peak in XPS spectra and the simultaneous reversible decrease of the carbon C1s peak for charged/discharged anode, the analysis of data from galvanostatic intermittent titration, in which it was found that the diffusion coefficient of potassium cations is significantly higher in the potential range > 0.4 V (in a study[238], the limit of the 0.3 V potential range is indicated as 0.3 V). Taken together, these data let the authors to conclude that the sloping region corresponds mainly to pseudo-capacitive processes (associated with sorption) and the plateau region corresponds to diffusion. However, according to the authors, these processes cannot be completely separated. Firstly, adsorption takes place at surface active sites and in the hard carbon pores, and at the same time, a some intercalation of potassium cations begins (sloping voltage region). After the hard carbon pores are filled with potassium, the potential drops to ~0.4 V and the previously adsorbed potassium cations intercalate between the layers of the hard carbon (plateau region) with simultaneous absorption of new potassium cations on the vacated pores.

A new approach to modelling the mechanisms of electrochemical interaction of K+ ions with hard carbon was proposed based on CV measurements, galvanostatic intermittent titration, ex situ X-ray diffraction data and in situ Raman spectroscopy[240-242]. The authors relate the mechanism of interaction to the carbonaceous microstructure and offer several models depending on the annealing temperature of hard carbon. It is noteworthy that this approach has much in common with the reported mechanism of interaction of hard carbon with sodium ions[87]. The amorphous component of hard carbon[240] stipulates adsorption of potassium cations, while graphite-like domains cause intercalation. Thus, for materials with a high proportion of amorphous component annealed at low temperatures of 800 – 1000 °C, the mechanism of interaction with K+ ions is adsorptive. In this case only a sloping region is observed on the charge-discharge curves. For the materials annealed at T = 1000 – 1500 °C, the galvanostatic curve shows a sloping region and a plateau region at a potential < 0.4 V, leading the authors to suggest an adsorption–intercalation mechanism of interaction of K+ ions with hard carbon. For the most ordered samples annealed at T = 1800 – 2900 °C, intercalation processes contribute most to the interaction with K+ ions. In this case, the charge-discharge curves reveal longer low-voltage plateau.

Lin et al.[242] suggested the adsorption mechanism of interaction with K+ ions both for low-temperature samples and for samples annealed in the temperature range of 1000 – 2000 °C. In the first case, the adsorption takes place on the defects of disordered structures and heteroatoms and moieties containing them, and in the second case, the adsorption sites are ordered carbon layers forming domains. The intercalation mechanism of interaction of K+ ions with hard carbon is proposed for materials annealed at temperatures above 2000 °C.

In all these studies, the of K+ ion storage mechanisms in hard carbon are determined by adsorption and intercalation processes. In case of PIBs, energy storage as a result of filling the pores of the material with metal cations in the case, in contrast to SIBs, is a controversial issue[240, 243], since to date there are no data unequivocally confirming this mechanism. Li et al.[241] suggest that nanopores of about 0.5 nm may act as adsorption sites for potassium ions, which have a size of 0.38 nm in the solvated state. Huang et al.[244] proposed a qualitatively different mechanism of potassium cation storage in hard carbon. The authors associate the sloping region with potassium adsorption on heteroatoms and defects, and the plateau region with pore filling. Using the CO2 adsorption/desorption method, the average pore size is estimated to be 0.5 nm. This value is smaller than the interplanar spacing of graphite with potassium cations intercalated therein (0.53 nm). Accordingly, the authors suggest that potassium cations fill the pores to form ordered structures similar to the potassium intercalation in graphite.

Further development of models for the potassium, and possibly sodium, cationstorage in hard carbon, can be based on the ability of these cations to form adducts with molecules of aromatic hydrocarbons containing curved planes due to the presence of five-membered rings. This ability of potassium cations to interact with corannulene has been reported[245-247].

Therefore, the storage mechanism of alkali metal cations in hard carbon is multi-stage and depends largely on the hard carbon microstructure. The microstructure of hard carbon is in turn determined by the method of preparation, i.e., annealing temperature, method of pre- and post-treatment, type of precursor. The most important parameters of the microstructure include the average interlayer distance, micropores radius, defects, presence of heteroatoms. In the above publications, the authors demonstrated the ability of alkali metal ions to penetrate into the interlayer spacing of hard carbon, interact with various defects and heteroatoms, and fill micropores to form metal clusters. The most contradictory point is the essential difference between the proposed models of K+ cation storage in hard carbon. However, this contradiction can be explained by the fact that the term ‘hard’ covers a variety of materials with different microstructures. Nevertheless, the determination of the exact nature of the relationship between the parameters of the material and its electrochemical properties remains an urgent task that requires the explanation of a series of experimental data.

6. Conclusion

In this review, the main research findings of the properties of hard carbon obtained under different synthesis conditions and the electrochemical performance of hard carbon anode for metal-ion batteries are discussed in detail.

The hard carbon microstructure implies several different reaction ways for interaction with alkali metal ions. Establishing the dependence of material parameters (presence of defects, interlayer distances, pore size, etc.) on synthesis conditions (nature of precursors, methods of their processing, annealing temperature, etc.) as well as correlations of the mechanism of interaction between hard carbon and cations of different metals with the physicochemical properties of the material remain relevant.

Currently, for hard carbon in SIBs, the average specific capacity is about 300 mA h g–1 and the ICE is ~90%, while in PIBs, the average specific capacity is about 250 mA h g–1 and the ICE is ~50%.

Graphite is still a benchmark carbon material used as an anode material for metal-ion batteries (since in the most developed technology of LIB production, its specific capacity is about 370 mA h g–1, ICE is over 95%, cycling stability is high). Developers of other carbon anodes suitable for SIBs and PIBs are targeting such electrochemical properties. The progress of recent years and the fundamental properties (reversible capacity, cycling stability, Coulombic efficiency) of hard carbon suggest that similar electrochemical performance can be achieved for it in SIBs in the near future. The development of technology for the production of hard carbon and anodes based thereon for PIBs will probably require more significant efforts.

In addition, the prospect for further increasing the capacity of anode materials for sodium and potassium ion batteries is the design of composites containing antimony, tin, lead and bismuth (similar to the graphite/silicon pair), as well as the creation of materials containing a deposited alkali metal[248]. The use of carbon black as a material for negative electrodes in SIBs and LIBs is also reported[97].

Although hard carbon is the most promising anode material for sodium-ion and potassium-ion batteries, there is currently no consensus in the scientific community on the optimal methods for producing it, reliable methods of studying it, or the mechanisms of its operation. Over the last few decades, there have been significant breakthroughs in the study and understanding of the peculiarities of this material, but so far the data, observations and regularities obtained using different methods on different samples often contradict each other. It is obvious that further development of the SIBs and PIBs will require a more thorough, detailed and methodical study of hard carbon.

7. List of abbreviations and designations

AFM — atomic force microscopy;
BET — Brunauer – Emmet – Teller equation;
CV — cyclic voltammetry;
DEC — diethyl carbonate;
DMC — dimethyl carbonate;
EC — ethylene carbonate;
EDS (EDX) — X-ray energy dispersive spectroscopy;
EPR — electron paramagnetic resonance;
FEC — fluoroethylene carbonate;
ICE — initial Coulombic efficiency;
LIB — lithium-ion battery;
PAN — polyacrylonitrile;
PANI — polyaniline;
PC — propylene carbonate;
PDF — pair distribution function;
PET — polyethylene terephthalate;
PFA — perfluoroalkoxyalkane;
PIB — potassium-ion battery;
PP — polypyrrol;
PTFE — polytetrafluoroethylene;
RS — Raman spectroscopy;
SAXS — small-angle X-ray scattering;
SEI — solid electrolyte interphase;
SEM — scanning electron microscopy;
SIB — sodium-ion battery;
SSABET — BET specific surface area;
SТМ — scanning tunneling microscopy;
TEM — transmission electron microscopy;
XPS — X-ray photoelectron spectroscopy;
XRD — X-ray diffraction.

Acknowledgements

This review was financially supported by the Russian Science Foundation (Project No. 17-73-30006). The authors are grateful to G.P. Lakienko and M.A. Novikov for their help in preparing illustrations.

References

1.
Electrochemical Energy Storage for Green Grid
Yang Z., Zhang J., Kintner-Meyer M.C., Lu X., Choi D., Lemmon J.P., Liu J.
Chemical Reviews, American Chemical Society (ACS), 2011
2.
Nickel as a key element in the future energy
Savina A.A., Boev A.O., Orlova E.D., Morozov A.V., Abakumov A.M.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2023
3.
Review—Hard Carbon Negative Electrode Materials for Sodium-Ion Batteries
Irisarri E., Ponrouch A., Palacin M.R.
Journal of the Electrochemical Society, The Electrochemical Society, 2015
4.
Review—Reference Electrodes in Li-Ion and Next Generation Batteries: Correct Potential Assessment, Applications and Practices
Cengiz E.C., Rizell J., Sadd M., Matic A., Mozhzhukhina N.
Journal of the Electrochemical Society, The Electrochemical Society, 2021
5.
Approaching high-performance potassium-ion batteries via advanced design strategies and engineering
Zhang W., Liu Y., Guo Z.
Science advances, American Association for the Advancement of Science (AAAS), 2019
6.
Recent advances in nanostructured carbon for sodium-ion batteries
Zhang H., Huang Y., Ming H., Cao G., Zhang W., Ming J., Chen R.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2020
7.
A sandwich-like porous hard carbon/graphene hybrid derived from rapeseed shuck for high-performance lithium-ion batteries
Li R., Huang J., Ren J., Cao L., Li J., Li W., Lu G., Yu A.
Journal of Alloys and Compounds, Elsevier, 2020
9.
Sodium-ion battery anodes: Status and future trends
Zhang W., Zhang F., Ming F., Alshareef H.N.
EnergyChem, Elsevier, 2019
10.
NaNbV(PO4)3: Multielectron NASICON-Type Anode Material for Na-Ion Batteries with Excellent Rate Capability
Khasanova N.R., Panin R.V., Cherkashchenko I.R., Zakharkin M.V., Novichkov D.A., Antipov E.V.
ACS applied materials & interfaces, American Chemical Society (ACS), 2023
11.
Towards K-Ion and Na-Ion Batteries as “Beyond Li-Ion”
Kubota K., Dahbi M., Hosaka T., Kumakura S., Komaba S.
Chemical Record, Wiley, 2018
12.
Tailoring sodium intercalation in graphite for high energy and power sodium ion batteries
Xu Z., Yoon G., Park K., Park H., Tamwattana O., Joo Kim S., Seong W.M., Kang K.
Nature Communications, Springer Nature, 2019
13.
Structure and function of hard carbon negative electrodes for sodium-ion batteries
Mittal U., Djuandhi L., Sharma N., Andersen H.L.
Journal of Physics Energy, IOP Publishing, 2022
14.
Hard‐Carbon Anodes for Sodium‐Ion Batteries: Recent Status and Challenging Perspectives
Shao W., Shi H., Jian X., Wu Z., Hu F.
Advanced Energy and Sustainability Research, Wiley, 2022
15.
Hard carbons for sodium-ion batteries: Structure, analysis, sustainability, and electrochemistry
Dou X., Hasa I., Saurel D., Vaalma C., Wu L., Buchholz D., Bresser D., Komaba S., Passerini S.
Materials Today, Elsevier, 2019
19.
The Structure of Glassy Carbon
Noda T., Inagaki M.
Bulletin of the Chemical Society of Japan, The Chemical Society of Japan, 1964
20.
Glass-like carbons
Noda T., Inagaki M., Yamada S.
Journal of Non-Crystalline Solids, Elsevier, 1969
21.
Structure of Polymeric Carbon
Mel'nichenko V.M., Sladkov A.M., Nikulin Y.N.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 1982
22.
Crystallite growth in graphitizing and non-graphitizing carbons
Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, The Royal Society, 1951
23.
THE INTERPRETATION OF DIFFUSE X-RAY DIAGRAMS OF CARBON
Franklin R.E.
Acta Crystallographica, International Union of Crystallography (IUCr), 1950
25.
Formation and structure of polymeric carbons
Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, The Royal Society, 1972
26.
Structure of non-graphitising carbons
Harris P.J.
International Materials Reviews, SAGE, 1997
27.
Fullerene-related structure of commercial glassy carbons
Harris † P.J.
The Philosophical Magazine, Taylor & Francis, 2004
28.
New Perspectives on the Structure of Graphitic Carbons
Harris P.J.
Critical Reviews in Solid State and Materials Sciences, Taylor & Francis, 2005
29.
Structure and Pore Size Distribution in Nanoporous Carbon
Wang Y., Fan Z., Qian P., Ala-Nissila T., Caro M.A.
Chemistry of Materials, American Chemical Society (ACS), 2022
30.
Towards an atomistic understanding of disordered carbon electrode materials
Deringer V.L., Merlet C., Hu Y., Lee T.H., Kattirtzi J.A., Pecher O., Csányi G., Elliott S.R., Grey C.P.
Chemical Communications, Royal Society of Chemistry (RSC), 2018
32.
Optimized hard carbon derived from starch for rechargeable seawater batteries
Kim Y., Kim J., Vaalma C., Bae G.H., Kim G., Passerini S., Kim Y.
Carbon, Elsevier, 2018
33.
Research Update: Hard carbon with closed pores from pectin-free apple pomace waste for Na-ion batteries
Dou X., Geng C., Buchholz D., Passerini S.
APL Materials, American Institute of Physics (AIP), 2018
34.
Defective Hard Carbon Anode for Na-Ion Batteries
Li Z., Chen Y., Jian Z., Jiang H., Razink J.J., Stickle W.F., Neuefeind J.C., Ji X.
Chemistry of Materials, American Chemical Society (ACS), 2018
35.
Low‐Defect and Low‐Porosity Hard Carbon with High Coulombic Efficiency and High Capacity for Practical Sodium Ion Battery Anode
Xiao L., Lu H., Fang Y., Sushko M.L., Cao Y., Ai X., Yang H., Liu J.
Advanced Energy Materials, Wiley, 2018
36.
Synthesizing higher-capacity hard-carbons from cellulose for Na- and K-ion batteries
Yamamoto H., Muratsubaki S., Kubota K., Fukunishi M., Watanabe H., Kim J., Komaba S.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2018
37.
D-Glucose Derived Nanospheric Hard Carbon Electrodes for Room-Temperature Sodium-Ion Batteries
Väli R., Jänes A., Thomberg T., Lust E.
Journal of the Electrochemical Society, The Electrochemical Society, 2016
38.
Pre-Oxidation-Tuned Microstructures of Carbon Anodes Derived from Pitch for Enhancing Na Storage Performance
Lu Y., Zhao C., Qi X., Qi Y., Li H., Huang X., Chen L., Hu Y.
Advanced Energy Materials, Wiley, 2018
40.
Properties and sodium insertion behavior of Phenolic Resin-based hard carbon microspheres obtained by a hydrothermal method
41.
Properties and lithium insertion behavior of hard carbons produced by pyrolysis of various polymers at 1000°C
Piotrowska A., Kierzek K., Rutkowski P., Machnikowski J.
Journal of Analytical and Applied Pyrolysis, Elsevier, 2013
42.
Pore Structure of Hard Carbon Made from Phenolic Resin Studied by129Xe NMR
Gotoh K., Ueda T., Eguchi T., Kawabata K., Yamamoto K., Murakami Y., Hayakawa S., Ishida H.
Bulletin of the Chemical Society of Japan, The Chemical Society of Japan, 2009
43.
High-Capacity Hard Carbon Synthesized from Macroporous Phenolic Resin for Sodium-Ion and Potassium-Ion Battery
Kamiyama A., Kubota K., Nakano T., Fujimura S., Shiraishi S., Tsukada H., Komaba S.
ACS Applied Energy Materials, American Chemical Society (ACS), 2019
44.
Valorizing low cost and renewable lignin as hard carbon for Na-ion batteries: Impact of lignin grade
Matei Ghimbeu C., Zhang B., Martinez de Yuso A., Réty B., Tarascon J.
Carbon, Elsevier, 2019
45.
Biowaste Lignin-Based Carbonaceous Materials as Anodes for Na-Ion Batteries
Marino C., Cabanero J., Povia M., Villevieille C.
Journal of the Electrochemical Society, The Electrochemical Society, 2018
47.
Hard carbon anode materials for sodium-ion batteries
El Moctar I., Ni Q., Bai Y., Wu F., Wu C.
Functional Materials Letters, World Scientific, 2018
48.
Hard carbons derived from pine nut shells as anode materials for Na-ion batteries*
Guo H., Sun K., Lu Y., Wang H., Ma X., Li Z., Hu Y., Chen D.
Chinese Physics B, IOP Publishing, 2019
49.
Microtubular Hard Carbon Derived From Willow Catkins as an Anode Material With Enhanced Performance for Sodium-Ion Batteries
Teng Y., Mo M., Li Y.
Journal of Electrochemical Energy Conversion and Storage, ASME, 2018
50.
Hard carbon micro-nano tubes derived from kapok fiber as anode materials for sodium-ion batteries and the sodium-ion storage mechanism
Yu Z., Lyu Y., Wang Y., Xu S., Cheng H., Mu X., Chu J., Chen R., Liu Y., Guo B.
Chemical Communications, Royal Society of Chemistry (RSC), 2020
51.
Hard carbons derived from waste tea bag powder as anodes for sodium ion battery
Arie A.A., Tekin B., Demir E., Demir-Cakan R.
Materials Technology, Taylor & Francis, 2019
52.
High capacity hard carbon derived from lotus stem as anode for sodium ion batteries
Zhang N., Liu Q., Chen W., Wan M., Li X., Wang L., Xue L., Zhang W.
Journal of Power Sources, Elsevier, 2018
53.
Hard carbons derived from green phenolic resins for Na-ion batteries
Beda A., Taberna P., Simon P., Matei Ghimbeu C.
Carbon, Elsevier, 2018
54.
N-doped porous hard-carbon derived from recycled separators for efficient lithium-ion and sodium-ion batteries
Wang Y., Li Y., Mao S.S., Ye D., Liu W., Guo R., Feng Z., Kong J., Xie J.
Sustainable Energy and Fuels, Royal Society of Chemistry (RSC), 2019
55.
Hard carbon derived from rice husk as low cost negative electrodes in Na-ion batteries
Rybarczyk M.K., Li Y., Qiao M., Hu Y., Titirici M., Lieder M.
Journal of Energy Chemistry, Elsevier, 2019
56.
Honeycomb-like Hard Carbon Derived from Pine Pollen as High-Performance Anode Material for Sodium-Ion Batteries
Zhang Y., Li X., Dong P., Wu G., Xiao J., Zeng X., Zhang Y., Sun X.
ACS applied materials & interfaces, American Chemical Society (ACS), 2018
58.
Hard carbon derived from corn straw piths as anode materials for sodium ion batteries
Zhu Y., Gu H., Chen Y., Yang D., Wei J., Zhou Z.
Ionics, Springer Nature, 2017
60.
Lotus Seedpod-Derived Hard Carbon with Hierarchical Porous Structure as Stable Anode for Sodium-Ion Batteries
Wu F., Zhang M., Bai Y., Wang X., Dong R., Wu C.
ACS applied materials & interfaces, American Chemical Society (ACS), 2019
61.
Nano Hard Carbon Anodes for Sodium-Ion Batteries
Kim D., Kim D., Kim S., Lee E., Park S., Lee J., Yun Y., Choi S., Kang J.
Nanomaterials, Multidisciplinary Digital Publishing Institute (MDPI), 2019
62.
Activated hard carbon from orange peel for lithium/sodium ion battery anode with long cycle life
Xiang J., Lv W., Mu C., Zhao J., Wang B.
Journal of Alloys and Compounds, Elsevier, 2017
63.
Biomass derived hierarchical porous carbons as high-performance anodes for sodium-ion batteries
Wang H., Yu W., Shi J., Mao N., Chen S., Liu W.
Electrochimica Acta, Elsevier, 2016
64.
Synthesis of hard carbon from argan shells for Na-ion batteries
Dahbi M., Kiso M., Kubota K., Horiba T., Chafik T., Hida K., Matsuyama T., Komaba S.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2017
65.
Expanded biomass-derived hard carbon with ultra-stable performance in sodium-ion batteries
Zhu Z., Liang F., Zhou Z., Zeng X., Wang D., Dong P., Zhao J., Sun S., Zhang Y., Li X.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2018
66.
Expanding Interlayer Spacing of Hard Carbon by Natural K+ Doping to Boost Na-Ion Storage
Wu F., Liu L., Yuan Y., Li Y., Bai Y., Li T., Lu J., Wu C.
ACS applied materials & interfaces, American Chemical Society (ACS), 2018
67.
Rambutan peel based hard carbons as anode materials for sodium ion battery
Arie A.A., Kristianto H., Muljana H., Stievano L.
Fullerenes Nanotubes and Carbon Nanostructures, Taylor & Francis, 2019
68.
Hierarchical Porous Carbon Derived from Sichuan Pepper for High-Performance Symmetric Supercapacitor with Decent Rate Capability and Cycling Stability
Zhang H., Xiao W., Zhou W., Chen S., Zhang Y.
Nanomaterials, Multidisciplinary Digital Publishing Institute (MDPI), 2019
69.
Old-loofah-derived hard carbon for long cyclicity anode in sodium ion battery
Yu C., Hou H., Liu X., Yao Y., Liao Q., Dai Z., Li D.
International Journal of Hydrogen Energy, Elsevier, 2018
70.
Sosnowskyi Hogweed-Based Hard Carbons for Sodium-Ion Batteries
Lakienko G.P., Bobyleva Z.V., Apostolova M.O., Sultanova Y.V., Dyakonov A.K., Zakharkin M.V., Sobolev N.A., Alekseeva A.M., Drozhzhin O.A., Abakumov A.M., Antipov E.V.
Batteries, Multidisciplinary Digital Publishing Institute (MDPI), 2022
71.
Optimization of Large Scale Produced Hard Carbon Performance in Na-Ion Batteries: Effect of Precursor, Temperature and Processing Conditions
Irisarri E., Amini N., Tennison S., Ghimbeu C.M., Gorka J., Vix-Guterl C., Ponrouch A., Palacin M.R.
Journal of the Electrochemical Society, The Electrochemical Society, 2018
72.
Correlation Between Microstructure and Na Storage Behavior in Hard Carbon
Zhang B., Ghimbeu C.M., Laberty C., Vix-Guterl C., Tarascon J.
Advanced Energy Materials, Wiley, 2015
73.
Recent Progress in Design of Biomass-Derived Hard Carbons for Sodium Ion Batteries
Górka J., Vix-Guterl C., Matei Ghimbeu C.
C – Journal of Carbon Research, Multidisciplinary Digital Publishing Institute (MDPI), 2016
74.
Hard carbon as a negative electrode material for potassium-ion batteries prepared with high yield through a polytetrafluoroethylene-based precursor
Abramova E.N., Marat N., Rupasov D.P., Morozova P.A., Kirsanova M.A., Abakumov A.M.
Carbon Trends, Elsevier, 2021
75.
Predicting capacity of hard carbon anodes in sodium-ion batteries using porosity measurements
78.
Unveiling the role of hydrothermal carbon dots as anodes in sodium-ion batteries with ultrahigh initial coulombic efficiency
Xie F., Xu Z., Jensen A.C., Ding F., Au H., Feng J., Luo H., Qiao M., Guo Z., Lu Y., Drew A.J., Hu Y., Titirici M.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2019
79.
The Role of Hydrothermal Carbonization in Sustainable Sodium‐Ion Battery Anodes
Xu Z., Wang J., Guo Z., Xie F., Liu H., Yadegari H., Tebyetekerwa M., Ryan M.P., Hu Y., Titirici M.
Advanced Energy Materials, Wiley, 2022
81.
High-Performance Hard Carbon Anode: Tunable Local Structures and Sodium Storage Mechanism
Jin Y., Sun S., Ou M., Liu Y., Fan C., Sun X., Peng J., Li Y., Qiu Y., Wei P., Deng Z., Xu Y., Han J., Huang Y.
ACS Applied Energy Materials, American Chemical Society (ACS), 2018
82.
Caramelization as a Key Stage for the Preparation of Monolithic Hard Carbon with Advanced Performance in Sodium-Ion Batteries
Bobyleva Z.V., Drozhzhin O.A., Alekseeva A.M., Dosaev K.A., Peters G.S., Lakienko G.P., Perfilyeva T.I., Sobolev N.A., Maslakov K.I., Savilov S.V., Abakumov A.M., Antipov E.V.
ACS Applied Energy Materials, American Chemical Society (ACS), 2022
83.
Pinecone biomass-derived hard carbon anodes for high-performance sodium-ion batteries
Zhang T., Mao J., Liu X., Xuan M., Bi K., Zhang X.L., Hu J., Fan J., Chen S., Shao G.
RSC Advances, Royal Society of Chemistry (RSC), 2017
84.
Impact of biomass inorganic impurities on hard carbon properties and performance in Na-ion batteries
Beda A., Le Meins J., Taberna P., Simon P., Matei Ghimbeu C.
Sustainable Materials and Technologies, Elsevier, 2020
85.
Hard carbon derived from hazelnut shell with facile HCl treatment as high-initial-coulombic-efficiency anode for sodium ion batteries
87.
A revised mechanistic model for sodium insertion in hard carbons
Au H., Alptekin H., Jensen A.C., Olsson E., O’Keefe C.A., Smith T., Crespo-Ribadeneyra M., Headen T.F., Grey C.P., Cai Q., Drew A.J., Titirici M.
Energy and Environmental Science, Royal Society of Chemistry (RSC), 2020
88.
Doctoral Dissertation, Departamento de Química Inorgánica, Universidad del País Vasco
P.S. Fontecoba
2017
89.
Laser Synthesis of Hard Carbon for Anodes in Na-Ion Battery
Zhang B., Deschamps M., Ammar M., Raymundo-Piñero E., Hennet L., Batuk D., Tarascon J.
Advanced Materials Technologies, Wiley, 2016
91.
Sieving carbons promise practical anodes with extensible low-potential plateaus for sodium batteries
Li Q., Liu X., Tao Y., Huang J., Zhang J., Yang C., Zhang Y., Zhang S., Jia Y., Lin Q., Xiang Y., Cheng J., Lv W., Kang F., Yang Y., et. al.
National Science Review, Oxford University Press, 2022
92.
Study of nano-porous hard carbons as anode materials for lithium ion batteries
Yang J., Zhou X., Li J., Zou Y., Tang J.
Materials Chemistry and Physics, Elsevier, 2012
93.
MgO‐Template Synthesis of Extremely High Capacity Hard Carbon for Na‐Ion Battery
Kamiyama A., Kubota K., Igarashi D., Youn Y., Tateyama Y., Ando H., Gotoh K., Komaba S.
Angewandte Chemie - International Edition, Wiley, 2021
94.
Enabling Fast Na + Transfer Kinetics in the Whole‐Voltage‐Region of Hard‐Carbon Anodes for Ultrahigh‐Rate Sodium Storage
Yin X., Lu Z., Wang J., Feng X., Roy S., Liu X., Yang Y., Zhao Y., Zhang J.
Advanced Materials, Wiley, 2022
95.
Zinc Single‐Atom Regulated Hard Carbons for High Rate and Low Temperature Sodium Ion Batteries
Lu Z., Wang J., Feng W., Yin X., Feng X., Zhao S., Li C., Wang R., Huang Q., Zhao Y.
Advanced Materials, Wiley, 2023
96.
New Template Synthesis of Anomalously Large Capacity Hard Carbon for Na‐ and K‐Ion Batteries
Igarashi D., Tanaka Y., Kubota K., Tatara R., Maejima H., Hosaka T., Komaba S.
Advanced Energy Materials, Wiley, 2023
97.
The recent progress of nitrogen-doped carbon nanomaterials for electrochemical batteries
Wu J., Pan Z., Zhang Y., Wang B., Peng H.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2018
98.
Nitrogen-doped carbon/graphene hybrid anode material for sodium-ion batteries with excellent rate capability
Liu H., Jia M., Cao B., Chen R., Lv X., Tang R., Wu F., Xu B.
Journal of Power Sources, Elsevier, 2016
99.
Nitrogen doped porous carbon fibres as anode materials for sodium ion batteries with excellent rate performance
Fu L., Tang K., Song K., van Aken P.A., Yu Y., Maier J.
Nanoscale, Royal Society of Chemistry (RSC), 2014
101.
Experimental design and theoretical calculation for sulfur-doped carbon nanofibers as a high performance sodium-ion battery anode
Jin Q., Li W., Wang K., Feng P., Li H., Gu T., Zhou M., Wang W., Cheng S., Jiang K.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2019
103.
Mechanism of Na‐Ion Storage in Hard Carbon Anodes Revealed by Heteroatom Doping
Li Z., Bommier C., Chong Z.S., Jian Z., Surta T.W., Wang X., Xing Z., Neuefeind J.C., Stickle W.F., Dolgos M., Greaney P.A., Ji X.
Advanced Energy Materials, Wiley, 2017
104.
Boric Acid Assisted Reduction of Graphene Oxide: A Promising Material for Sodium-Ion Batteries
Wang Y., Wang C., Wang Y., Liu H., Huang Z.
ACS applied materials & interfaces, American Chemical Society (ACS), 2016
105.
Hard carbon nanoparticles as high-capacity, high-stability anodic materials for Na-ion batteries
Xiao L., Cao Y., Henderson W.A., Sushko M.L., Shao Y., Xiao J., Wang W., Engelhard M.H., Nie Z., Liu J.
Nano Energy, Elsevier, 2016
106.
Nitrogen-Doped Porous Carbon Nanosheets as Low-Cost, High-Performance Anode Material for Sodium-Ion Batteries
Wang H., Wu Z., Meng F., Ma D., Huang X., Wang L., Zhang X.
ChemSusChem, Wiley, 2012
108.
25th Anniversary Article: Chemically Modified/Doped Carbon Nanotubes & Graphene for Optimized Nanostructures & Nanodevices
Maiti U.N., Lee W.J., Lee J.M., Oh Y., Kim J.Y., Kim J.E., Shim J., Han T.H., Kim S.O.
Advanced Materials, Wiley, 2013
109.
Sulfur/Oxygen Codoped Porous Hard Carbon Microspheres for High-Performance Potassium-Ion Batteries
Chen M., Wang W., Liang X., Gong S., Liu J., Wang Q., Guo S., Yang H.
Advanced Energy Materials, Wiley, 2018
111.
Nitrogen and oxygen co-doping carbon microspheres by a sustainable route for fast sodium-ion batteries
Zhang F., Qin D., Xu J., Liu Z., Zhao Y., Zhang X.
Electrochimica Acta, Elsevier, 2019
112.
A Large Scalable and Low‐Cost Sulfur/Nitrogen Dual‐Doped Hard Carbon as the Negative Electrode Material for High‐Performance Potassium‐Ion Batteries
Liu Y., Dai H., Wu L., Zhou W., He L., Wang W., Yan W., Huang Q., Fu L., Wu Y.
Advanced Energy Materials, Wiley, 2019
113.
A SAXS outlook on disordered carbonaceous materials for electrochemical energy storage
Saurel D., Segalini J., Jauregui M., Pendashteh A., Daffos B., Simon P., Casas-Cabanas M.
Energy Storage Materials, Elsevier, 2019
114.
Hydrothermal carbon from biomass: a comparison of the local structure from poly- to monosaccharides and pentoses/hexoses
Titirici M., Antonietti M., Baccile N.
Green Chemistry, Royal Society of Chemistry (RSC), 2008
115.
Correlation of Structure and Performance of Hard Carbons as Anodes for Sodium Ion Batteries
Gomez-Martin A., Martinez-Fernandez J., Ruttert M., Winter M., Placke T., Ramirez-Rico J.
Chemistry of Materials, American Chemical Society (ACS), 2019
116.
Modulating the Graphitic Domains and Pore Structure of Corncob-Derived Hard Carbons by Pyrolysis to Improve Sodium Storage
Song N., Guo N., Ma C., Zhao Y., Li W., Li B.
Molecules, Multidisciplinary Digital Publishing Institute (MDPI), 2023
117.
Typha-derived hard carbon for high-performance sodium ion storage
Shen Y., Sun S., Yang M., Zhao X.
Journal of Alloys and Compounds, Elsevier, 2019
118.
Rapid Sodium-Ion Storage in Hard Carbon Anode Material Derived from Ganoderma lucidum Residue with Inherent Open Channels
Lu P., Xia J., Dong X.
ACS Sustainable Chemistry and Engineering, American Chemical Society (ACS), 2019
119.
Synthesis and characterization of PVDF‐coated cotton‐derived hard carbon for anode of Li‐ion batteries
Li L., Fan C., Tang Y., Zeng B.
International Journal of Energy Research, Wiley, 2019
121.
Biomorphic carbon derived from corn husk as a promising anode materials for potassium ion battery
Wang Q., Gao C., Zhang W., Luo S., Zhou M., Liu Y., Liu R., Zhang Y., Wang Z., Hao A.
Electrochimica Acta, Elsevier, 2019
122.
Hierarchically porous structured carbon derived from peanut shell as an enhanced high rate anode for lithium ion batteries
Murali G., Harish S., Ponnusamy S., Ragupathi J., Therese H.A., Navaneethan M., Muthamizhchelvan C.
Applied Surface Science, Elsevier, 2019
123.
Sodium storage in hard carbon with curved graphene platelets as the basic structural units
Wang K., Xu Y., Li Y., Dravid V., Wu J., Huang Y.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2019
124.
Facile synthesis of Camellia oleifera shell-derived hard carbon as an anode material for lithium-ion batteries
Ma B., Huang Y., Nie Z., Qiu X., Su D., Wang G., Yuan J., Xie X., Wu Z.
RSC Advances, Royal Society of Chemistry (RSC), 2019
125.
Three-dimensional hard carbon matrix for sodium-ion battery anode with superior-rate performance and ultralong cycle life
Yuan Z., Si L., Zhu X.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2015
126.
Hard Carbon Microtubes Made from Renewable Cotton as High-Performance Anode Material for Sodium-Ion Batteries
127.
Hard carbon derived from sepals of Palmyra palm fruit calyx as an anode for sodium-ion batteries
Damodar D., Ghosh S., Usha Rani M., Martha S.K., Deshpande A.S.
Journal of Power Sources, Elsevier, 2019
128.
Regulating Pore Structure of Hierarchical Porous Waste Cork‐Derived Hard Carbon Anode for Enhanced Na Storage Performance
Li Y., Lu Y., Meng Q., Jensen A.C., Zhang Q., Zhang Q., Tong Y., Qi Y., Gu L., Titirici M., Hu Y.
Advanced Energy Materials, Wiley, 2019
129.
Waste Beverage Coffee-Induced Hard Carbon Granules for Sodium-Ion Batteries
Lee M.E., Kwak H.W., Jin H., Yun Y.S.
ACS Sustainable Chemistry and Engineering, American Chemical Society (ACS), 2019
130.
Zhou Z.
International Journal of Electrochemical Science, Electrochemical Science Group, University of Belgrade, 2019
131.
Structural Analysis of Sucrose-Derived Hard Carbon and Correlation with the Electrochemical Properties for Lithium, Sodium, and Potassium Insertion
Kubota K., Shimadzu S., Yabuuchi N., Tominaka S., Shiraishi S., Abreu-Sepulveda M., Manivannan A., Gotoh K., Fukunishi M., Dahbi M., Komaba S.
Chemistry of Materials, American Chemical Society (ACS), 2020
132.
Features of the Synthesis of Functional Carbon Materials from Plant Carbohydrates
Bobyleva Z.V., Apostolova M.O., Lakienko G.P., Alekseeva A.M., Drozhzhin O.A., Antipov E.V.
Chemistry and Technology of Fuels and Oils, Springer Nature, 2022
133.
Hard carbon derived from cellulose as anode for sodium ion batteries: Dependence of electrochemical properties on structure
Simone V., Boulineau A., de Geyer A., Rouchon D., Simonin L., Martinet S.
Journal of Energy Chemistry, Elsevier, 2016
134.
Low-Cost and High-Performance Hard Carbon Anode Materials for Sodium-Ion Batteries
Wang K., Jin Y., Sun S., Huang Y., Peng J., Luo J., Zhang Q., Qiu Y., Fang C., Han J.
ACS Omega, American Chemical Society (ACS), 2017
135.
Studies on electrochemical sodium storage into hard carbons with binder-free monolithic electrodes
Hasegawa G., Kanamori K., Kannari N., Ozaki J., Nakanishi K., Abe T.
Journal of Power Sources, Elsevier, 2016
136.
Revealing sodium ion storage mechanism in hard carbon
Alvin S., Yoon D., Chandra C., Cahyadi H.S., Park J., Chang W., Chung K.Y., Kim J.
Carbon, Elsevier, 2019
137.
Non-Aqueous K-Ion Battery Based on Layered K0.3MnO2and Hard Carbon/Carbon Black
Vaalma C., Giffin G.A., Buchholz D., Passerini S.
Journal of the Electrochemical Society, The Electrochemical Society, 2016
138.
Controlled synthesis of macroscopic three-dimensional hollow reticulate hard carbon as long-life anode materials for Na-ion batteries
Li W., Huang J., Feng L., Cao L., Ren Y., Li R., Xu Z., Li J., Yao C.
Journal of Alloys and Compounds, Elsevier, 2017
139.
New Mechanistic Insights on Na-Ion Storage in Nongraphitizable Carbon
Bommier C., Surta T.W., Dolgos M., Ji X.
Nano Letters, American Chemical Society (ACS), 2015
140.
Following the in-plane disorder of sodiated hard carbon through operando total scattering
Mathiesen J.K., Väli R., Härmas M., Lust E., Fold von Bülow J., Jensen K.M., Norby P.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2019
141.
Mechanistic insights into sodium storage in hard carbon anodes using local structure probes
Stratford J.M., Allan P.K., Pecher O., Chater P.A., Grey C.P.
Chemical Communications, Royal Society of Chemistry (RSC), 2016
142.
Correlating Local Structure and Sodium Storage in Hard Carbon Anodes: Insights from Pair Distribution Function Analysis and Solid-State NMR
Stratford J.M., Kleppe A.K., Keeble D.S., Chater P.A., Meysami S.S., Wright C.J., Barker J., Titirici M., Allan P.K., Grey C.P.
Journal of the American Chemical Society, American Chemical Society (ACS), 2021
143.
Neutron scattering for lithium-ion batteries: analysis of materials and processes
Balagurov A.M., Bobrikov I.A., Samoylova N.Y., Drozhzhin O.A., Antipov E.V.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2014
144.
Highly disordered hard carbon derived from skimmed cotton as a high-performance anode material for potassium-ion batteries
He X., Liao J., Tang Z., Xiao L., Ding X., Hu Q., Wen Z., Chen C.
Journal of Power Sources, Elsevier, 2018
145.
Evolution of glassy carbon under heat treatment: correlation structure–mechanical properties
Jurkiewicz K., Pawlyta M., Zygadło D., Chrobak D., Duber S., Wrzalik R., Ratuszna A., Burian A.
Journal of Materials Science, Springer Nature, 2017
146.
Raman microspectroscopy of soot and related carbonaceous materials: Spectral analysis and structural information
147.
Studying disorder in graphite-based systems by Raman spectroscopy
Pimenta M.A., Dresselhaus G., Dresselhaus M.S., Cançado L.G., Jorio A., Saito R.
Physical Chemistry Chemical Physics, Royal Society of Chemistry (RSC), 2007
148.
Double Resonant Raman Scattering in Graphite
Thomsen C., Reich S.
Physical Review Letters, American Physical Society (APS), 2000
149.
In Nanokremnii: Svoistva, Poluchenie, Primenenie, Metody Issledovaniya i Kontrolya. (Nanosilicon: Properties, Production, Application, Methods of Research and Control). (Moscow: Fizmatlit). 405 p.
A.A.Ishenko, G.V, Fetisov, L.A.Aslanov
2011
151.
Sodiation energetics in pore size controlled hard carbons determined via entropy profiling†
Mercer M., Nagarathinam M., Gavilán Arriazu E.M., Binjrajka A., Panda S., Au H., Crespo-Ribadeneyra M., Titirici M., Leiva E., Hoster H.E.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2023
152.
Hard carbon porosity revealed by the adsorption of multiple gas probe molecules (N2, Ar, CO2, O2 and H2)
Beda A., Vaulot C., Matei Ghimbeu C.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2021
153.
The role of specific and active surface areas in optimizing hard carbon irreversible capacity loss in sodium ion batteries
Beda A., Vaulot C., Rabuel F., Morcrette M., Matei Ghimbeu C.
Energy Advances, Royal Society of Chemistry (RSC), 2022
154.
Hard carbon key properties allow for the achievement of high Coulombic efficiency and high volumetric capacity in Na-ion batteries
Beda A., Rabuel F., Morcrette M., Knopf S., Taberna P., Simon P., Matei Ghimbeu C.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2021
155.
Nanoporous hard carbon microspheres as anode active material of lithium ion battery
Jafari S.M., Khosravi M., Mollazadeh M.
Electrochimica Acta, Elsevier, 2016
156.
Enhancement of First Cycle Coulombic Efficiency of Hard Carbon Derived from Eucalyptus in a Sodium Ion Battery
Nakabayashi K., Yi H., Ryu D., Chung D., Miyawaki J., Yoon S.
Chemistry Letters, The Chemical Society of Japan, 2019
157.
In Porous Media: Fluid Transport and Pore Structure. (New York: Academic Press). P. 180
F.A.L. Dullien
2012
158.
In Adsorbtsiya. Tekstura Dispersnykh i Poristykh Materialov. (Adsorption. Texture of Dispersed and Porous Materials) (Novosibirsk: Nauka). 470 p.
A.P. Karnaukhov
1999
159.
Izmerenie Ploshchadi Poverkhnosti i Poristosti Metodom Kapillyarnoi Kondensatsii Azota. Metedologicheskaya Razrabotka. (Measurement of Surface Area and Porosity Using Capillary Nitrogen Condensation. Methodological Development). (Moscow: MSU). P. 18
A.S.Vyacheslavov, E.A.Pomerantseva
2006
160.
Modeling of type IV and V sigmoidal adsorption isotherms
Buttersack C.
Physical Chemistry Chemical Physics, Royal Society of Chemistry (RSC), 2019
161.
Doctoral Dissertation. Universite Paris-Est
Z. Zhang
2014
162.
Hard Carbon Microspheres: Potassium-Ion Anode Versus Sodium-Ion Anode
Jian Z., Xing Z., Bommier C., Li Z., Ji X.
Advanced Energy Materials, Wiley, 2015
163.
Defective Graphene as a High-Capacity Anode Material for Na- and Ca-Ion Batteries
Datta D., Li J., Shenoy V.B.
ACS applied materials & interfaces, American Chemical Society (ACS), 2014
165.
Interaction of the Stone-Wales defects in graphene
Openov L.A., Podlivaev A.I.
Physics of the Solid State, Pleiades Publishing, 2015
166.
Correlation of the irreversible lithium capacity with the active surface area of modified carbons
Béguin F., Chevallier F., Vix-Guterl C., Saadallah S., Bertagna V., Rouzaud J.N., Frackowiak E.
Carbon, Elsevier, 2005
167.
Enhanced Capacity and Rate Capability of Nitrogen/Oxygen Dual-Doped Hard Carbon in Capacitive Potassium-Ion Storage
Yang J., Ju Z., Jiang Y., Xing Z., Xi B., Feng J., Xiong S.
Advanced Materials, Wiley, 2017
169.
High capacity hard carbon anodes for sodium ion batteries in additive free electrolyte
Ponrouch A., Goñi A.R., Palacín M.R.
Electrochemistry Communications, Elsevier, 2013
170.
New Paradigms on the Nature of Solid Electrolyte Interphase Formation and Capacity Fading of Hard Carbon Anodes in Na‐Ion Batteries
Bommier C., Leonard D., Jian Z., Stickle W.F., Greaney P.A., Ji X.
Advanced Materials Interfaces, Wiley, 2016
171.
Biomass-derived carbon: synthesis and applications in energy storage and conversion
Deng J., Li M., Wang Y.
Green Chemistry, Royal Society of Chemistry (RSC), 2016
172.
Hard Carbon Anodes for Na‐Ion Batteries: Toward a Practical Use
Hasegawa G., Kanamori K., Kannari N., Ozaki J., Nakanishi K., Abe T.
ChemElectroChem, Wiley, 2015
175.
Optimization of Na-Ion Battery Systems Based on Polyanionic or Layered Positive Electrodes and Carbon Anodes
Dugas R., Zhang B., Rozier P., Tarascon J.M.
Journal of the Electrochemical Society, The Electrochemical Society, 2016
176.
Insertion compounds and composites made by ball milling for advanced sodium-ion batteries
Zhang B., Dugas R., Rousse G., Rozier P., Abakumov A.M., Tarascon J.
Nature Communications, Springer Nature, 2016
177.
Towards high energy density, low cost and safe Na-ion full-cell using P2–Na0.67[Fe0.5Mn0.5]O2 and Na2C4O4 sacrificial salt
Martínez De Ilarduya J., Otaegui L., Galcerán M., Acebo L., Shanmukaraj D., Rojo T., Armand M.
Electrochimica Acta, Elsevier, 2019
178.
Presodiation Strategies and Their Effect on Electrode–Electrolyte Interphases for High-Performance Electrodes for Sodium-Ion Batteries
Moeez I., Jung H., Lim H., Chung K.Y.
ACS applied materials & interfaces, American Chemical Society (ACS), 2019
179.
Chemically Presodiated Hard Carbon Anodes with Enhanced Initial Coulombic Efficiencies for High-Energy Sodium Ion Batteries
Liu M., Zhang J., Guo S., Wang B., Shen Y., Ai X., Yang H., Qian J.
ACS applied materials & interfaces, American Chemical Society (ACS), 2020
180.
Solution-based chemical pre-alkaliation of metal-ion battery cathode materials for increased capacity
Kapaev R.R., Stevenson K.J.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2021
181.
Improving the Performance of Hard Carbon//Na3V2O2(PO4)2F Sodium-Ion Full Cells by Utilizing the Adsorption Process of Hard Carbon
Shen B., You Y., Niu Y., Li Y., Dai C., Hu L., Guo B., Jiang J., Bao S., Xu M.
ACS applied materials & interfaces, American Chemical Society (ACS), 2018
182.
Patent RF 2731884C1
2020
183.
Low-surface-area hard carbon anode for na-ion batteries via graphene oxide as a dehydration agent.
Luo W., Bommier C., Jian Z., Li X., Carter R., Vail S., Lu Y., Lee J., Ji X.
ACS applied materials & interfaces, American Chemical Society (ACS), 2015
184.
Self-supported binder-free hard carbon electrodes for sodium-ion batteries: insights into their sodium storage mechanisms
Beda A., Villevieille C., Taberna P., Simon P., Matei Ghimbeu C.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2020
186.
Hard–Soft Carbon Composite Anodes with Synergistic Sodium Storage Performance
Xie F., Xu Z., Jensen A.C., Au H., Lu Y., Araullo‐Peters V., Drew A.J., Hu Y., Titirici M.
Advanced Functional Materials, Wiley, 2019
188.
Bio-derived hard carbon nanosheets with high rate sodium-ion storage characteristics
Asfaw H.D., Gond R., Kotronia A., Tai C., Younesi R.
Sustainable Materials and Technologies, Elsevier, 2022
189.
Electrolytes and Interphases in Sodium‐Based Rechargeable Batteries: Recent Advances and Perspectives
Eshetu G.G., Elia G.A., Armand M., Forsyth M., Komaba S., Rojo T., Passerini S.
Advanced Energy Materials, Wiley, 2020
190.
Variable-resistance materials for lithium-ion batteries
Beletskii E.V., Alekseeva E.V., Levin O.V.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2022
191.
Role of electrolyte in stabilizing hard carbon as an anode for rechargeable sodium-ion batteries with long cycle life
Hirsh H.S., Sayahpour B., Shen A., Li W., Lu B., Zhao E., Zhang M., Meng Y.S.
Energy Storage Materials, Elsevier, 2021
192.
Unraveling the Mechanism of Different Kinetics Performance between Ether and Carbonate Ester Electrolytes in Hard Carbon Electrode
Yi X., Li X., Zhong J., Wang S., Wang Z., Guo H., Wang J., Yan G.
Advanced Functional Materials, Wiley, 2022
193.
Revisit Electrolyte Chemistry of Hard Carbon in Ether for Na Storage
Pan J., Sun Y., Yan Y., Feng L., Zhang Y., Lin A., Huang F., Yang J.
JACS Au, American Chemical Society (ACS), 2021
195.
Research Development on K-Ion Batteries
Hosaka T., Kubota K., Hameed A.S., Komaba S.
Chemical Reviews, American Chemical Society (ACS), 2020
196.
Unusual Passivation Ability of Superconcentrated Electrolytes toward Hard Carbon Negative Electrodes in Sodium-Ion Batteries
Takada K., Yamada Y., Watanabe E., Wang J., Sodeyama K., Tateyama Y., Hirata K., Kawase T., Yamada A.
ACS applied materials & interfaces, American Chemical Society (ACS), 2017
197.
Moderately concentrated electrolyte improves solid–electrolyte interphase and sodium storage performance of hard carbon
Patra J., Huang H., Xue W., Wang C., Helal A.S., Li J., Chang J.
Energy Storage Materials, Elsevier, 2019
198.
Beyond the Norm: Synthesis and Electrochemical Study of High Concentrated NaPF6 Electrolytes
Ould D.M., Menkin S., O'Keefe C.A., Coowar F., Barker J., Grey C.P., Wright D.S.
ECS Meeting Abstracts, The Electrochemical Society, 2022
199.
Reversible and High-rate Hard Carbon Negative Electrodes in a Fluorine-free Sodium-salt Electrolyte
MORIKAWA Y., YAMADA Y., DOI K., NISHIMURA S., YAMADA A.
Electrochemistry, The Electrochemical Society of Japan, 2020
200.
A Halogen‐Free and Flame‐Retardant Sodium Electrolyte Compatible with Hard Carbon Anodes
Colbin L.O., Mogensen R., Buckel A., Wang Y., Naylor A.J., Kullgren J., Younesi R.
Advanced Materials Interfaces, Wiley, 2021
201.
Optimization of Sodium Bis(oxalato)borate (NaBOB) in Triethyl Phosphate (TEP) by Electrolyte Additives
Welch J., Mogensen R., van Ekeren W., Eriksson H., J. Naylor A., Younesi R.
Journal of the Electrochemical Society, The Electrochemical Society, 2022
202.
Hard Carbon Anode with a Sodium Carborane Electrolyte for Fast-Charging All-Solid-State Sodium-Ion Batteries
Niitani K., Ushiroda S., Kuwata H., Ohata H.N., Shimo Y., Hozumi M., Matsunaga T., Nakanishi S.
ACS Energy Letters, American Chemical Society (ACS), 2021
203.
Assessment of the Electrochemical Stability of Carbonate-Based Electrolytes in Na-Ion Batteries
Yan G., Alves-Dalla-Corte D., Yin W., Madern N., Gachot G., Tarascon J.
Journal of the Electrochemical Society, The Electrochemical Society, 2018
204.
A New Electrolyte Formulation for Securing High Temperature Cycling and Storage Performances of Na‐Ion Batteries
Yan G., Reeves K., Foix D., Li Z., Cometto C., Mariyappan S., Salanne M., Tarascon J.
Advanced Energy Materials, Wiley, 2019
205.
Origins of irreversible capacity loss in hard carbon negative electrodes for potassium-ion batteries
Katorova N.S., Luchkin S.Y., Rupasov D.P., Abakumov A.M., Stevenson K.J.
Journal of Chemical Physics, American Institute of Physics (AIP), 2020
206.
Effect of Concentrated Diglyme-Based Electrolytes on the Electrochemical Performance of Potassium-Ion Batteries
Katorova N.S., Fedotov S.S., Rupasov D.P., Luchinin N.D., Delattre B., Chiang Y., Abakumov A.M., Stevenson K.J.
ACS Applied Energy Materials, American Chemical Society (ACS), 2019
207.
Superior potassium storage behavior of hard carbon facilitated by ether-based electrolyte
Dai H., Zeng Z., Yang X., Jiang M., Wang Y., Huang Q., Liu L., Fu L., Zhang P., Wu Y.
Carbon, Elsevier, 2021
208.
The roles of electrolyte chemistry in hard carbon anode for potassium-ion batteries
Wu Z., Zou J., Shabanian S., Golovin K., Liu J.
Chemical Engineering Journal, Elsevier, 2022
211.
Research progress of biomass carbon materials as anode materials for potassium-ion batteries
Li X., Zhou Y., Deng B., Li J., Xiao Z.
Frontiers in Chemistry, Frontiers Media S.A., 2023
212.
Nitrogen doping and graphitization tuning coupled hard carbon for superior potassium-ion storage
213.
Fast-Charging of Hybrid Lithium-Ion/Lithium-Metal Anodes by Nanostructured Hard Carbon Host
Gong H., Chen Y., Chen S., Xu C., Yang Y., Ye Y., Huang Z., Ning R., Cui Y., Bao Z.
ACS Energy Letters, American Chemical Society (ACS), 2022
214.
Quasi-metallic lithium encapsulated in the subnanopores of hard carbon for hybrid lithium–ion/lithium metal batteries
Su K., Jin T., Zhang C.H., Wang R., Yuan S., Li N.W., Yu L.
Chemical Engineering Journal, Elsevier, 2022
215.
A hybrid lithium storage mechanism of hard carbon enhances its performance as anodes for lithium-ion batteries
Wang K., Xu Y., Wu H., Yuan R., Zong M., Li Y., Dravid V., Ai W., Wu J.
Carbon, Elsevier, 2021
216.
Hard Carbon Anodes for Next‐Generation Li‐Ion Batteries: Review and Perspective
Xie L., Tang C., Bi Z., Song M., Fan Y., Yan C., Li X., Su F., Zhang Q., Chen C.
Advanced Energy Materials, Wiley, 2021
217.
Unveiling pseudocapacitive behavior of hard carbon anode materials for sodium-ion batteries
Bobyleva Z.V., Drozhzhin O.A., Dosaev K.A., Kamiyama A., Ryazantsev S.V., Komaba S., Antipov E.V.
Electrochimica Acta, Elsevier, 2020
218.
The Mechanisms of Lithium and Sodium Insertion in Carbon Materials
Stevens D.A., Dahn J.R.
Journal of the Electrochemical Society, The Electrochemical Society, 2002
219.
Chemical Origins of a Fast-Charge Performance in Disordered Carbon Anodes
Ahn S., Lagnoni M., Yuan Y., Ogarev A., Vavrinyuk E., Voynov G., Barrett E., Pelli A., Atrashchenko A., Platonov A., Gurevich S., Gorokhov M., Rupasov D., Robertson A.W., House R.A., et. al.
ACS Applied Energy Materials, American Chemical Society (ACS), 2023
220.
Impact of Sodium Metal Plating on Cycling Performance of Layered Oxide/Hard Carbon Sodium-ion Pouch Cells with Different Voltage Cut-offs
Hijazi H., Ye Z., Zhang L., Deshmukh J., Johnson M., Dahn J.R., Metzger M.
Journal of the Electrochemical Society, The Electrochemical Society, 2023
221.
High Capacity Anode Materials for Rechargeable Sodium-Ion Batteries
Stevens D.A., Dahn J.R.
Journal of the Electrochemical Society, The Electrochemical Society, 2002
222.
Electrochemical Na Insertion and Solid Electrolyte Interphase for Hard-Carbon Electrodes and Application to Na-Ion Batteries
Komaba S., Murata W., Ishikawa T., Yabuuchi N., Ozeki T., Nakayama T., Ogata A., Gotoh K., Fujiwara K.
Advanced Functional Materials, Wiley, 2011
223.
Carbon Microspheres Obtained from Resorcinol-Formaldehyde as High-Capacity Electrodes for Sodium-Ion Batteries
Alcántara R., Lavela P., Ortiz G.F., Tirado J.L.
Electrochemical and Solid-State Letters, The Electrochemical Society, 2005
224.
NMR study for electrochemically inserted Na in hard carbon electrode of sodium ion battery
Gotoh K., Ishikawa T., Shimadzu S., Yabuuchi N., Komaba S., Takeda K., Goto A., Deguchi K., Ohki S., Hashi K., Shimizu T., Ishida H.
Journal of Power Sources, Elsevier, 2013
225.
Sodium Ion Insertion in Hollow Carbon Nanowires for Battery Applications
Cao Y., Xiao L., Sushko M.L., Wang W., Schwenzer B., Xiao J., Nie Z., Saraf L.V., Yang Z., Liu J.
Nano Letters, American Chemical Society (ACS), 2012
226.
Carbon Nanosheet Frameworks Derived from Peat Moss as High Performance Sodium Ion Battery Anodes
Ding J., Wang H., Li Z., Kohandehghan A., Cui K., Xu Z., Zahiri B., Tan X., Lotfabad E.M., Olsen B.C., Mitlin D.
ACS Nano, American Chemical Society (ACS), 2013
227.
Manipulating Adsorption–Insertion Mechanisms in Nanostructured Carbon Materials for High‐Efficiency Sodium Ion Storage
Qiu S., Xiao L., Sushko M.L., Han K.S., Shao Y., Yan M., Liang X., Mai L., Feng J., Cao Y., Ai X., Yang H., Liu J.
Advanced Energy Materials, Wiley, 2017
228.
Insights on the Na+ ion storage mechanism in hard carbon: Discrimination between the porosity, surface functional groups and defects
Matei Ghimbeu C., Górka J., Simone V., Simonin L., Martinet S., Vix-Guterl C.
Nano Energy, Elsevier, 2018
229.
Hard carbons for sodium-ion batteries and beyond
Xie F., Xu Z., Guo Z., Titirici M.
Progress in Energy, IOP Publishing, 2020
231.
Combination of solid state NMR and DFT calculation to elucidate the state of sodium in hard carbon electrodes
Morita R., Gotoh K., Fukunishi M., Kubota K., Komaba S., Nishimura N., Yumura T., Deguchi K., Ohki S., Shimizu T., Ishida H.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2016
232.
Mechanism of Sodium Storage in Hard Carbon: An X‐Ray Scattering Analysis
Morikawa Y., Nishimura S., Hashimoto R., Ohnuma M., Yamada A.
Advanced Energy Materials, Wiley, 2019
233.
Elucidating the Sodiation Mechanism in Hard Carbon by Operando Raman Spectroscopy
Weaving J.S., Lim A., Millichamp J., Neville T.P., Ledwoch D., Kendrick E., McMillan P.F., Shearing P.R., Howard C.A., Brett D.J.
ACS Applied Energy Materials, American Chemical Society (ACS), 2020
234.
Insight into Sodium Insertion and the Storage Mechanism in Hard Carbon
Anji Reddy M., Helen M., Groß A., Fichtner M., Euchner H.
ACS Energy Letters, American Chemical Society (ACS), 2018
235.
Alkali metal insertion into hard carbon – the full picture
Euchner H., Vinayan B.P., Reddy M.A., Fichtner M., Groß A.
Journal of Materials Chemistry A, Royal Society of Chemistry (RSC), 2020
236.
Raman spectroscopic investigations of activated carbon materials
Shimodaira N., Masui A.
Journal of Applied Physics, American Institute of Physics (AIP), 2002
238.
Revealing the Intercalation Mechanisms of Lithium, Sodium, and Potassium in Hard Carbon
Alvin S., Cahyadi H.S., Hwang J., Chang W., Kwak S.K., Kim J.
Advanced Energy Materials, Wiley, 2020
240.
Correlation between Potassium-Ion Storage Mechanism and Local Structural Evolution in Hard Carbon Materials
Xu J., Fan C., Ou M., Sun S., Xu Y., Liu Y., Wang X., Li Q., Fang C., Han J.
Chemistry of Materials, American Chemical Society (ACS), 2022
241.
Microstructure‐Dependent K + Storage in Porous Hard Carbon
Li W., Zhang R., Chen Z., Fan B., Xiao K., Liu H., Gao P., Wu J., Tu C., Liu J.
Small, Wiley, 2021
244.
Storage Mechanism of Alkali Metal Ions in the Hard Carbon Anode: an Electrochemical Viewpoint
Huang Y., Wang Y., Bai P., Xu Y.
ACS applied materials & interfaces, American Chemical Society (ACS), 2021
245.
In Tekhnicheskii Uglerod: Morfologiya, Svoistva, Proizvodstvo. (Carbon Black: Morphology, Properties, Production) (Moscow: Kauchuk i Pezina). 586 p.
T.G.Gul’misaryan, V.M.Kapustin, I.P.Levenberg
2017
246.
Ion-Induced Soot Nucleation Using a New Potential for Curved Aromatics
Bowal K., Martin J.W., Misquitta A.J., Kraft M.
Combustion Science and Technology, Taylor & Francis, 2019
247.
Nanoglobular carbon and palladium–nanoglobular carbon catalysts for liquid-phase hydrogenation of organic compounds
Mironenko R.M., Likholobov V.A., Belskaya O.B.
Russian Chemical Reviews, Autonomous Non-profit Organization Editorial Board of the journal Uspekhi Khimii, 2022
248.
Homogenous metallic deposition regulated by defect-rich skeletons for sodium metal batteries
Xu Z., Guo Z., Madhu R., Xie F., Chen R., Wang J., Tebyetekerwa M., Hu Y., Titirici M.
Energy and Environmental Science, Royal Society of Chemistry (RSC), 2021