Next Article in Journal
A Stable Porous Aluminum Electrode with High Capacity for Rechargeable Lithium-Ion Batteries
Next Article in Special Issue
Grain Boundary Characterization and Potential Percolation of the Solid Electrolyte LLZO
Previous Article in Journal
High-Performance Metal–Chalcogen Batteries
Previous Article in Special Issue
In Situ Solidified Gel Polymer Electrolytes for Stable Solid−State Lithium Batteries at High Temperatures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress and Perspectives of Solid State Na-CO2 Batteries

1
School of Chemical & Environmental Engineering, China University of Mining and Technology-Beijing, Beijing 100083, China
2
21C Innovation Laboratory, Contemporary Amperex Technology Ltd. (21C LAB), Ningde 352100, China
3
Key Laboratory of Advanced Energy Materials Chemistry (Ministry of Education), College of Chemistry, Nankai University, Tianjin 300071, China
*
Authors to whom correspondence should be addressed.
Batteries 2023, 9(1), 36; https://doi.org/10.3390/batteries9010036
Submission received: 23 November 2022 / Revised: 26 December 2022 / Accepted: 30 December 2022 / Published: 4 January 2023

Abstract

:
Solid state Na-CO2 batteries are a kind of promising energy storage system, which can use excess CO2 for electrochemical energy storage. They not only have high theoretical energy densities, but also feature a high safety level of solid-state batteries and low cost owing to abundant sodium metal resources. Although many efforts have been made, the practical application of Na-CO2 battery technology is still hampered by some crucial challenges, including short cycle life, high charging potential, poor rate performance and lower specific full discharge capacity. This paper systematically reviews the recent research advances in Na-CO2 batteries in terms of understanding the mechanism of CO2 reduction, carbonate formation and decomposition reaction, design strategies of cathode electrocatalysts, solid electrolytes and their interface design. In addition, the application of advanced in situ characterization techniques and theoretical calculation of metal–CO2 batteries are briefly introduced, and the combination of theory and experiment in the research of battery materials is discussed as well. Finally, the opportunities and key challenges of solid-state Na-CO2 electrochemical systems in the carbon-neutral era are presented.

Graphical Abstract

1. Introduction

Metal–CO2 batteries (such as Li/Na/Zn/K-CO2 batteries) are a high energy density energy storage and power supply technology that enables CO2 fixation and conversion [1,2]. Among various kinds of metal–CO2 batteries, Na-CO2 batteries have attracted more attention because of the abundant resources of sodium metal and similar physical and chemical properties to lithium. Na-CO2 batteries exhibit better comprehensive performance, including high energy density (1.13 kWh kg−1), and relatively high working voltage (2.35 V) [3]. In addition, the abundance of metal sodium is 1352 times that of metal lithium, so the cost is lower (the price of metal sodium is 20-times cheaper than that of metal lithium) [4,5]. In addition, a Na-CO2 battery with the reaction of 4 Na + 3 CO2 ↔ 2 Na2CO3 + C (∆rG0m = −905.6 kJ mol−1) has a low reaction Gibbs free energy, which means that Na-CO2 batteries may have a low charging voltage and inhibit electrolyte decomposition, which is conducive to improving round-trip efficiency and prolonging service life [6,7]. Compared with Li+, Na+ as a charge carrier has other advantages, for example, sodium has a larger ionic radius and atomic mass than lithium, but the Stokes radius of sodium is smaller than lithium, so it has higher mobility and ionic conductivity, resulting in a smaller polarization [8]. Therefore, Na-CO2 batteries are widely considered to be a promising next generation energy storage power supply technology. However, the research of Na-CO2 batteries is still in its infancy, and there are problems, for example, the reaction mechanism is still unclear, the types of cathodic materials and catalytic activity are relatively limited, the ionic conductivity and interfacial stability of the solid electrolyte still need to be improved, the impedance between the electrode and the interface is large and dendrites grow at the metal anode interface during the reaction. These problems lead to unsatisfactory electrochemical performance, such as cycling stability, overpotential, rate capability and specific full discharge capacity of Na-CO2 batteries [9,10,11]. To solve these problems, it is necessary to comprehensively understand the reaction mechanism of Na-CO2 battery and clarify the possible origins of the issues.

2. Mechanism of Na-CO2 Electrochemistry

In 2011, Asaoka’s group found that adding CO2 to Li/Na-O2 batteries could increase its discharge capacity and energy density, demonstrating the feasibility of metal–CO2 batteries for the first time [12]. In 2013, Das et al. first designed a rechargeable O2-assisted Na-CO2 battery [13]. Under different partial pressures of CO2, the discharge capacity of rechargeable Na-CO2 (O2) battery with tetraethlene glycol dimethyl ether (TEGDME) and ionic liquid (IL) electrolyte increased by 2.6 times and 2.1 times, respectively (Figure 1b,c), the optimal CO2/O2 ratio of the maximum discharge capacity ranges from 40% to 70% (Figure 1c) [13]. The ex situ Fourier transformed infrared (FTIR) spectroscopy test and X-ray powder diffraction (XRD) analysis showed that Na2CO3 and Na2C2O4 exist simultaneously in TEGDME-based electrolyte; Na2C2O4 is the main discharge product in IL electrolyte. These results proved that the electrolyte solvent may determine the final discharge product by affecting the intermediate stability.
In 2016, Hu et al. first proposed and demonstrated the electrochemical reaction mechanism of rechargeable Na-(pure) CO2 batteries:
3CO2 + 4Na ↔ 2Na2CO3 + C
which is inconsistent with the conjecture proposed by Das that Na2C2O4 or Na2CO3 and CO is obtained in pure CO2 atmosphere [6]. The Na-CO2 battery was designed based on NaClO4/TEGDME electrolyte and treated multi-walled carbon nanotube (MWCNT) cathode, which was successfully cycled 200 times in a pure CO2 atmosphere, and the reaction mechanism was verified by a series of tests including XRD and X-ray photoelectron spectroscopy (XPS). Combined with the research results of Li-CO2 system similar to Na-CO2 system [14], it is reasonable to speculate that the reaction mechanism of rechargeable Na-pure CO2 system is as follows: CO2 molecules directly capture e to form C2O42−, the unstable C2O42− undergoes a two-step disproportionation reaction to form CO32− and C, and finally the discharge products Na2CO3 and C are produced as follows:
2CO2 + 2e → C2O42−
C2O42− → CO22− + CO2
C2O42− + CO22− → 2CO32− + C
CO32− + 2Na+ → Na2CO3
The schematic diagram of Na-CO2 battery structure is shown in Figure 2. Its electrochemical reaction route can be proposed as the following chemical equations:
Cathode reactions: 4Na+ + 3CO2 + 4e → 2Na2CO3 + C
Anode reactions: Na → Na+ + e
Overall reaction equation: 4Na + 3CO2 → 2Na2CO3 + C
In addition, a rechargeable Na-O2 (CO2) battery with an IL-propylene carbonate-based electrolyte supplemented with 10% SiO2 nanoparticles was reported by Archer et al. in 2014 [15]. The discharge product was detected as NaHCO3, while CO2 and O2 were released during charging. Although it was speculated that the H in NaHCO3 might come from the introduction of trace amounts of H2O during electrolyte preparation, there was no sufficient experimental evidence to prove it. Obviously, the exploration of the Na-CO2 reaction mechanism is rather limited and further studies are needed.

3. Cathode Material/Catalysts of Na-CO2 Batteries

The charging and discharging process of Na-CO2 batteries is accompanied by the adsorption and desorption of CO2, as well as the deposition and decomposition of the insulated discharge product Na2CO3 on the cathode surface. Inadequate reaction kinetics is the main obstacle, resulting in large overpotential, poor reversibility, poor rate performance and poor cycling stability, etc. [16]. Therefore, exploring an efficient, stable and low-cost electrocatalyst to facilitate CO2 reduction and carbonate decomposition is one of the key issues in the development of this technology. The following issues should be considered during the research of cathode materials/catalysts [17,18,19]: to realize proper CO2 adsorption properties, reasonably designed porous and macroporous structures are essential elements for designing cathode materials, so as to facilitate Na+ and CO2 diffusion, reduce the activation energy of the rate-controlling step and accommodate the insulation discharge product Na2CO3. In addition, the selection of efficient catalytic materials and the design of rich catalytic sites are the key factors to reduce the overpotential in the discharge process and improve the electrochemical performance. Finally, the factors such as abundant raw material resources, environmentally friendly preparation process and easy preparation are the prerequisites for the practical application of Na-CO2 batteries. To realize the commercialization of Na-CO2 batteries, the exploration of efficient catalytic materials is of broad significance. In the past decades, Na-CO2 batteries have achieved rewarding results, especially in the design of efficient electrocatalysts. In this section, we review and discuss the research progress of cathode catalysts based on their chemical composition and microstructure from the three major categories: carbon and heteroatom-doped carbon materials, metal-loaded composite catalytic materials and single-atom catalysts, and then analyze their structure-activity relationship and the overall performance of Na-CO2 batteries

3.1. Carbon Materials and Heteroatom-Doped Carbon Materials

Carbon materials have been widely used in various electrochemical energy storage devices, especially metal–O2 batteries and metal–CO2 batteries, due to their high electronic conductivity, large specific surface area, stable chemical and electrochemical properties, controllable pore structure and adjustable surface chemistry (defect engineering and heteroatom doping) [20,21,22,23,24]. Super P, Ketjen black (KB), activated carbon, carbon nanotubes, graphene, metal–organic framework materials (MOFs) and nitrogen-doped porous carbon materials have been widely used as cathodes for Li-CO2 batteries, and they are also very suitable for Na-CO2 batteries [13,15,25].

3.1.1. Commercial Carbon Materials

Commercial carbon materials can be used as cathode materials for metal–CO2 batteries due to their good electronic conductivity, large surface area, relatively chemical stability, low cost, as well as their mature and scalable preparation process. Common commercial carbon materials, such as Ketjen black (KB) [12] and Super P [26,27], have been developed as porous cathode materials for Li-CO2 batteries. However, the electrochemical performance of commercial carbon materials is not ideal owing to the inherent defects of low electrical conductivity, small pore volume, relatively small specific surface area and limited active sites [28].
Super P is first applied to the metal–CO2 battery as the cathode material. Notably, the electrolyte plays a key role in determining the battery performance when the electrode materials are the same. As reported by Archer et al. Super P has almost no discharge capacity when it was used as a cathode material in ionic liquid electrolytes [26]; while Yang et al. applied a Super P cathode in combination with an ether-based electrolyte to a Li-CO2 battery, showing a greatly enhanced discharge capacity at 100 mA g−1, reaching 6062 mAh g−1 (further increase in discharge capacity with the addition of Ru metal, Ru@Super P, 8229 mAh g−1) [27]. The reason for the significant difference in discharge capacity was ascribed to the electrolyte solvent could affect the stability of the intermediate discharge products and the formation mechanism of the final discharge products, thus the final discharge products and discharge capacities differed with different electrolytes as reported in the earliest studies on Na-CO2(O2) batteries with Super P as the cathode material and ionic liquid electrolytes or ether-based electrolytes [13]. Therefore, the selection of a suitable electrolyte and cathode material plays a key role in determining the battery performance. Advanced characterization methods can be used to track these intermediate species and theoretical calculations can also further provide sufficient evidence.

3.1.2. Nanocarbon Materials

In addition to the above commercial activated carbon materials, nanostructured carbon materials (such as carbon nanotubes and graphene) are more widely explored in Na-CO2 batteries [6,29,30]. They not only have novel structures, high electronic conductivity and high specific surface area, but also have excellent physical and chemical properties due to their unique quantum size effects and surface chemical states, so they have excellent electrochemical activity [31].
Chen’s group prepared activated multi-walled carbon nanotube (a-MWCNT) cathode with three-dimensional tri-continuous porous structure (Figure 3a), high electronic conductivity and good wettability to electrolyte by boiling MWCNT with TEGDME at 100 °C and coating it on Ni network, which effectively improved its reactivity and reduced electrochemical polarization [6]. The rechargeable Na-CO2 battery assembled with this material as cathode electrode catalyst and ether-based liquid electrolyte has a maximum reversible capacity of 60,000 mAh g−1 at 1 A g−1, can be cycled 200 times and the initial discharge/charge voltage difference is only 0.6 V, and gradually increases to 1.3 V after 200 cycles (Figure 3b). Transmission electron microscopy (TEM) images (Figure 3b) shows grape-like discharge nanoparticles (50 nm) were randomly deposited on the t-MWCNT surface during the discharge process, and the grape-like discharge products disappeared after charging (Figure 4c), clearly indicating that the reaction is extremely reversible. A series of electrochemical test methods such as selected area electron diffraction (SAED), Raman spectroscopy (Figure 3d), XPS (Figure 3e), electron energy loss spectroscopy (EELS) (Figure 3f) similarly demonstrated the discharge products of Na-CO2 batteries as Na2CO3 and C. Subsequently, this group reported solid-state Na-CO2 batteries via acid-treated MWCNTs cathodes and NaF-modified anode [29]. The successful introduction of -COOH and -OH groups (Figure 3h) by acid treatment of MWCNTs before use can improve the adsorption capacity of CO2 and thus the reaction kinetics of CO2 electrochemical reduction was enhanced.
In general, carbon materials have remarkable characteristics such as large specific surface area, rich surface chemical properties, high intrinsic electronic conductivity, high chemical and electrochemical stability and low cost, and are the most commonly used cathode materials in Na-CO2 batteries [17,29]. Although the incorporation of new nanostructured carbon materials such as graphene and carbon nanotubes has greatly improved the discharge capacity of Na-CO2 batteries, their catalytic activity and resulting rate performance and cycling performance are still unsatisfactory, and require effective strategies, such as heteroatom doping to are needed to adjust the microstructure, porosity, defects and surface charge distribution.

3.1.3. Heteroatom-Doped Carbon Materials

It is widely accepted that the activation of CO2 at the early stage of discharge depends mainly on the interaction of CO2 with the chemically inert surface of the carbon material, i.e., the adsorption/desorption of CO2. In the field of Li-CO2 batteries, it has been demonstrated that heteroatom doping can modulate the charge distribution of nanostructured carbon material, strengthen nearby positively charged carbon atoms [32]. This has a significant impact on the adsorption mode and adsorption energy between the gas molecules (CO2) and the carbon material surface, thus significantly enhancing the gas reduction kinetics and promoting the decomposition kinetics of the solid discharge products on the cathode surface of the metal–CO2 batteries [20,33,34].
N element doping is an important strategy for catalyst design [35]. Due to the high electronegativity of nitrogen, N doping can break the charge neutrality of the carbon skeleton and cause charge redistribution, which gives the material excellent electronic and catalytic properties. For example, Hu et al. obtained nitrogen-doped carbon cathode nanomaterials with unique structures by calcining zeolite imidazolium salt framework (ZIF-8) at a certain temperature and then washing with dilute hydrochloric acid [25]. This material has higher CO2 absorption and CO2 adsorption properties, as well as better cycling stability. As shown in Figure 4b, the CO2 absorption of the nitrogen-doped sample is much higher than that of carbon black, but its SSA is relatively small. According to the calculated results (Figure 4d), the surface of N-doped has stronger interaction with CO2 bond, which can promote the reduction in CO2 and the formation of discharge products. The electrochemical impedance spectrum shows that the resistance of the cell with the optimized nitrogen-doped nanocarbon (NC900) cathode increases only slightly after 80 cycles, which is much better than that of the cell with the carbon black cathode (about a six-fold increase after seven cycles), as shown in Figure 4c. Their assembled solid-state Na-CO2 batteries Na || liquid-free PEO-based polymer electrolyte || optimized nitrogen-doped nanocarbon (NC900) cathode material exhibited better electrochemical performance: lower overpotential at 50 °C, higher discharge capacity of 10,500 mAh g−1, the energy density of 180 Wh kg−1 and stable cycling for 320 h (at a capacity of 1000 mAh g−1), as shown in Figure 4a. In addition, X-ray photoelectron spectroscopy (XPS) revealed that the signal of CO32− was detected after the NC900 cathode was discharged and disappeared after charging, confirming the reversible formation and decomposition of Na2CO3 (Figure 4e). These excellent properties are attributed to the efficient catalytic effect of porous and highly conductive N-doped nanocarbons. Furthermore, the optimized nitrogen-doped nanocarbons facilitate the formation of sheet-like discharge products, which are easily decomposed into CO2 after charging.
In 2019, Sun’s group prepared nitrogen-doped single-walled carbon nanotubes (N-SWCNH) catalytic materials with a unique structure [36]. The excellent electrocatalytic performance of metal-free N-SWCNH for CO2 reduction is mainly attributed to the unique structure of single-walled carbon nanotubes and nitrogen doping. The porous nature and unique conical structure of SWCNH provide sufficient storage space for discharge products; the highly dispersed nitrogen doping provides a large number of structural defect sites for CO2 adsorption and electron transfer, which contributes to the electron affinity and CO2 adsorption/desorption ability and improves catalyst activity and reversibility. They successfully prepared refillable hybrid Na-CO2 batteries using N-doped single-walled carbon nanotubes (N-SWCNH) as the cathode catalyst and Na superionic conductor (NASICON) solid electrolyte as the separation medium for the hybrid electrolyte system. The use of aqueous electrolyte is also beneficial to the dissolution of discharge products, greatly improving the electrochemical reaction kinetics. As shown in Figure 4f,g, compared with the highly promising N-MWCNTs and Au NPs catalytic materials, N-SWCNH exhibited better performances and the prepared Na-CO2 hybrid battery not only exhibited a low discharge/charge voltage difference of 0.49 V at a current density of0.1 mA cm−1 (Figure 4f). It provides a high discharge capacity of 2293 mAh g−1 at a cut-off voltage and a current density of 0.2 mA cm−2, as shown in Figure 4g. It can be cycled for more than 100 times at a current density of 0.1 mA cm−1 (Figure 4h).
Figure 4. (a) Schematic of the structure of an all-solid-state Na-CO2 cell with N-doped carbon cathode. (b) Low-pressure CO2 adsorption isotherm of N-doped nanocarbon and carbon black at 273 K. The adsorption amount at 1 atm is shown in parentheses. (c) Nyquist plots of the NC900 cathode and carbon black cathode of the Na-CO2 batteries in different charge states. (d) Density function theory (DFT) calculations of the interaction of undoped, graphitic N-doped and pyridine N-doped nanocarbon with CO2 molecules and the adsorption of one CO2 molecule on undoped and doped nanocarbon binding energy calculations. (e) XPS characterization of the discharged products. Reproduced with permission from [25]. Copyright 2020, American Chemical Society. (fh) Electrochemical performance of Na-CO2 batteries. (f) Discharge–charge voltage curves of the cells with N-MWCNTs, AuNPs and N-SWCNH as catalysts at 0.1 mA cm−2 current density. (g) Discharge capacity curves at a current density of 0.2 mA cm−2. (h) Rate performances at different current densities. Reproduced with permission from [36]. Copyright 2020, Elsevier Ltd.
Figure 4. (a) Schematic of the structure of an all-solid-state Na-CO2 cell with N-doped carbon cathode. (b) Low-pressure CO2 adsorption isotherm of N-doped nanocarbon and carbon black at 273 K. The adsorption amount at 1 atm is shown in parentheses. (c) Nyquist plots of the NC900 cathode and carbon black cathode of the Na-CO2 batteries in different charge states. (d) Density function theory (DFT) calculations of the interaction of undoped, graphitic N-doped and pyridine N-doped nanocarbon with CO2 molecules and the adsorption of one CO2 molecule on undoped and doped nanocarbon binding energy calculations. (e) XPS characterization of the discharged products. Reproduced with permission from [25]. Copyright 2020, American Chemical Society. (fh) Electrochemical performance of Na-CO2 batteries. (f) Discharge–charge voltage curves of the cells with N-MWCNTs, AuNPs and N-SWCNH as catalysts at 0.1 mA cm−2 current density. (g) Discharge capacity curves at a current density of 0.2 mA cm−2. (h) Rate performances at different current densities. Reproduced with permission from [36]. Copyright 2020, Elsevier Ltd.
Batteries 09 00036 g004
In addition to N element doping, density function theory (DFT) calculations show that F doping and B doping also improve the performance [37]. In addition, multi-element co-doping, N/S co-doping [38] and N/B co-doping [39] are also used. Cheng et al. adopted the S/N co-doping strategy to improve the physical and chemical properties of the cathode material [40]. They designed electrophilic S vacancies and nucleophilic N-doped active centers on the surface of ReS2, and used the synergistic coupling effect between heteroatoms to adjust the interaction of the catalyst with Li atoms and C/O atoms to show suitable adsorption during charging and discharging processes of Li-CO2 batteries, respectively, thus reducing the activation energy of the rate-determining step and thus increasing the reaction rate.
In summary, the design of three-dimensional (3D) structures and heteroatom doping is an effective strategy to facilitate the diffusion of reactants and improve the catalytic activity of carbon nanomaterials for CO2 reduction and Na2CO3 decomposition. However, the optimal design for carbon materials cannot yield the desired high catalytic activity. Therefore, an in-depth analysis of the active center of heteroatom doping and the related CO2 reduction mechanism with the help of advanced characterization is beneficial for exploring the methods to achieve uniform doping and controlled preparation.

3.2. Metal-Loaded Composites

Despite the demonstrated applicability of nitrogen-doped carbon materials, the catalytic activity for the reversible reaction of the cathode in Na-CO2 batteries is quite limited [41]. Therefore, various carbon loaded metal composites were designed as cathodes for activity modulation and thus to improve the performance of Na-CO2 batteries. The results show that metal-loaded composite catalysts appear to be a good choice for reducing the battery charging potential and improving the electrochemical performance of the battery compared to pure carbon materials [38,42,43,44,45,46,47].

3.2.1. Precious Metals and Their Composites

Due to the inherent electronic configuration of half-filled anti-bonds and high electrical conductivity of noble metals, noble metals and their composites usually have the advantages of good adsorption, low resistance and low overpotential as electrocatalysts [2,48].
Among all precious metal elements, ruthenium (Ru) is one of the most widely studied precious metal catalysts [27,49]. For example, Guo et al. obtained Ru@KB composites by in situ reductions of RuCl3 on porous Ketjen Black [16]. As shown in Figure 5a–c, compared with pure KB as the cathode of Na-CO2 battery, Ru@KB can significantly improve the discharge capacity up to 11,537 mAh g−1, cycle stability for more than 130 cycles and coulomb efficiency of 94.1% of Na-CO2 battery. Their electrochemical performance and ex situ characterization confirmed that ruthenium nanoparticles can significantly reduce the charging overpotential, promote the reversible reaction between Na2CO3 and carbon, and further improve the cycling stability of Na-CO2 batteries. Besides the excellent catalytic activity of metallic Ru, its corresponding oxide RuO2 has also been shown to have significant catalytic ability for reversible metal–CO2 batteries. In addition, Pt [50], Ag, Ir [51,52], Au [53], Pd [54] and other noble metals and their oxides have been shown to be very effective in bifunctional catalysis for metal–CO2 batteries, and need further be explored.
Compared with pure carbon-based materials, the composite by loaded with noble metal-based catalysts help to reduce the active energy required for CO2 reduction and decomposition of discharging products, exhibit excellent catalytic activity in facilitating the electrochemical reaction of Na-CO2 batteries, allowing for smaller overpotentials and higher energy efficiency, and significantly increase the discharge/charge capacity of the batteries. Unfortunately, the expensive cost and limited resources seriously hinder the commercial application of noble metal-based catalysts. Therefore, reasonable strategies such as developing noble metal-based single-atom catalysts or other inexpensive and abundant transition metal-based catalysts should be explored to meet their industrial development.

3.2.2. Transition Metals and Their Composites

Transition metals Ni [21], Co [43,55], Mn [56], Cu [57] and Fe [58] are supported on carbon-based materials with high specific surface area and high electronic conductivity due to their unique adjustable structure and multivalent characteristics, providing rich active sites for electrochemical reactions. In recent years, single metal composites, alloy type composites and transition metal oxide composites of transition metals have been widely reported. Moreover, due to the advantages of rich reserves and low cost of transition metals, transition metal-based composites are a feasible solution for future controlled scale up production [17].
In 2020, Xu et al. obtained an efficient active material (Co/Co9S8 @SNHC) for hybrid system Na-CO2 batteries by anchoring Co/Co9S8 active nanoparticles on biomass-derived S and N-doped graded porous carbon via a microporous/mesoporous domain-limited synthesis strategy, as shown in Figure 5d [45]. The Na-CO2 battery with Co/Co9S8 @SNHC not only exhibits a low overcharge potential of ~0.32 V and a charge/discharge voltage difference of only 0.65 V (Figure 5e). As shown in Figure 5f, it also shows better rate performance, cycle stability (cycled for more than 200 cycles at a current density of 0.1 mA cm−1) and higher specific area discharge capacity (~18.9 mA cm−2). These excellent electrochemical properties are attributed to the meso- and mesoporous-limited domains of the biomass carbon skeleton, which not only effectively inhibit the agglomeration of Co/Co9S8 nanoparticles, but also provide diffusion channels for CO2 and Na+ as well as sufficient space for storing the discharge products. In addition, the effective synergistic interactions between the effective catalytically active sites (Co/Co9S8, C-N, C-S bonds) and the defect-rich carbon interfaces (S, N doping) similarly prevent the agglomeration and separation of Co/Co9S8 nanoparticles, enhance the catalytic activity and improve the stability.
In addition to monometallic composites, bimetallic composites are also potential candidates for CO2 reduction reaction (CO2RR) and CO2 electroreduction reaction (CO2ER). In 2021, Xu et al. obtained bimetallic nitrogen-doped carbon materials of Fe-Cu-N-C with dense bimetallic active sites as catalysts for Na-CO2 batteries with mixed air by introducing Fe3+ and Cu2+ regulated in situ pyrolytic growth of carbon nanotubes via solid-phase reactions [58]. They suggested that the excellent electrocatalytic activity of Fe-Cu-N-C is attributed to the synergy between the N-doped carbon framework with more defects and a large number of active sites in the Fe-Nx, Cu-Nx and Fe/Fe3C nanocrystals. Moreover, Fe3+ is the key to catalyze the conversion of g-C3N4 to CNT conformation, while Cu2+ gives the carbon nanotubes a good structure and uniform diameter. As shown in Figure 5g,h, the Fe-Cu-N-C materials synthesized at a pyrolysis temperature of 700 °C exhibit an ultra-low voltage gap of 0.44 V and a cycle efficiency of 83.2%, as well as a large discharge capacity of 8411 mAh g−1 and a long-term cycling performance of 1550 cycles (over 600 h).
Figure 5. Na-CO2 battery cycling behavior of KB and Ru@KB cathode at 200 mA g−1: (a) KB cathode. (b) Ru@KB composite cathode. (c) Discharge–charge profiles of Na-CO2 batteries with KB and Ru@KB composite cathodes at a current density of 100 mA g−1 in the first cycle. Reproduced with permission from [16]. Copyright 2019, Royal Society of Chemistry. (d) Schematic illustration of catalytic cathode of hybrid Na-CO2 battery. (e) Discharge–charge voltage curves at a current density of 0.2 mA cm−2. (f) Discharge capacity curves of hybrid Na-CO2 batteries with Co/Co9S8@SNHC or SNHC at a current density of 0.5 mA cm−2. Reproduced with permission from [45]. Copyright 2021, Elsevier Ltd. (gi) Electrochemical performances of the as-obtained Fe-Cu-N-C, catalysts: (g) Discharge–charge voltage curves at 0.05 mA cm−2; (h) Discharge–charge cycling curves; (i) Discharge–charge curves of Fe-Cu-N-C. Reproduced with permission from [58]. Copyright 2021, Royal Society of Chemistry.
Figure 5. Na-CO2 battery cycling behavior of KB and Ru@KB cathode at 200 mA g−1: (a) KB cathode. (b) Ru@KB composite cathode. (c) Discharge–charge profiles of Na-CO2 batteries with KB and Ru@KB composite cathodes at a current density of 100 mA g−1 in the first cycle. Reproduced with permission from [16]. Copyright 2019, Royal Society of Chemistry. (d) Schematic illustration of catalytic cathode of hybrid Na-CO2 battery. (e) Discharge–charge voltage curves at a current density of 0.2 mA cm−2. (f) Discharge capacity curves of hybrid Na-CO2 batteries with Co/Co9S8@SNHC or SNHC at a current density of 0.5 mA cm−2. Reproduced with permission from [45]. Copyright 2021, Elsevier Ltd. (gi) Electrochemical performances of the as-obtained Fe-Cu-N-C, catalysts: (g) Discharge–charge voltage curves at 0.05 mA cm−2; (h) Discharge–charge cycling curves; (i) Discharge–charge curves of Fe-Cu-N-C. Reproduced with permission from [58]. Copyright 2021, Royal Society of Chemistry.
Batteries 09 00036 g005
Due to the excellent catalytic activity, low cost and simple preparation methods, transition metal oxides such as NiO [59,60], MnO [37], MnO2 [61], have been extensively studied in the field of Na-CO2 batteries. Fang et al. obtained in situ grown (CMO@CF) composites of Co2MnO4 on carbon fibers by a simple hydrothermal method and high temperature annealing [55] which achieved a reversible charge/discharge process and remained stable after 75 cycles at 200 mA g−1. In addition, XRD, XPS, Raman spectroscopy and SEM characterization demonstrated that the high catalytic activity of CMO@CF electrodes is mainly due to the homogeneous morphological, chemical and structural stability and the hybridized Co2+/Co3+ and Mn2+/Mn3+ redox pairs. ZnCo2O@CNT materials were synthesized by the same hydrothermal method as that of preparing ZnCo2O4 porous carbon nanorod cathodes [43]. The discharge capacity of 500 mAh g−1 at 100 mA g−1 can be charged and discharged stably for at least 150 cycles. Theoretical calculations show that there is a strong adsorption energy for CO2, Na and Na2CO3 on three surfaces of ZnCo2O4, namely the [001] surface of ZnCo2O4, the [111] surface with only Co atoms exposed and the [111] surface with Co and Zn atoms exposed. In addition, the exposed Co atoms on these three surfaces of the density of states (DOS) calculated surface are the real catalytically active sites for the CO2 electrochemical reaction.
Apparently, the transition metal-based composites combine the high SSA and electronic conductivity of porous carbon materials and the unique multivalent characteristics of transition metals, which enable an effective catalytic effect in the application of Na-CO2 batteries, which are considered as the most attractive catalytic cathodes. Therefore, in future research, the catalytic mechanism of transition metal materials should be understood in depth, and more attention should be paid to and design of composite carbon-based materials with synergistic effects of transition metals, transition metal oxides and multi-transition metals, and their controlled and large-scale production should be realized.

3.2.3. Other Types of Composite Materials

In addition to metal nanoparticles loaded on porous carbon materials, several composites have been applied to metal–CO2 batteries, such as molybdenum-based electrode materials [62], metal–organic complexes, polymers and self-supporting freestanding cathodes.
Molybdenum carbide (MoxC) has a d-band electronic structure similar to that of precious metals, especially metallic palladium, and is considered to be a “similar catalyst to precious metals”. In early studies, Mo2C was first applied to lithium–oxygen batteries and was shown to improve the Coulombic efficiency and cycle life of the batteries [63]. After that, Chen’s group prepared Mo2C/CNT composites by a simple carbon thermal reduction method and applied them to Li-CO2 batteries [64]. In this work, the Li2C2O4 intermediate was stabilized by forming a Li2C2O4-Mo2C species on the surface of the catalyst. Finally, amorphous Li2C2O4 final discharge products with thin film morphology were obtained instead of Li2CO3 species. In this case, Mo2C makes the amorphous Li2C2O4 components readily decomposable at very low potentials below 3.5 V. Moreover, it shows excellent cycling properties. However, the specific mechanism of Mo2C catalyzing the formation of amorphous Li2C2O4 remains unclear yet, and some advanced characterization tools as well as theoretical simulation calculations are needed in this regard to strongly elucidate the effects of Mo2C on the performances of CO2 batteries. In addition, MoS2 is also an effective electrocatalyst for the decomposition of Li2CO3 [65]. In fact, we believe that it is a worthwhile effort to study the crystal structure of MoxC and to develop new molybdenum-based catalysts/materials for Na-CO2 batteries.
Metal–organic backbone (MOF) is a crystalline porous organic-inorganic hybrid material with periodic network structure self-assembled by inorganic metal centers (usually metal ions or metal clusters) and bridging organic ligands [24]. Furthermore, MOFs not only combine the stiffness of inorganic materials with the flexibility of organic materials, but also have a porous structure and a metal–nonmetal structure, offering attractive prospects in the field of metal–air batteries research [66,67,68]. Successful applications of MOFs have been reported for Li-CO2 batteries [24] and Na-CO2 batteries [25]. The results show that MOFs have significant potential to increase the discharge capacity and reduce the polarization potential of Li-CO2 batteries. However, the electronic conductivity of MOF is still insufficient, and the catalytic efficiency of the cathode needs to be further improved. We believe that the use of conductive MOFs with high electronic conductivity, along with metal ion doping to synergistically promote catalytic activity, can be attempted. In addition, MOF-derived single-atom catalysts are expected to be a promising new class of MOF-based materials in the laboratory due to their excellent activity in ORR [46,69].
Due to the insulator nature of discharge product, Na2CO3, of the Na-CO2 batteries higher external voltages are typically required to decompose Na2CO3 during the charging process. However, high charging voltages typically lead to degradation of the battery components (e.g., electrolytes, carbon additives and binders), resulting in a decrease in overall battery performance [14]. Therefore, the development of carbon-free and binder-free cathodes is also an important task.

3.3. Single-Atom Catalysts

Large-scale applications of precious metals are hindered by their high price and scarce resources. Non-precious metal materials are promising candidates for noble metals, but with lower catalytic performance. In order to reduce the use of noble metals and further enhance the effective activity of metal atoms, single-atom catalysts (SACs) have emerged and attracted extensive attention [70]. Compared with nanomaterials, SACs can maximize the use of metal atoms; in addition, due to quantum confinement effects, these sub-nanometer catalysts have unexpected catalytic properties and can be used for various energy conversion reactions, and their applications in metal–CO2 batteries are also highly anticipated [71,72,73]. Despite these advantages, SACs still suffer from some drawbacks, such as synthesis and characterization difficulties.
When metal particles are dispersed on the support at the atomic/cluster level, the properties of the catalyst, such as surface free energy, unsaturated coordination environment, quantum size effect and metal–support interaction change drastically, and SACs are easily agglomerated during preparation and application, leading to catalyst deactivation (Figure 6) [69]. Theoretically, increasing the loading of single-atom catalysts and avoiding atomic agglomeration are mainly achieved by increasing the surface area of the carrier and enhancing the interaction of the metal carriers [73]. Researchers have obtained “bottom-up” and “top-down” synthesis strategies for SACs [74]. Typical “bottom-up” strategies include atomic layer deposition (ALD), wet chemical methods, electrochemical deposition and chemical vapor deposition [75,76,77,78]. Atomic layer deposition (ALD) allows for precise control of the film thickness by evaporating the active precursor into a vapor and depositing it layer by layer on a carrier in a self-limiting manner using sophisticated equipment [79,80]. Wet chemical methods consisting of wet impregnation and co-precipitation are considered a promising technique for the synthesis of SACs because of their simplicity and the possibility of mass production of SACs [81]. However, it should be noted that the SACs obtained by this strategy usually exhibit the disadvantages of low metal loading and easy aggregation. In addition, electrochemical deposition has proven to be a universal method for the preparation of single-atom catalysts. Single-atom catalysts on various substrates such as transition group metals and oxides, sulfides, selenides and carbon materials were obtained using electrochemical deposition by Zeng et al. [82]. In contrast to the bottom-up strategy, top-down strategies are generally ball milling, high-temperature atom capture and pyrolysis [83,84,85]. Ball milling is a simple and versatile method that enables large-scale preparation by breaking/reconstructing the bonding, thus efficiently producing SA [83]. The pyrolysis strategy often uses various metal–organic complex precursors with molecular confinement and organic ligand coordination characteristics to obtain uniformly dispersed M-N-C site catalysts by pyrolysis, which are considered to be the most promising non-precious metal-based ORR catalysts [86]. Despite the consumption of organic reagents, it shows clear advantages: simpler preparation process, controllable tuning of the performance of SACs by controlling the pyrolysis environment and inducing defects [87,88].
With the development of nanotechnology and characterization science, advanced characterization techniques provide rich information and reliable evidence for studying the composition, structure and performance relationships of catalysts, which plays a crucial role in the rapid development of single-atom catalysts [89,90]. In addition, theoretical calculations can provide a reasonable reaction model for uniformly dispersed active sites and calculate the activation energy of the reaction. Combining advanced characterization techniques and data analysis of theoretical calculations, detailed information on the geometry and electronic structure of the active sites of single-atom catalysts can be obtained, thus elucidating the structure and performance relationships of single-atom materials in the catalytic process and guiding the design of single-atom catalysts [91,92]. The contents of characterization technology and theoretical calculation will be elaborated in the following section.
SACs are a promising redox electrocatalyst. However, the application of SACs in Na-CO2 batteries is relatively new and much research is needed to explore the potential of SACs for practical applications in the next generation of rechargeable batteries. Theoretically, SACs ensure 100% atom utilization efficiency, which is very valuable for saving metal resources, especially precious metals. Recently, Zhu et al. successfully prepared single-atom Pt deposited on nitrogen-doped carbon nanotubes (Pt@NCNT) as a cathode material for Na-CO2 batteries [44]. The single-atom dispersed Pt catalytic site, with a unique electronic structure and low coordination environment (Figure 7a,b), can achieve high activity and high selectivity due to the well-dispersed and fully exposed active site. Compared with the air batteries with pure NCNT as the positive electrode, the Pt single-atom catalyst can effectively improve the discharge reaction rate. The schematic diagram of Figure 7c illustrates the reaction mechanism based on Pt-SA in the discharge/charge electrochemical process of Na-CO2 battery.
Given the two important factors of catalyst performance and cost, non-precious metal SACs with similar activity to noble metal SACs are more attractive [93,94]. Future work should target the combination of nitride or carbide supports with non-precious metal single atoms, which may provide unique electronic interactions with the metal and thus improve anode electrochemical catalytic performance [95]. In addition, SACs applied to Na-CO2 batteries require harsh operating conditions, and the development of industrial fabrication methods for low-cost, stable, highly metal-loaded SACs is critical.

4. Progress of Solid Electrolyte and Interface Research

Many Na-CO2 batteries reported still use the conventional organic liquid electrolyte [6,43], because Na-CO2 batteries operate in a semi-open system, there are great safety hazards in actual application, such as the inherent volatility and flammability of organic solvents, and potential problems such as reduced battery life due to interfacial side reactions under high pressure [7]. Replacing the liquid electrolyte with a solid electrolyte can not only effectively solve the above safety hazards, but also inhibit the growth of dendrites and prevent CO2 from corroding the metal Na anode during the long-term cycling process [11]. In addition, through scientific and reasonable structural design, solid electrolyte can also provide wider electrochemical window, higher energy density and longer cycle life [96,97,98].
For the solid electrolyte of the Na-CO2 batteries, in addition to meeting the necessary energy requirements for a solid sodium ion conductor, e.g., high ionic conductivity, non-conductivity, interface compatibility, simplicity of preparation, low cost and environmental friendliness, it needs to be highly resistant to superoxide and peroxide to facilitate the smooth Na-CO2 reaction [11]. It is difficult for the reported sodium ion solid electrolyte for Na-CO2 batteries to meet all the above requirements at the same time, further research is needed [99,100]. In this section, we provide a brief overview of several important applications of solid electrolyte in the Na-CO2 batteries, and analyze the research on solid electrolyte interface. Sodium-based electrolytes can be classified into three categories according to the chemical composition of the electrolyte and the transport mechanism of sodium ion, namely inorganic solid electrolytes (ISEs), polymer electrolytes (PEs) and composite polymer electrolytes (CPEs).

4.1. Inorganic Solid Electrolytes

Inorganic solid electrolytes (ISEs) are usually characterized by high ionic conductivity (10−5~10−2 S cm−1), high ionic mobility number, mechanical rigidity, non-flammability and non-flow characteristics due to their structural properties [101]. In order to understand the reasons for the high ionic conductivity of ISEs and to further improve its ionic conductivity, it is necessary to understand its ionic transport mechanisms and properties. The three principal migration mechanisms shown in Figure 8a are vacancy leap, gap-site leap and linkage leap [102]. According to the mechanism of vacancy leap and gapsite leap ion transport, the ionic transport is mainly related to the activation energy of the material and the number of defects in the vacancy, so ion doping can effectively improve the ionic conductivity of the material [103,104,105]. The ionic transport in linkage leap is not directly jumping between vacancies, but knocking the ions of adjacent sites to make them migrate to vacancies, so the ionic transport potential of linkage leap mechanism is lower than that of vacancy or gap site leap, so increasing the concentration of sodium ions in ISEs can improve the ionic conductivity [106,107,108,109]. ISEs can be classified according to their components and crystal structure types: NASICON, Na-beta-Al2O3, sulfide electrolytes and complex hydride electrolytes.

4.1.1. NASICON Structured Electrolytes

Na1+xZr2SixP3−xO12 (0 ≤ x ≤ 3) was proposed by Hong and Goodenough in 1976 and consists of NaZr2(PO4)3 and Na4Zr2(SiO4)3 solid solution; the NASICON-structured fast ion conductor material with a 3D transport channel for Na+ was obtained by partially replacing the pentavalent P in NaZr2(PO4)3 by the tetravalent Si and introducing Na+ to balance the charge [110,111]. Among them, Na1+xZr2SixP3−xO12 (0 ≤ x ≤ 3) fast ionic conductors are CO2-stable ISEs. Due to its high ionic conductivity, good thermal and chemical stability, Na Superionic conductor (NASICON) structure electrolytes is the most widely studied inorganic material in the solid electrolyte of Na-CO2 batteries [36,42,112,113]. Kim et al. [114] demonstrated that after NASICON materials contact with water, H3O+ occupies part of the Na+ sites, causing a shift in the NASICON peak position, as revealed by XRD analysis. Although NASICON reacts with water, it does not cause the fracture and decomposition of the NASICON ceramic sheet, and there is no drastic decay in ionic conductivity.
According to the crystal structure analysis of this series of materials by Hong (Figure 8b) [111], the crystal structure of the material is monoclini with space group C2/c when 1.8 ≤ x ≤ 2.2 at room temperature; when x is outside this range, the crystal structure becomes tripartite (rhombic structure) with space group R-3c; with the change in temperature, the two crystal structures can be converted, and the phase transition temperature depends on the specific components of the material, usually at 150~200 °C. In the triangular structure (R-3c), SiO4 or PO4 tetrahedra are connected to ZrO6 octahedra at the top corners, forming the three-dimensional skeleton for Na+ transport. Therefore, the most commonly used method to increase the ionic conductivity is doping. Heterogeneous element doping is used to enhance ionic conductivity by changing the local structure of the crystal, making the Na+ transport channel larger and lowering the ion diffusion potential barrier [107]. Among Na1−xZr2SixP3−xO12 (0 ≤ x ≤ 3), Na3Zr2Si2PO12 (x = 2) with the monoclinic phase structure is more stable. It exhibits the highest ionic conductivity of ~10−4 and ~10−1 S cm−1 at room temperature and 300 °C, respectively [110]. As shown in Figure 8c, Song et al. investigated the crystal structure and ionic conductivity of NASICON doped with different contents of alkaline earth metal ions (Mg, Ca, Sr, Ba) [105]. It was found that Mg ion doping improved its conductivity most significantly, and the ionic conductivity of Na3.1Zr1.95Mg0.05Si2PO12 could reach 3.5 × 10−3 S cm−1 at room temperature. However, Ba2+ ion substituted compounds by the elements with larger ionic radii exhibit narrower bottlenecks than the original compounds, suggesting that the ionic radius of the substituent plays an important role [105]. Lu et al. prepared Na3.2Zr1.9Mg0.1Si2PO12 matrix materials by doping Mg2+ with Zr sites, showing an ionic conductivity of 1.16 mS cm−1 at room temperature in order to obtain high-performance Na-CO2 battery [113] (Figure 8d). In addition, heterogeneous elements with similar ionic radii to Zr (including calcium ions [115] and aluminum ions [116]) doped NASICON-based composite electrolytes can significantly improve the ionic conductivity of NZSP by changing the Na+ concentration and crystal localization structure, thus improving the Na+ transport channels [117].
Another way to improve the ionic conductivity of NASICON materials is to reduce the grain boundary resistance by adjusting the grain size and grain boundary chemistry [100]. Ihlefeld et al. investigated the size effect on the grain boundary resistance of Na1+xZr2SixP3−xO12 (0.25 < x < 1.0) [118]. The results showed that changing the Si/P ratio, increasing the process temperature and decreasing the annealing temperature all reduced its grain size and thus the ionic conductivity is increased. Hu et al. synthesized La3+ doped NASICON by adding La(CH3COO)3 to the precursor through a self-growth strategy [103]. XRD results showed that several new phases of Na3La(PO4)2, La2O3 and LaPO4 appeared at the grain boundaries of the final electrolyte material (Figure 8e). The newly appeared phases adjusted the chemical composition of the grain boundaries and increased the ionic conductivity at the grain boundaries.

4.1.2. Na-Beta-Al2O3

Na-beta-Al2O3 has become one of the most studied sodium ion solid electrolytes due to its high ionic conductivity and suitable mechanical properties, and was first applied in solid-state Na-S batteries [119,120,121]. Na-beta-Al2O3 has two crystal structure types as shown in Figure 8f, both are layered structures made of alternating stacks of spinel and sodium-conducting layers, with Na+ conducting in two dimensions between two adjacent spinel stacks, called sodium-conducting layers. one crystal structure of Na-beta-Al2O3 is a hexagonal crystal system structure made of two spinel structures stacked with space group P6m3/mmc, labeled beta-Al2O3, composed of Na2O-(8~11) Al2O3; the other is a tripartite crystal structure, consisting of three spinel structures stacked in the space group R3m, labeled β″-Al2O3, composed of Na2O-(5~7) Al2O3.
A higher proportion of β″-Al2O3 phase is usually desired in the material because the ionic conductivity of the β″-Al2O3 phase is higher than that of the β-Al2O3 phase. This is not difficult to analyze: the two adjacent layers of spinel structure are connected by O2− in the Na+ conducting sodium layer, forming an Al-O-Al bond and Na+ two-dimensional ion transport along the ab plane. For the β-Al2O3 phase, O2− in the Na+ conduction layer has a higher electrostatic gravitational force on the surrounding Na+ and can accommodate a smaller amount of Na+, while for the β″-Al2O3 phase, O2− in the Na+ conduction layer has a lower electrostatic gravitational force on the surrounding Na+ and can accommodate more Na+, thus the ionic conductivity of the β″-Al2O3 phase is higher than that of the β-Al2O3 phase. Single-crystal β″-Al2O3 can exhibit ionic conductivity up to 1 S cm−1 at 300 °C, while polycrystalline structures exhibit 0.002 S cm−1 at room temperature and 0.2–0.4 S cm−1 at 300 °C. The lower ionic conductivity exhibited by polycrystalline structures is due to the high grain boundary resistance in polycrystalline β″-Al2O3. However, pure β″-Al2O3 is a thermodynamically sub-stable phase, which decomposes into Al2O3 and β-Al2O3 at 1500 °C and has relatively poor mechanical properties (200 MPa). Na-beta-Al2O3 is generally prepared by solid-phase method; however, the solid-phase synthesized Na-β″-Al2O3 powder has residual NaAlO2 at the grain boundaries of the β″-Al2O3 phase and β-Al2O3 phase, which is unstable in air and easily reacts with CO2 and H2O [122,123]. Virkar et al. used steam-assisted method to obtain dense ceramic flakes by sintering Y-ZrO2 and α-Al2O3 at 1450 °C to enhance their chemical stability [124].
The research of this material focuses on achieving stabilization in CO2 atmosphere and reducing the impedance of grains and grain boundaries by suitable preparation methods, enhancing the ratio of β″-Al2O3 and reducing unwanted by-products. Doping is usually used to stabilize the β″-Al2O3 phases, such as Li+, Mg2+, Ni2+ and Ti4+; in addition, the overall mechanical strength can be enhanced by synthesizing mixed crystals of β″-Al2O3 and β-Al2O3 or by adding zirconium oxide [125].

4.1.3. Sulfide Electrolytes

Compared with oxide solid electrolytes, sulfides as electrolyte materials have higher ionic conductivities and lower grain boundary resistances, which is due to the intrinsic characteristics of sulfur [126]. S is less electronegative than O, which is less binding to Na+ and facilitates the free movement of Na+, thus it has lower grain boundary resistance; S has a larger ionic radius than O, and S replaces O to expand the lattice structure and form channels that facilitate the diffusion of Na+ and thus has a high ionic conductivity [106,127]. In addition, sulfide electrolytes also have the advantages of mild synthesis conditions, good mechanical strength, good ductility, etc. However, sulfide is unstable in humid air, easy to absorb water and easy to decompose with water in air, releasing toxic H2S gas [128,129,130]. Therefore, it is necessary to improve the ionic conductivity and air stability by designing new sulfide-based electrolytes for the application in metal–CO2 batteries.
Na3PS4 has two crystal structures, the tetragonal phase (P-421c, a = b = 6.9520 Ǻ, c = 7.0757 Ǻ) and the cubic phase (I-43m, a = b = c = 7.0699 Ǻ) [131,132]. The Figure 8g shows that in the cubic phase, Na+ is distributed in two distorted tetrahedral interstitial sites with space group I-43m, while in the tetragonal phase, Na+ is distributed in a tetrahedral site and an octahedral site with space group P-421c. Typically, Na3PS4 exists as a tetragonal phase, which can be transformed into a cubic phase at about 530 K. The cubic phase monomer has the lowest activation energy and has been studied the most. In 2012, the Hayashi and colleagues reported glass-ceramic sulfide electrolytes with an ionic conductivity of 2 × 10−4 S cm−1 at room temperature, and the high ionic conductivity can be attributed to the stability of cubic Na3PS4 in microcrystalline glass electrolytes at room temperature [126]. It was found that the ionic conductivity of several crystalline phases of Na3PS4 sulfide electrolytes, has the following pattern: ionic conductivity of the cubic phase > ionic conductivity of the tetragonal phase > ionic conductivity of the glass-ceramic phase > ionic conductivity of the glass phase. Therefore, how to obtain a stable cubic phase Na3PS4 crystal structure at room temperature is one of the methods to improve the ionic conductivity of sulfide solid electrolytes.
Tuning the size of unit cell/channel by introducing Na vacancies, gaps or modulating the interaction between Na+ and anion skeleton in the lattice by means of elemental doping is an important method to improve the ionic conductivity of sulfide electrolytes. The doping of P sites with homologous As5+ of larger ionic radius expands the lattice and introduces Na vacancies while increasing the distance between Na-S [128,133]. In the tetragonal phase of Na3PS4, replacing S2− with negative monovalent F, Cl, Br and I can also introduce Na vacancies according to the charge balance theory, thus increasing the migration probability of Na+ from one site to neighboring sites and thus increasing the ionic conductivity [134,135]. De Klerk and Wagemaker investigated the effect of Na+ vacancies on the ionic conductivity of Na3PS4 by molecular dynamics (MD) simulations: a significant increase at 300 K compared to pure cubic Na3PS4 (0.17 S cm−1) [136]. Experimental results show that the ionic conductivity of chlorine-doped tetragonal sulfide (Na2.9375PS3.9375Cl0.0625) at 303 K is 1.14 × 10−1 S cm−1, which is much higher than that of the pristine tetragonal Na3PS4 (5 × 10−5 S cm−1) [135]. The first-principles study shows that the Na+ gap defect structure can also increase the Na3PS4 carrier density, i.e., in the cubic phase of Na3PS4, doping with tetravalent ions M4+ (M = Si, Sn, Ti, Ge) replaces the P5+ sites, while more Na+ will be introduced to maintain the electroneutrality and broaden the Na+ channel size, thus reducing the gap migration barrier and increasing its ionic conductivity [127]. The silica sulfide doped microcrystalline glass electrolytes (94Na3PS4·6Na4SiS4) also show higher ionic conductivity (7.4 × 10−4 S cm−1) compared to the undoped state [120]. In addition, the interaction between Na+ and the anionic backbone has a significant effect on the ionic conductivity of the electrolyte. Theoretical calculations show that Se doping can increase the polarizability of the anionic framework, smooth the lattice and lowering the activation potential barrier [137].
However, to apply sulfide electrolytes in metal–air batteries, it is important to make them stable in air. In 2016, Liang et al. proposed an air-stable sulfide electrolyte [138]. When the P-site of Na3PS4 is completely replaced by Sb to obtain Na3SbS4, the unit cell and channels are expanded to further improve the ionic conductivity of Na3PS4, exhibiting a high ionic conductivity of 1 mS cm−1 at 25 °C and a good compatibility with sodium metal anodes. More importantly, Na3SbS4 is stable when exposed to O2 and H2O, which can be well explained by the theory of soft and hard acids and bases [138,139]. In Na3PS4, the interaction between O2− and P5+ (hard acid) is stronger, so S2− (soft base) is easily replaced by O2− (hard acid), while in Na3SbS4, Sb5+ (soft acid) and S2−. The interaction between them is strong and not easy to be destroyed by O2. When exposed to air, pure Na3SbS4 tends to form Na3SbS4-xH2O and H2O may reversibly lose after heating at 150 °C for 1 h (Figure 8h). Despite the attractive stability and ionic conductivity of Na3SbS4 (~10−3 S cm−1), the toxicity of Sb should be considered. At present, the research on sodium sulfide ionic solid electrolytes is still at the early stage and further studies are needed to improve their ionic conductivity, chemical stability and electrochemical stability.

4.1.4. Complex Hydride Electrolytes

In addition to the inorganic solid electrolytes mentioned above, complex hydride electrolytes are also good conductors of Na ions. In 2012, Orimo and colleagues first reported complex hydrides as Na+ ion solid electrolytes [140]. The reported complex hydride electrolytes have high ionic conductivities and ion mobility numbers, but complex hydrides are usually prone to water absorption, difficult to maintain stably in air. The electrochemical window is also not meet the requirement. If its chemical stability can be improved by modification, it may provide a new solution for the application of Na-CO2 batteries.
Complex hydrogen compounds consist of a metal cation Na+ and a complex anion consisting of a central atom and a ligand hydrogen atom, which have been reported as Na2(BH4)(NH2) [141], Na2B12H12 [142,143], NaCB11H12 [144], NaCB9H10 [145], Na2(B12H12)0.5(B10H10)0.5 [146,147], Na2(CB9H10)(CB11H12) [144], Na2B10H10 [148] and Na3OBH4 [149], etc. For example, Na2B12H12 exhibits high ionic conductivity (>0.1 S cm−1 at 573 K) due to its high-temperature disordered bulk-centered cubic phase (cation vacancy rich structure). However, due to the high phase transition temperature, the complex hydride electrolytes of large anions cannot meet the requirements of practical applications in different fields. It is necessary to reduce or eliminate the phase transition temperature of complex hydrides. It has been shown that the phase transition temperature can be significantly lowered after the introduction of C by chemical modification of the anion [145,150], for example, NaCB11H12 vs. Na2B12H12: 380 and 529 K; NaCB9H10 vs. Na2B10H10: 290 and 380 K, with ionic conductivity of 7 × 10−2 S cm−1 at room temperature (Figure 8i). In addition, mixing different anions helps to reduce or eliminate the transition temperature due to the introduction of geometric hindrance [140]. Tang and colleagues found that the transition temperature can also be effectively reduced by ball milling by reducing the grain size and disordering, and that a more homogeneous mixing of atoms is less prone to phase transitions [151]. Further studies should focus on investigating the stability and ionic conductivity of composite hydrides and applying them to solid-state Na-CO2 batteries.
Figure 8. (a) Arrows indicate the three typical migration mechanisms: vacancy site leap, gap site leap and linkage leap. The circles indicate cations in stable (green) and substable (orange) positions in the model lattice. The dashed lines indicate the transition states of cation jumps imposed by the anion skeleton (not explicitly shown). Reproduced with permission from [102]. Copyright 2019, Springer Nature. (b) Crystal structures of representative NASICON (Na3Zr2Si2PO12) with rhombic and monoclinic crystalline phase. Reproduced with permission from [111]. Copyright 1976, Elsevier Ltd. (c) Lattice parameter, volume of unit cell and area of the bottleneck (marked as T1 in the inset) of Na3Zr2Si2PO12 and Na3.1Zr1.95M0.05Si2PO12 (M = Mg, Ca, Sr, Ba). Reproduced with permission from [105]. Copyright 2016, Springer Nature. (d) Simplified view of the structure of Na3.2Zr1.9Mg0.1Si2PO12 approximately along the [101] direction. Reproduced with permission from [113]. Copyright 2022, John Wiley and Sons. (e) X-ray diffraction patterns. The X-ray diffraction patterns of the nominal composition Na3+xLaxZr2−xSi2PO12 (x = 0, 0.05, 0.10, 0.15, 0.20, 0.25, 0.30, 0.35, 0.40) solid electrolytes. The peaks in the circled by the dotted box represent the Na3La(PO4)2 minor phase. The peaks marked by the black solid spheres and black asterisks denote the La2O3 and LaPO4, respectively. Reproduced with permission from [103]. Copyright 2017, John Wiley and Sons. (f) Crystal structures of β-Al2O3 and β″-Al2O3. Reproduced with permission from [100]. Copyright 2018, Elsevier Ltd. (g) Crystal structures of Na3PS4 with cubic and tetragonal phase. Reproduced with permission from [131]. Copyright 2017, Reproduced with permission from. (h) XRD patterns of pristine Na3SbS4·9H2O, as-prepared Na3SbS4, air-exposed Na3SbS4 and reheated air-exposed Na3SbS4 (150 °C for 1 h under vacuum). Reproduced with permission from [138]. Copyright 2016, John Wiley and Sons. (i) Relative geometries of the B12H122−, B10H102−, CB11H12 and CB9H10 anions. Reproduced with permission from [148]. Copyright 2014, John Wiley and Sons.
Figure 8. (a) Arrows indicate the three typical migration mechanisms: vacancy site leap, gap site leap and linkage leap. The circles indicate cations in stable (green) and substable (orange) positions in the model lattice. The dashed lines indicate the transition states of cation jumps imposed by the anion skeleton (not explicitly shown). Reproduced with permission from [102]. Copyright 2019, Springer Nature. (b) Crystal structures of representative NASICON (Na3Zr2Si2PO12) with rhombic and monoclinic crystalline phase. Reproduced with permission from [111]. Copyright 1976, Elsevier Ltd. (c) Lattice parameter, volume of unit cell and area of the bottleneck (marked as T1 in the inset) of Na3Zr2Si2PO12 and Na3.1Zr1.95M0.05Si2PO12 (M = Mg, Ca, Sr, Ba). Reproduced with permission from [105]. Copyright 2016, Springer Nature. (d) Simplified view of the structure of Na3.2Zr1.9Mg0.1Si2PO12 approximately along the [101] direction. Reproduced with permission from [113]. Copyright 2022, John Wiley and Sons. (e) X-ray diffraction patterns. The X-ray diffraction patterns of the nominal composition Na3+xLaxZr2−xSi2PO12 (x = 0, 0.05, 0.10, 0.15, 0.20, 0.25, 0.30, 0.35, 0.40) solid electrolytes. The peaks in the circled by the dotted box represent the Na3La(PO4)2 minor phase. The peaks marked by the black solid spheres and black asterisks denote the La2O3 and LaPO4, respectively. Reproduced with permission from [103]. Copyright 2017, John Wiley and Sons. (f) Crystal structures of β-Al2O3 and β″-Al2O3. Reproduced with permission from [100]. Copyright 2018, Elsevier Ltd. (g) Crystal structures of Na3PS4 with cubic and tetragonal phase. Reproduced with permission from [131]. Copyright 2017, Reproduced with permission from. (h) XRD patterns of pristine Na3SbS4·9H2O, as-prepared Na3SbS4, air-exposed Na3SbS4 and reheated air-exposed Na3SbS4 (150 °C for 1 h under vacuum). Reproduced with permission from [138]. Copyright 2016, John Wiley and Sons. (i) Relative geometries of the B12H122−, B10H102−, CB11H12 and CB9H10 anions. Reproduced with permission from [148]. Copyright 2014, John Wiley and Sons.
Batteries 09 00036 g008

4.2. Polymer Electrolytes

The concept of polymer electrolytes (PEs) was first proposed in 1973 when Fenton et al. discovered that alkali metal salts dissolved in polyethylene oxide were found to form conductive complexes [152]. Subsequently, Feuillade et al. introduced organic plasticizers into the polymer–salt binary system to obtain a quasi-solid electrolyte [153]. Compared with inorganic solid electrolytes, electrolytes with polymer matrix usually have good flexibility, easy processing and high tolerance to vibration, shock and mechanical deformation and better interfacial contact and compatibility between electrodes and electrolytes.
For the ion mobility number and ionic conductivity of PEs, the methods of increasing amorphous regions and fixing anions are usually used. Soluble metal salts with anion stabilizing effect are generally preferred [11], such as NaCF3SO3, NaPF6 and NaClO4; in addition, a certain percentage of inorganic fillers are also added to enhance the amorphous region and ionic transport of organic electrolytes, thus improving ionic conductivity, mechanical strength and electrochemical properties. Polymer electrolytes are classified into solvent-free polymer electrolytes (SPEs) and gel polymer electrolytes (GPEs) based on their composition and physical form, and then explains the ion transport mechanisms, basic properties and their applications in Na-CO2 batteries [99].

4.2.1. Solvent-Free Polymer Electrolytes

Solvent-free polymer electrolytes (SPEs) are usually consisted of only polymer matrix and lithium salt as solute without adding liquid solvent as plasticizer, and can be easily fabricated by solvent casting, thermoforming or extrusion techniques. SPEs that have been reported for Na-CO2 batteries are mainly of the polyethylene oxide (PEO) type [25,30]. The carbonate electrolytes are susceptible to nucleophilic attack by superoxide radicals and are not suitable for metal–air batteries, while other types of SPEs are still to be developed.
PEO is the earliest and most studied, and its chemical structure is H-(O-CH2-CH2)n -OH, a polyether compound with the advantages of good chemical stability, good compatibility with alkali metal negative electrode, good flexibility, good water solubility, low density, good viscoelasticity and easy film formation, and tolerance to superoxide radicals, which is a more suitable organic material for Na-CO2 battery electrolyte. However, the current PEO-based solid electrolytes during application are often troubled by problems such as: the relatively high degree of room temperature crystallization of PEO (PEO chains are mainly crystallized at 65 °C), resulting in low room temperature ionic conductivity (~10−8 S/cm), thus requiring operation at higher temperatures; low upper limit of electrochemical stability potential (≤4.2 V), thus preventing the use of high voltage cathode materials; poor dimensional thermal stability (softening point of 55–64 °C); low mechanical strength (≤10 MPa).
According to the ionic transport mechanism of PEO-based PEs (Figure 9a), the metal salts dissociate and delocalize metal cation (Li+, Na+), metal ions with polar groups on the polymer chain, such as O and N. Under the action of electric field and above the glass transition temperature, the polymer molecular chain segments in the amorphous region (amorphous region) are able to vibrate, and with the movement of the polymer chain, the metal cations are continuously complexed and dissociated with the groups on the polymer chain segments. Metal cation jumps from one coordination site (usually composed of more than three electron-giving groups) to the next, thus enabling the transfer of metal ions [154]. However, since the movement of chain segments transfers both metal cation and anions, the metal cation migration number of PEs is usually less than 0.5 and their ionic conductivity is generally inferior to that of ISEs [155,156,157].
The following aspects can be used to improve the comprehensive performance of PEO-based polymer electrolytes: (1) to facilitate the migration of metal ions, the polymer-cation interaction must be a compromise between sufficient strength (ensuring salt solubility through cation solventization) and sufficient instability (facilitating ion hopping from one coordination site to another); (2) selecting a polymer matrix with a dielectric constant a polymer matrix with a large dielectric constant, a high dielectric constant facilitates effective charge separation of the metal salt and thus a high Na+ concentration; (3) the higher backbone flexibility and motility of the PEO chain facilitates the segmental movement of the polymer chain and (4) the high molecular weight of the polymer matrix is also desirable to obtain a polymer electrolyte with better mechanical strength.
In 2018, Chen’s team obtained a high-performance PEO/NaClO4/SiO2 all-solid polymer electrolyte by adding inorganic filler nano-SiO2 and assembled an all-solid flexible Na-CO2 batteries with sodium as the anode, PEO/NaClO4/SiO2 as the electrolyte and multi-walled carbon nanotubes as the cathode [30]. In the all-solid polymer electrolyte, PEO acts as a Na ion conductor to transport Na between the Na anode and the gas cathode, and the addition of nano-SiO2 can reduce the crystallinity and promote the dissociation of NaClO4, thus improving the ionic conductivity, in addition to improving its mechanical strength and thermal stability. The highest sodium ionic conductivity of 6.4 × 10−4 S cm−1 and the highest ion transfer number of 0.56 were obtained at an operating temperature of 70 °C when the SiO2 content in the polymer electrolyte was 3 wt% (200 μm thickness). Assembled flexible batteries with good bendability (21,000 times), foldability and shape customizability, as well as in the bending state (0~360°), stable operation time (80 h), high cycling stability (240 cycles with −0.4 V increase in overpotential), high capacity (450 mAh g−1) and high energy density (173 Wh kg−1) were obtained in electrochemical tests. Furthermore, Sun et al. developed a non-aqueous solution of expanded perfluorosulfonic ethylene carbonate and propylene carbonate (EC-PC) with a hexagonal structure of Na0.67Ni0.23Mg0.1Mn0.67O2 as the cathode, and metal sodium as the anode. It proved to be a safe and durable all-solid-state Na-ion battery [158].
The research on SPEs for Na-CO2 batteries is obviously insufficient, and subsequent studies need to further explore and investigate the Na+ transport mechanism of SPEs to elucidate the intrinsic connection between Na+ and polymer matrix, and to further develop polymer electrolyte matrix that can be applied to Na-CO2 battery systems by simple blending, copolymerization, hyperbranching or cross-linking methods to find electrolyte materials with better overall performance.

4.2.2. Gel Polymer Electrolytes

Gel polymer electrolytes (GPEs) is a semi-solid electrolyte in a gel state, consisting of a polymer matrix, electrolyte salt and plasticizer. For GPEs, ion transport is the result of synergistic transport between the solid polymer matrix and the liquid electrolyte, and their room temperature ionic conductivity is higher than that of SPEs due to the presence of a certain amount of solvent; however, the safety in batteries is poorer than that of SPEs due to the presence of organic solvents. In addition, GPEs have both polymeric properties and are more flexible and easier to process than glass-ceramic type solid electrolytes. Therefore, GPEs exhibit good mechanical properties and good compatibility with electrodes. In quasi-solid polymer electrolytes, sodium ion transport occurs mainly in the liquid plasticizer containing dissolved lithium salts, while the polymer matrix provides mechanical strength to the GPEs and keeps it quasi-solid, thus minimizing the safety risk caused by leakage of liquid components. During battery charging and discharging, the plasticizer in the GPEs reacts on the electrode surface to form a solid electrolyte interface (SEI) film, similar to a liquid electrolyte. In contrast, electrochemically inert polymeric matrices are typically not involved in SEI formation.
PVDF-based gel electrolytes are the most widely studied matrix materials for GPEs due to their good film formation, large dielectric constant, high glass transition temperature and strong electron-absorbing groups. In 2017, Hu et al. [159] reported a quasi-solid Na-CO2 battery with an applied PVDF-HFP-4% SiO2/NaClO4 -TEGDME CPE with advantages such as high ionic conductivity (1.0 mS cm−1), strong toughness, non-flammability (automatically extinguished within 1 s even if ignited for 5 s) and operated at a capacity of 1000 mAh g−1, 400 cycles at 200 °C, as shown in Figure 9b,c. The product has shown higher safety and good performance.
In addition to the above PVDF-based gel polymers, PEO-based and cyano polymer-based GPEs are widely used in lithium batteries, and related sodium ion gel electrolytes and their applications in sodium batteries are yet to be developed.
PEO-based GPEs have a high degree of crystallization at room temperature and low ionic conductivity at room temperature, making them difficult to apply in practical batteries. Suitable plasticizers can be added, such as polyethylene glycol [153] or crown ether [160], to reduce the crystallinity and enhance the room temperature ionic conductivity. Co-blending [161], cross-linking [162] or the addition of inorganic fillers to improve the mechanical strength and thermal stability.
Cyano (e.g., polyacrylonitrile) [153] cyano (such as polyacrylonitrile and cyanoethyl polyvinyl ether) is a polar group with a high dipole moment [163], and is a strong electron-absorbing group with a dielectric constant of about 30. Cyano has high oxidation resistance and its introduction into the polymer matrix can increase the oxidative decomposition voltage of the electrolyte. Feuillade et al. [153] prepared polyacrylonitrile electrolytes with ionic conductivity close to 1 × 10−3 S cm−1 at room temperature. In polyacrylonitrile gel electrolytes, the ionic conductivity increases with increasing plasticizer and salt content. However, due to the strong polarity of the cyano group, there is a large passivation effect on the metal negative electrode leading to an increase in the interfacial resistance, thus requiring an effective negative interface protection.

4.3. Composite Polymer Electrolytes

Composite polymer electrolytes (CPEs) combining the soft mechanical properties of organic solid electrolytes and the high ionic conductivity of inorganic solid electrolytes offer a promising direction for highly stable Na-CO2 battery electrolytes [97,115,164]. PEs have a generally low ionic conductivity. Inorganic powders incorporated into solid polymer electrolytes to obtain composite solid electrolytes can effectively reduce the orderly arrangement of polymer chains and prevent polymer crystallization; and more ion transport channels are formed in the surface region of nanoparticles, which helps to improve ionic conductivity. In 1988, Skaarup et al. added fast ionic conductor particles Li3N as fillers to polyethylene oxide (PEO)-LiCF3SO3 matrix to obtain a composite polymer electrolyte, which not only improved its mechanical strength, but also had a higher ionic conductivity than the pure polymer electrolyte [165]. More interestingly, Wieczorek et al. found that the addition of nonionic fillers such as Al2O3 also improved the ionic conductivity of PEO-based polymer electrolytes, mainly due to the increase in the amorphous phase [166]. In addition, the addition of inorganic filler can improve the mechanical properties of electrolyte and maintain good flexibility, which improves the interfacial contact between electrolyte and electrode.
Inorganic fillers are mainly divided into two major categories: one is inert fillers, which do not have ion transport capability by themselves, such as Al2O3 [167], SiO2 and TiO2 [168], etc.; the other is active fillers, which have ion transport capability by themselves, such as Na2SiO3 [169], NASICON [170], β″-Al2O3 [171] and other inorganic solid electrolytes. Sun’s group prepared a high performance PVDF-HFP-Na3.2Zr1.9Mg0.1Si2PO12 organic-inorganic composite solid electrolyte [113]. The Na3.2Zr1.9Mg0.1Si2PO12 inorganic material with an ionic conductivity of 1.16 mS cm−1 at room temperature was obtained by replacing the Zr ion in Na3Zr2Si3PO12 with Mg2+, which was then combined with structurally stable PVDF-HFP with excellent mechanical properties as a solid electrolyte for Na-CO2 batteries. After 120 cycles at cut-off capacities of 200 mA g−1 and 500 mAh g−1, the intermediate gap voltage was below 2 V. The Coulomb efficiency, multiplicative performance and cycling performance of the batteries were significantly improved.
In addition to the above method of compounding inorganic particles directly with polymers, composite solid electrolytes can also be prepared by compounding functional polymers, functional inorganic particles, or polymer/inorganic particles together with a polymer matrix. For example, blending cellulose in sodium ionomer solid polymers [172] or glass fiber [173] can effectively enhance the mechanical properties of the electrolyte film; or adding functional inorganic particles to immobilize the anions on the polymer matrix chains and inhibit the formation of polarization centers, thus promoting the movement of cations [174].
Figure 9. (a) Schematic illustration of Li ion transport mechanism in PEO-based SPEs. Reproduced with permission from [154]. Copyright 2016, Royal Society of Chemistry. (b) Composition of the CPE. Inset: Transmission electron microscopy (TEM) image of fumed SiO2. (c) Ionic conductivity of CPE with various contents of SiO2. Reproduced with permission from [159]. Copyright 2017, the American Association for the Advancement of Science. (d) Charge/discharge curves tested at different current densities. (e) Cut-off capacity of NZM1SP-PVDF-HFP electrolytes at 200 mA g−1 for 500 mAh g−1. (f) Corresponding changes in intermediate charging and discharging voltages of cells with NZSP-PVDF-HFP and NZM1SP-PVDFHFP electrolytes. Reproduced with permission from [113]. Copyright 2022, John Wiley and Sons.
Figure 9. (a) Schematic illustration of Li ion transport mechanism in PEO-based SPEs. Reproduced with permission from [154]. Copyright 2016, Royal Society of Chemistry. (b) Composition of the CPE. Inset: Transmission electron microscopy (TEM) image of fumed SiO2. (c) Ionic conductivity of CPE with various contents of SiO2. Reproduced with permission from [159]. Copyright 2017, the American Association for the Advancement of Science. (d) Charge/discharge curves tested at different current densities. (e) Cut-off capacity of NZM1SP-PVDF-HFP electrolytes at 200 mA g−1 for 500 mAh g−1. (f) Corresponding changes in intermediate charging and discharging voltages of cells with NZSP-PVDF-HFP and NZM1SP-PVDFHFP electrolytes. Reproduced with permission from [113]. Copyright 2022, John Wiley and Sons.
Batteries 09 00036 g009

4.4. Interfaces of Solid Electrolytes

During the electrochemical cycling of a solid-state Na-CO2 batteries, the interface problems between the solid electrolyte and the electrode include two main ones: interfacial close contact and interfacial compatibility (no chemical reaction).
First, the close contact between the interfaces will facilitate the rapid transport of electrons or ions. In the conventional liquid electrolyte battery, the electrolyte has extremely high fluidity and good wettability to the electrode material, which ensures a low contact impedance inside the battery. In contrast, in solid-state batteries, the effective contact area is greatly reduced due to the point to point solid–solid contact between the electrolyte and the electrode, thus the interfacial impedance between the solid electrolyte and the electrode is very high (Figure 10a). Therefore, the current research work on solid-state batteries is mainly focused on improving the contact between solid electrolyte and electrode material and reducing the interfacial impedance. Sun et al. developed a monolithic symmetric cell for solid-state sodium batteries (SSSBs). A three-dimensional (3D) electronic and ionic conductive network is formed by integrating sodium anodes into NZSP-type monolithic structure. The interfacial resistance of the monolithic symmetric cell was significantly reduced, and exhibited a stable sodium plating/strip cycle with a low polarization of over 600 h [175].
Tong et al. artificially improved the interfacial stability and interfacial contact on the positive side of NASICON by obtaining a butanedionitrile-based plastic crystal electrolyte in situ on the cathode interface side (Figure 10b). This interfacial layer can be reversibly deformed with the change in particle volume, which reduces the interfacial charge transfer resistance and NASICON electrolyte is protected from successive interfacial side reactions, and thus the prepared solid-state Na-CO2 battery can stably cycle for more than 50 cycles at a limited capacity of 500 mAh g−1 [29]. In addition, solution casting method and in situ polymerization are also effective measures to enhance their interfacial contact.
In addition, chemical and electrochemical compatibility between different components will contribute to the stable operation of solid-state batteries (Figure 10c). Usually, the electrolyte and electrode will react at the interface, and dendrites may also form at the interface during the electroplating and stripping process, leading to short-circuiting of the battery [176,177]. Effective measures to solve the above problems include: (1) Optimizing the composition of the solid electrolyte or electrode, where the electrolyte reacts with the electrode to form an intermediate phase, but the interfacial products are stable and can reduce the interfacial impedance. For example, Tong et al. found that on the anode interface side, Na3Zr2Si2PO12 reacts with Na metal to form a thin passivation layer (indicated by NaO), which not only prevents further reaction between electrolyte and Na metal, but also improves the interfacial contact between electrolyte and anode. (2) Optimizing the composition of electrode, by reducing the local current density and homogenizing the electric field distribution, so as to guide the sodium ion fluxes and improve the problem of battery short circuit caused by interface dendrites. Based on this idea, Sun et al. modified the surface oxidation functional groups of three-dimensional carbon cloth to make the carbon cloth sodium-friendly, and then injected molten sodium to obtain a stable Na @ CC anode which inhibited dendrite growth. The composite anode greatly improves the reversibility and safety of the electrode [178]. (3) To introduce artificial interfacial layers to reduce the interfacial impedance caused by the chemical or electrochemical reaction of the intermediate phase (Figure 10d). Goodenough et al. [179] heated sodium metal and NASICON to 380 °C, chemically reacted to get stable interface layer and then reduced to room temperature, and the negative metal Na showed good wettability to NASICON, and the symmetric batteries of sodium achieved good cycling at 65 °C and at current densities of 0.15 mA cm−2 and 0.25 mA cm−2. Sun et al. added 1,1,1,3,3,3-hexafluoroisopropylmethyl ether (HFPM) to 0.1 M NaPF6 in 1,2-dimethoxyethane (DME) and fluoroethylene carbonate (FEC) electrolyte to form a new fluorine-containing organic layer, effectively stabilized the surface of Na anode, and showed excellent cycling performance [180].
Recently, a new approach can solve the above interfacial close contact and interfacial compatibility simultaneously. For example, Sui et al. and others constructed a thin self-reinforcing GPE in situ by polymerizing 1,3-dioxolane (DOL) in the nanofiber skeleton [181]. The framework is composed of polydopamine modified PVDF-HFP (PDA/PVDF-HFP) nanofiber membrane (Figure 10e). It has good affinity with PDOL and DOL, and can improve ionic conductivity. The DOL precursor solution can be firmly absorbed in the porous membrane, making the PVDF-HFP chain well swelled and forming good contact with the electrode interface, and be more stable to the negative metal Na [182]. After polymerization, there is no residual organic liquid GPE to ensure the full safety of the battery. Similarly, Sun et al. injected methylmethacrylate (MMA) into porous Na3Zr2Si2PO12-polymer vinylidene fluoride-hexafluoropropylene (NZSP-PVDF-HFP) composite membrane and obtained in situ polymerization polymethylmethacrylate (PMMA)-filled composite electrolyte membrane with excellent electrochemical performance [183].
Figure 10. (a) Schematic diagram of the “point-to-point” contact between the porous cathode and the NZSP. Reproduced with permission from [42]. Copyright 2021, Elsevier Ltd. (b) SN wetting on the porous cathode surface. Reproduced with permission from [29]. Copyright 2019, Royal Society of Chemistry. (c) Na dendrite growth process. (d) Contact model of ceramic solid electrolyte with metallic sodium in the process of sodium plating. Reproduced with permission from [184]. Copyright 2020, Elsevier Ltd. (e) Preparation of 3D PDOL@PDA/PVDF-HFP gel polymer electrolyte via in situ polymerization. Reproduced with permission from [182]. Copyright 2022, John Wiley and Sons.
Figure 10. (a) Schematic diagram of the “point-to-point” contact between the porous cathode and the NZSP. Reproduced with permission from [42]. Copyright 2021, Elsevier Ltd. (b) SN wetting on the porous cathode surface. Reproduced with permission from [29]. Copyright 2019, Royal Society of Chemistry. (c) Na dendrite growth process. (d) Contact model of ceramic solid electrolyte with metallic sodium in the process of sodium plating. Reproduced with permission from [184]. Copyright 2020, Elsevier Ltd. (e) Preparation of 3D PDOL@PDA/PVDF-HFP gel polymer electrolyte via in situ polymerization. Reproduced with permission from [182]. Copyright 2022, John Wiley and Sons.
Batteries 09 00036 g010

5. In Situ Characterization Techniques and Theoretical Calculations

The electrochemical performance of metal–air batteries is closely related to the composition and structure of each component and the process of battery fabrication. A detailed understanding of the mechanism of action of each component is required to assemble a battery with excellent performance [185]. The combination of theoretical calculations and characterization techniques can help to better understand the electrochemical reaction mechanisms and the constitutive relationships of batteries materials in the cell system, which in turn can guide the design of safer batteries with higher energy density and longer life [186,187].

5.1. In Situ Characterization Techniques

During charge and discharge, the transport of Na+ ions between the cathode and anode usually induces various changes in the electrode materials and the electrode/electrolyte interface, such as material volume expansion and contraction, structural phase changes, morphological evolution and surface reconstruction phenomena. With the help of in situ characterization technology, the structural development process of Na-CO2 batteries can be monitored in real time on multiple scales, providing a strong theoretical basis for studying the structural evolution of Na-CO2 batteries [188,189,190,191]. This section summarizes typical in situ characterization techniques, including in situ diffraction (XRD, NPD and PDF), in situ microscopy (SEM, TEM and AFM), and in situ spectroscopy (FTIR, Raman and XAS), and gives some representative application examples. The key importance of each technique is highlighted in the study of metal–air batteries [185,189,191,192,193].

5.1.1. In Situ Diffraction Technique

The diffraction technique is a common method for characterizing the composition of crystalline materials and phases. The basic idea is as follows: any phase has its own characteristic diffraction spectrum; the diffraction patterns of any two phases cannot be identical; the diffraction peaks of multi-phase samples are the simple superposition of the phases [194,195]. Therefore, the position and intensity of the peaks are closely related to the composition, arrangement and type of atoms in the crystal. Typical in situ diffraction techniques include in situ XRD, in situ NPD and in situ PDF.
XRD originates from the interference of X-rays scattering crystals to produce diffraction patterns. In situ XRD can be used to monitor information on phase changes, lattice parameter changes and intermediate phase formation of the batteries’ material during charge and discharge [194,196]. The development of in situ XRD technology offers the possibility to explore the reaction mechanism of metal–CO2 batteries [197,198,199]. The battery structure for in situ XRD is shown in Figure 11a [196]. Chen and colleagues used in situ XRD to observe the growth of Na2O, the discharge product of Na-O2 batteries in different electrolytes [198]. It proved that the electrolyte has an important influence on the reaction mechanism of Na-O2 batteries, i.e., NaO2 is the main discharge product in high and medium donor number electrolytes, but is rapidly converted to Na2O2 in low donor number electrolytes.
In addition to the conventional laboratory X-ray light source, the synchrotron X-ray light source can provide higher intensity and larger photon energy, which can collect higher signal-to-noise diffraction data in a short period of time and thus better study the lattice changes and structural evolution of battery materials during charging and discharging, reflecting more real-time change information [200].
Similar to in situ XRD techniques, in situ neutron diffraction allows for unique information for detecting the structural evolution of materials because neutrons are more sensitive to lighter elements (e.g., H [201,202], Li [97,203], O [204,205,206], Na [207], Al [208]) and can easily distinguish elements with similar atomic numbers (e.g., Fe and Mn) [209]. Specific applications include component determination, crystallinity, lattice constants, ion diffusion, expansion tensor, bulk modulus, phase transition and structure refinement and determination. However, the disadvantage of neutron diffraction is that it requires a strong neutron source, often a large sample size and a long data collection time [210]. Using NPD technology, Sun et al. found that there is an opportunity to create more sodium ion diffusion paths and improve the sodium ion diffusion capacity by applying perturbations, such as doping elements [113]. However, in situ XRD or NPD methods to study the long-range crystal structure of materials are unable to determine phase transitions in amorphous electrode materials or local structural changes during battery cycling.
Unlike the above diffraction techniques, the pair distribution function (PDF) can provide information about the local structure in the material, and can study both crystalline and amorphous materials [211]. The pair distribution function is also based on X-ray or neutron light source tests, but it provides information on the structure of the material at the atomic scale, especially on the atom–atom interactions in the structure. Fourier transform allows to obtain a real-space pair distribution function map, based on peak positions that can directly correspond to adjacent atomic spacings in real space; the peak intensity corresponds to the relative abundance of adjacent atomic spacings. For amorphous materials, local structural information, such as bond lengths and coordination numbers, can be extracted from the pair distribution familiarity map; for crystalline materials, the corresponding structural information can be fitted by taking the corresponding crystal structure model [212,213]. For crystalline materials, the corresponding crystal structure model can be adopted to fit the corresponding structural information. Unfortunately, there are few reports on the use of in situ PDF for characterization, and we believe that these studies will play a more important role in the design and development of solid state Na-CO2 batteries in the future.
In situ diffraction techniques are important to obtain information on the structural changes in battery materials. The application of in situ diffraction techniques such as in situ XRD, in situ NPD and in situ PDF in the study of metal–air batteries are emphasized. However, some microscopy and spectroscopy techniques are still needed to obtain more comprehensive information about the structure and composition of cell materials. For example, while synchrotron radiation-based X-ray imaging techniques can provide morphological and chemical information over a wide length scale from tens to hundreds of nanometers. Although synchrotron X-ray imaging techniques can provide morphological and chemical information on scales from tens of nanometers to a few millimeters, imaging at the atomic length scale (submicron) still requires transmission electron microscopy (TEM) techniques. In addition, spectroscopic techniques can also provide important complementary information for structural analysis of cellular materials.

5.1.2. In Situ Microscopy Techniques

In situ microscopy techniques, such as Scanning Electron Microscopy (SEM), Transmission Electron Microscope (TEM) and Atomic Force Microscope (AFM), allow visual observation of particle size and morphological changes on the material surface and is often used to observe the fine structure of the electrochemical reaction process in metal–air batteries and thus to determine the (catalytic) reaction mechanism. In situ microscopy combined with other spectroscopic techniques, such as energy dispersive spectroscopy (EDS) and electron energy loss spectroscopy (EELS), gives information on the chemical composition, electronic structure and chemical bonding of the sample. TEM-EDS is essential for understanding the chemical composition and microstructure of materials. Scanning transmission electron microscopy-electron energy loss spectroscopy (STEM-EELS) also allows for selective imaging of various parts of the electron energy loss spectrum to obtain information on the structure, interfacial features, diffusion pathways and electronic structure at the atomic scale in the sample.
SEM uses a focused electron beam to scan and image the surface of a specimen point by point, and obtains information about the surface morphology of the test specimen by accepting, amplifying and displaying imaging of these signals. Usually, SEM collects secondary electrons from the sample surface, and its lining degree reflects the surface morphology and roughness of the sample. The spatial resolution of SEM can reach 10 nm, and the actual resolution is limited by the conductivity of the sample and the environment of the electron microscope cavity, with some time-resolved capability for in situ characterization of the battery. Neelam et al. studied the degradation mechanism of two sulfide-based solid electrolytes, b-Li3PS4 and Li6PS5Cl, during the operation of batteries by in situ SEM [214]. As shown in Figure 11b, compared with Li6PS5Cl, b-Li3PS4 shows huge plating, faster dendrite formation, cracks and ultimately cell failure.
TEM has extremely high spatial resolution, can realize the observation of diffraction patterns in tiny regions (several nanometers), and is suitable for the study of crystal structure of microcrystals, surfaces and thin films. Because electrons are easily scattered or absorbed by objects with low penetration, TEM must prepare ultra-thin samples, usually in thickness of 50–100 nm; on the other hand, the intense irradiation of the electron beam tends to damage the sample and bring artifacts. The in situ TEM technique has been applied to the study of the reaction mechanism and catalyst of metal–CO2 batteries [44,215,216,217]. Zhu et al. directly studied the morphological changes in the cathode surface of single Pt atom-loaded nitrogen-doped carbon nanotubes (Pt@NCNT) during the charging and discharging of Na-CO2 batteries using in situ ambient transmission electron microscopy (TEM) under CO2 atmosphere (Figure 11c) [44]. Han et al. investigated the mechanism of the reversible redox reaction in Na-O2 batteries with CuS cathodes using in situ TEM [215,216]. The results show that Na2O2 is the main final ORR product, which uniformly covers the whole linear cathode (Figure 11d). In addition, Huang et al. studied the reaction process of Na-O2/CO2 battery in situ using spherical aberration-corrected environmental transmission electron microscopy, and characterized the structure and composition of the discharging and partially charging products by using annular dark field images (ADF) and EELS [187]. Figure 12a,b showed the ADF diagrams characterizing the cathode surface: the discharge and charge products had similar spherical profiles in the low-loss and core-loss spectra. As shown in Figure 12c,d, EELS showed that: only Na and Na2O2 are present in the discharge and charge spheres, and the shell layer shows the formulation of Na2CO3.
Figure 11. (a) Schematic diagram of a typical battery device for in-situ XRD testing;. Reproduced with permission from [196]. Copyright 2019, John Wiley and Sons. (b) Plating, cracking and dendrite formation in b-Li3PS4 and Li6PS5C. Reproduced with permission from [214]. Copyright 2021, Cambridge University Press. (c) Morphology evolution of individual platinum-doped NCNTs (Pt@NCNTs) during electrochemical discharge and charge of Na-CO2 nanobatteries. Reproduced with permission from [44]. Copyright 2020, Elsevier Ltd. (d) In situ TEM micrographs of the OER process in Na-O2 batteries. Reproduced with permission from [215]. Copyright 2020, American Chemical Society.
Figure 11. (a) Schematic diagram of a typical battery device for in-situ XRD testing;. Reproduced with permission from [196]. Copyright 2019, John Wiley and Sons. (b) Plating, cracking and dendrite formation in b-Li3PS4 and Li6PS5C. Reproduced with permission from [214]. Copyright 2021, Cambridge University Press. (c) Morphology evolution of individual platinum-doped NCNTs (Pt@NCNTs) during electrochemical discharge and charge of Na-CO2 nanobatteries. Reproduced with permission from [44]. Copyright 2020, Elsevier Ltd. (d) In situ TEM micrographs of the OER process in Na-O2 batteries. Reproduced with permission from [215]. Copyright 2020, American Chemical Society.
Batteries 09 00036 g011
Compared with SEM and TEM, AFM can more easily simulate cell environment conditioning, and be used for in situ operational characterization to monitor the microscopic morphology of the electrode surface in real time, provide physicochemical information of the electrode surface at the nanoscale, and provide an experimental basis for the optimization of electrode materials and electrolyte modification. In addition, unlike scanning electron microscopy which can only provide two-dimensional images, AFM can provide true three-dimensional images. At the same time, scanning electron microscopy needs to operate under high vacuum conditions, while AFM can work under atmospheric pressure or even in liquid environments. Cort’es et al. used improved AFM without putting AFM in glove box, visualized the formation and decomposition process on thin Li2O2 carbon cathode as shown in Figure 12e. This equipment provides a technical support for future studies of new cathode materials [218].
Figure 12. Annular dark field images (ADF) and EELS characterization of the structure and composition of the discharge and partial charging products: (a,b) ADF images of the time-dependent structural evolution of the air cathode of the Na-O2/CO2 batteries. During the discharge reaction, two balls emerged at 561 s at the CNT-Na substrate-O2/CO2 triple point, which then grew under a constant −3 V bias, showing the core-shell structure of the ball. (c,d) Low-loss and core-loss EELS spectra. Reproduced with permission from [187]. The low-loss and core-loss spectra of both the discharged (red profile) and charged (blue profile) balls show similar profile: in the Low-Loss region, there are three multiple plasma peaks at 5.7, 11.5 and 17.3 eV respectively, and another three plasma peaks at 22.9, 31.1 and 37.6 eV, and combined with side absorption peaks of O-K and Na-K in the Core-Loss region, indicating the presence of only Na and Na2O2 exist in the discharge sphere and charge sphere. The Core-Loss of the shell layer shows the presence of Na, C and O, indicating the formation of Na2CO3.Copyright 2020, American Chemical Society. (e) Schematic diagram of the flow electrochemical atomic force microscope (FE-AFM). Reproduced with permission from [218]. Copyright 2021, Elsevier Ltd. (f,g) In situ SERS characterization using LiCF3SO3-TEGDME with and without Ru catalyst electrodes during discharge and recharge in CO2. Reproduced with permission from [27]. Copyright 2017, Royal Society of Chemistry.
Figure 12. Annular dark field images (ADF) and EELS characterization of the structure and composition of the discharge and partial charging products: (a,b) ADF images of the time-dependent structural evolution of the air cathode of the Na-O2/CO2 batteries. During the discharge reaction, two balls emerged at 561 s at the CNT-Na substrate-O2/CO2 triple point, which then grew under a constant −3 V bias, showing the core-shell structure of the ball. (c,d) Low-loss and core-loss EELS spectra. Reproduced with permission from [187]. The low-loss and core-loss spectra of both the discharged (red profile) and charged (blue profile) balls show similar profile: in the Low-Loss region, there are three multiple plasma peaks at 5.7, 11.5 and 17.3 eV respectively, and another three plasma peaks at 22.9, 31.1 and 37.6 eV, and combined with side absorption peaks of O-K and Na-K in the Core-Loss region, indicating the presence of only Na and Na2O2 exist in the discharge sphere and charge sphere. The Core-Loss of the shell layer shows the presence of Na, C and O, indicating the formation of Na2CO3.Copyright 2020, American Chemical Society. (e) Schematic diagram of the flow electrochemical atomic force microscope (FE-AFM). Reproduced with permission from [218]. Copyright 2021, Elsevier Ltd. (f,g) In situ SERS characterization using LiCF3SO3-TEGDME with and without Ru catalyst electrodes during discharge and recharge in CO2. Reproduced with permission from [27]. Copyright 2017, Royal Society of Chemistry.
Batteries 09 00036 g012

5.1.3. In Situ Spectroscopy Technique

Spectroscopic techniques such as XPS, FTIR, Raman and XAS are sensitive to changes in the local chemical environment, have elemental resolution and are suitable for compositional analysis of crystalline and amorphous phases.
X-ray photoelectron spectroscopy (XPS) is based on the photoelectric effect, using the interaction of X-rays with the surface of the sample to generate photoelectrons, and using an energy analyzer to analyze the kinetic energy of the photoelectrons, the X-ray photoelectron spectrum is thus obtained. Based on the measured kinetic energy of photoelectrons, it is possible to determine which elements are present on the surface and to know the content of an element on the surface based on the intensity of photoelectrons with a certain energy, with an error of about 20%. It is possible to determine both the relative concentration of elements and the relative concentration of different oxidation states of the same element. XPS not only determines the composition of all elements on the surface except H and He, but also gives information on the chemical state of each element, with high energy resolution and a certain spatial resolution (currently on the micron scale) and temporal resolution (on the minute scale). XPS also gives information on the surface, tiny areas and depth distribution information. In situ atmospheric pressure XPS (APXPS) is an effective tool to study the chemical state changes on the surface of solid electrolytes and catalysts. Wang et al. used synchrotron in situ APXPS to study the redox reaction of CO2 on the surface of porous carbon electrode in ionic liquid electrolytes [219]. The results showed that the reduction in pure CO2 at porous carbon electrodes has no electrochemical activity at room temperature. In contrast, when O2 (CO2:O2 = 2:1) is added to the ionic liquid electrolyte of Li-CO2 batteries, CO2 is reduced to low-valent amorphous carbon and Li2O2/Li2O. Compared with the reaction of Li2CO3, the charging reaction of amorphous carbon, Li2O2 and Li2O is faster. The main reason why the ultimately discharge product is not Li2CO3 is that the strong solvation between ionic liquid electrolyte and Li+ stabilizes the intermediate anion of metastable CRR reaction.
In situ Fourier infrared reflectance absorption spectroscopy (FTIR) is particularly advantageous in identifying different functional groups and components during electrochemical reactions, and thus in situ FTIR techniques can be used to analyze reaction intermediates and determine reaction mechanisms. Although the application of in situ FTIR in metal–air batteries has not been widely reported, it can be expected to play an important role in future research in this field.
In situ Raman spectroscopy is based on the detection of laser-induced vibrational, rotational and other low-frequency leap patterns in electrode materials, and is a fingerprint spectrum of the structure of matter [192]. Raman spectroscopy is suitable for vibration mode measurements of symmetric vibrations, non-polar groups and homoatomic bonds, such as S=S, S-S, N=N, C=C and O=O. In addition to identifying substance types and chemical components, Raman spectroscopy mainly measures molecular vibration frequencies to quantitatively understand intermolecular forces and intra-molecular forces, and to infer information such as symmetry, geometry and arrangement of atoms in molecules. However, conventional Raman spectroscopy signals are relatively weak, and surface-enhanced Raman spectroscopy (SERS) is able to detect lower concentrations of species in the surface layers of bulk materials.
Zhou et al. used in situ SERS to study the catalytic activity of Ru in promoting CO2 reduction to Li2CO3 and C on ruthenium-free sputtered gold electrodes and gold-ruthenium electrodes [27]. By observing figures f and g, it can be seen that the peak Raman intensity corresponding to Li2CO3 at 1080 cm−1 decreases during the charging process and completely disappears at the end of the charging process. Comparing Figure 12f,g, the difference is mainly the Raman peak corresponding to the G-band of carbon at 1580 cm−1. As shown in Figure 12f, the peak value of the carbon G band corresponding to 1580 cm−1 in the gold electrode hardly changes during the whole charging process. In the Au-Ru electrode (Figure 12g), the peak intensity of the G-band in carbon (1580 cm−1) decreases during the charging process and disappears completely at the end of the charging process, which is similar to the Raman peak of Li2CO3. This result shows that the ruthenium catalyst can significantly promote the reaction between Li2CO3 and carbon during charging (charging reaction), while the absence of ruthenium leads to the self-decomposition of Li2CO3 and leads to an irreversible discharge–recharging cycle of the battery.
However, SERS can only characterize information from a few atomic layers on the surface. In contrast, X-ray absorption spectroscopy (XAS) can effectively capture deeper signal variations in bulk materials. By analyzing and fitting the data to the X-ray absorption spectra, the most accurate characterization of the sample as a whole can be obtained, including interatomic distances, number and type of neighboring atoms and information on the average oxidation state of the elements in the coordination environment [189,220]. In addition, unlike in situ diffraction techniques, in situ XAS can provide real-time element-specific information on crystalline and amorphous evolution and phase transitions. XAS can be divided into X-ray absorption near edge structure (XANES) spectroscopy and extended X-ray absorption fine structure (EXAFS) [221]. EXAFS is commonly used in the range of 150–2000 eV to obtain quantitative local structural information such as bond length, coordination number and disorder of the central and coordination atoms. On the other hand, XANES can quickly identify the chemical state of elements in the low kinetic energy range of 5–150 eV. In addition, XANES can be used for time-resolved experiments and high-temperature in situ chemistry experiments because of its violent vibrations, short spectrum acquisition time and temperature dependence. The self-reinforced catalysis of LiCoO2 as an electrocatalyst for Li-O2 batteries was reported by Zhou et al. [222]. The evolution of Co electrons and local structure was investigated using XAFS, illustrating that the intercalation/extraction of Li+ in LiCoO2 not only induces changes in Co valence and regulates the electron/crystal structure, but also the surface disorder, lattice strain and local symmetry, thus promoting bidirectional catalysis. The in situ technique XAFS has not been reported.
Photoelectronic information (XPS), which is sensitive to the material surface, can detect the changes in the element composition and chemical state information on the battery surface in real time. Raman spectroscopy is derived from the inelastic scattering of incident laser by the molecules on the surface of the object. In situ Raman spectroscopy can detect the changes in material composition and structure on the electrode or solid electrolyte surface of the all-solid battery. In situ XAS plays an irreplaceable role in the rapid and high precision analysis of the elements and their valence states and their respective distributions in solid state batteries. With the construction of more and more advanced synchronous light sources, in situ spectroscopy will play an increasingly important role in the research of solid state metal–air batteries and other energy materials.
Since each characterization method has its own advantages and disadvantages, combining the advantages of different characterization techniques to study the physicochemical changes and failure mechanisms of solid Na-CO2 electrode materials and interfaces during operation can provide an important information for further understanding and optimization of material performance, and provide strong support for subsequent improvement of battery performance including energy density, cycle life and safety. With the development of modern technology, these in situ characterization techniques and data acquisition and analysis systems will be further improved and intelligent, and more in situ characterization techniques will be available for more systematic real-time inspection of all-solid-state metal–air batteries to guide the design of solid-state metal–air batteries.

5.2. Theoretical Calculations and Simulations

Theoretical calculations and simulations also play important roles in the development of battery technology. With the rapid development and wide application of computer simulation techniques in quantum and atomic scale-based materials science research, it provides the possibility to understand the reaction mechanism of microscopic metal–air batteries and to develop and design new battery materials efficiently. For catalytic materials, theoretical calculations and simulations can calculate adsorption energy, migration energy, charge distribution, electronic structure and even defect states to understand the reaction mechanism; for electrode materials, theoretical calculations and simulations can calculate the energy band gap, sodium storage voltage, bulk deformation and charge compensation mechanism of positive and negative electrodes and electrolyte materials. In addition, theoretical calculations and simulations can study the electrochemical window, ionic conductivity, ion mobility and ion diffusion barrier and transport mechanism. With the dramatic increase in computer computing power, methods such as high-throughput computing and machine learning have also started to develop rapidly, making the research and development of battery materials tend to be more efficient and intelligent. The combination of theoretical calculations and experiments helps to use the deep understanding of electrochemical theory to design and control electrochemical reactions in turn, thus advancing the battery research. This section outlines some common computational and simulation methods based on quantum mechanical theory in battery research, and briefly introduces the application of materials genetic engineering and machine learning in battery materials development.

5.2.1. Theoretical Calculation and Simulation Methods Based on Quantum Mechanics

Density Function Theory (DFT) was developed by Kohn, the 1998 Nobel Prize Chemistry awardee, which enables efficient solution of the many-body Schrödinger equation. In recent years, theoretical calculations based on density DFT and molecular dynamics (MD) simulations have greatly contributed to the understanding of the reaction mechanism of metal–air batteries at the atomic scale and to the efficient design of battery materials. The calculation of various properties of materials by inputting structural information of crystals has provided a theoretical basis for understanding the reaction mechanism, catalytic reaction kinetics and ionic transport in electrolytes in metal–air batteries.
Due to the multi-electron composite process of oxygen reduction reaction (ORR) and oxygen evolution reaction (OER), an in-depth understanding of the catalytic reaction process of cathode catalysts is essential to improve the reaction kinetics by changing the growth path and morphological evolution of the cathode surface [223]. In recent years, first-principles calculations have become a routine approach to understand complex catalytic phenomena and experiments at the electronic level, and have been further extended to find and design novel catalysts [25,61,224]. Jiang et al. used DFT calculations to investigate the contribution of point defects on carbon surfaces to the reaction of non-aqueous Li-O2 batteries [225]. In this work, five representative defect structures were considered, including SV (point defects), DV5-8-5 (two pentagons and one octagon), DV555-777 (three pentagons and three heptagons), DV5555-6-7777 (four pentagons, one hexagon and four heptagons) and SW (Stone–Wales) defects. The calculated free energy results (Figure 13a–f) showed that the DV555-777, DV5555-67777 and SW type defects can increase the discharge voltage and decrease the charge voltage, while promoting the ORR and OER processes. In addition, the discharge voltage of DV555-777 is the highest, and the charging voltage of DV5555-6-7777 and SW is the lowest. Therefore, non-aqueous Li-O2 batteries require carbon materials containing DV5555-6-7777 and SW type defects. Deng et al. obtained porous self-supporting cathodes for Li-CO2 batteries by directly anchoring two-dimensional (2D) cobalt-doped CeO2 nanosheets on graphene aerogels. Combined with the experimental results and DFT calculations, it was proved that co-doped CeO2 nanosheets could effectively improve the conductivity and adsorption capacity of CO2. In addition, they obtained the possible reaction path of non-aqueous Li-O2 batteries from the perspective of thermodynamics, and the analysis results are shown in Figure 13g [224]. Recently, Sun et al. constructed a new autocatalytic system for lithium–air batteries using an in situ electrochemical doping strategy combined with theoretical computational simulations and systematically investigated its reaction mechanism and battery performance. The work was based on density function theory (DFT) for the theoretical study, and it was found that the batteries’ performance could be significantly enhanced by doping the discharge product Li2O2 with metal ions, and then the theoretical calculations were used to guide the experimental screening of the most effective doping structure [226].
For solid electrolytes, polymeric solid electrolytes are difficult to establish standard models for calculations due to their complex structures. In contrast, inorganic solid electrolytes, such as Na3Zr2Si2PO12, are easier to model due to their simple structure, and their corresponding theoretical simulations and calculations are more studied [107,227]. Park et al. investigated the effect of adding excess Na to standard NASICON and the mechanism of Na+ ion transport [107]. Both experimental and theoretical calculations demonstrated that the main mechanisms of Na+ ion transport in the NASICON structure are, grain boundary diffusion at low temperatures and grain diffusion at high temperatures, respectively (Figure 14a). In addition, it was found that after adding 10% excess Na, the expansion of the bottleneck of polycrystalline particle sodium diffusion channel was conducive to improving the bulk conductivity of NASICON. DFT calculation results show that by adding 10% excess Na, the minimum bottleneck area was expanded from 4.9922 Ǻ to 5.4086 Ǻ; the activation energy was significantly reduced. The minimum bottleneck area is the key factor to determine the conductivity of Na ion in NASICON electrolytes, which is basically consistent with the experimental results. Figure 14b shows that the ionic conductivity of NAICON has been significantly improved after adding excessive sodium, resulting from the enlargement of the bottleneck areas in the Na diffusion channels of polycrystalline grains.
The combination of DFT calculations and MD simulations has a strong potential in facilitating the development of metal–air batteries: the mass transfer dynamics of gas molecules and metal ions throughout battery system can be studied at the atomic scale. Notably, most of the current theoretical calculations are based on predictions under 0 K and vacuum conditions. However, powerful predictive tools should model the actual battery environment and conduct extensive experiments to verify theoretical simulations and improve accuracy.

5.2.2. Introduction to Material Genetic Engineering and Machine Learning

Materials genetic engineering is a new concept and method for materials research that has emerged in recent years and is at the forefront of materials science and engineering in the world today. Materials genetic engineering is a technological integration and synergy of high-throughput materials computing, high-throughput materials synthesis and characterization experiments and databases, which can at least double the speed of materials discovery, manufacturing and application, and reduce the cost by at least half. Materials genetic engineering is a leap forward in materials science and technology, replacing the traditional trial-and-error method of multiple sequential iterations with high-throughput parallel iterations, gradually changing from “experience-guided experiments” to a new model of “theoretical prediction and experimental verification” in materials research, and finally realizing materials design on demand.
There are some important steps in the process of high-throughput computational screening of materials [228]. First, a database of experimental crystal structures containing a large number of materials is needed as the seven points for high-throughput screening; second, structural information is input in bulk and data are screened; then the screened crystal structure information is generated into a data format that can be calculated using density functional theory and put into the server for calculation; then the calculation files are output and kept together to form a database after the calculation is completed; finally, the data from the database are used to analyze the various properties of the material, which can be used to guide the material design or to filter the data for the next cycle.
In recent years, the combination of machine learning and materials genetic engineering has led to the advancement of materials informatics and promoted the development of materials science. Currently, the use of data-driven machine learning algorithms to build material performance prediction models and then apply them to material screening and new material development has attracted much attention from research community. Machine learning can be seen as an umbrella term for a class of algorithms that can mine potential laws from large amounts of experimental data, acquire new knowledge or skills, build corresponding data analysis models and allow machines to repeatedly analyze the corresponding content and reorganize existing knowledge structures through the input of new data to continuously improve their performance. In the field of materials development, machine learning can show great potential and advantages, especially in the design of new catalysts, organic and inorganic battery materials, etc. [229]. A machine learning toolkit for genetic engineering improves the performance of solid-state Na-CO2 batteries by searching for high-performance solid electrolytes.
Materials genetic engineering and machine learning aim to establish a new model of data-driven, computational simulation and theoretical prediction first, experimental validation second, thus replacing the existing empirical and experimental-based materials R&D paradigm, which is a powerful tool for the future development of metal–air batteries.

6. Outlook

Metal–CO2 batteries are considered as a promising clean technology because of their potential for high energy density energy storage and power supply, as well as their ability to convert and immobilize CO2 as well as mitigate global warming trends. After nearly a decade of development, although the research of Na-CO2 battery has made great progress, compared with other energy storage systems, the research of Na-CO2 battery is still in the infancy stage, facing the challenges of unclear electrochemical reaction mechanism, slow catalytic reaction of CO2 gas at the cathode and the discharge product carbonate has difficulties in reversible transformation due to problems such as good thermodynamic stability, limited research on electrolyte materials, lack of suitable electrolytes and interface incompatibility. To solve these problems and realize the practical application of solid-state Na-CO2 batteries operating at room temperature, we propose the following aspects for further research. First, the analysis of the electrochemical mechanism of Na-CO2 batteries, especially the formation and decomposition process of the discharge products in the rate-determined step, is of great significance for the study of Na-CO2 batteries, and should be considered in conjunction with in situ experimental studies, such as the electrochemical scanning microscope system combined with Fourier transform infrared spectroscopy, to further rationalize the design of the Li/Na-CO2 cell system by analyzing the carbonate formation and decomposition pathways. Secondly, the sluggish reaction kinetics is a non-negligible challenge for Na-CO2 batteries, which will lead to large overpotential and poor reversibility. Therefore, efficient catalysts with low cost, reasonable structure and high catalytic ability need to be developed to fill this gap. Considering the cost and performance, heteroatom doping, noble metal single-atom catalysts and transition metal complex catalysts are very promising.
Third, the applications of solid electrolytes in Na-CO2 batteries are rather limited, and there is an urgent need for researchers to design and develop more solid electrolyte materials with high performance (e.g., high ionic conductivity and good chemical/electrochemical stability) based on fundamental principles and theoretical calculations, and to create stable interfaces with good contacts. In fact, ion transport in electrolytes and at interfaces is a multi-scale process, including atomic scale, micro and mesoscopic scale and device scale [102]. Monitoring the ion conduction and interfacial reactions at different scales, together with atomic-scale characterization techniques and theoretical modeling, are important for the design and fabrication of high-performance batteries. To improve the interfacial contact and form a three-dimensional (3D) electronic and ionic conducting network monolithic cell architecture may be a good choice. Finally, how to scale up the production of high-performance solid electrolytes by a simple, scalable and cost-effective method is also an important consideration for practical applications.
Finally, with the rapid development of in situ characterization technology and high performance computing, more comprehensive real-time inspection, data acquisition and analysis systems will be available and with the powerful predictive capabilities of computer simulation technology (especially theoretical calculations and simulations, genetic engineering of materials and machine learning), the development of battery materials will tend to be more efficient and intelligent, which is a powerful tool for the future development of metal–air batteries.
In summary, Na-CO2 batteries demonstrate great application potential in CO2 fixation as well as energy storage. They involve a cross multidiscipline. We believe that the development of practical metal–CO2 batteries will bring new hope for achieving the strategic goal of carbon neutralization and carbon peaking.

Author Contributions

Proposing and supervising this project, C.S.; Writing—review and editing, Z.W., L.L., L.J. and C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Foundation of Key Laboratory of Advanced Energy Materials Chemistry (Ministry of Education), Nankai University, 21C Innovation Laboratory, Contemporary Amperex Technology Ltd. by project No. 21C-OP-202212, the Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (Grant No. 2022-K15), China University of Mining & Technology (Beijing), Beijing National Laboratory for Condensed Matter Physics, and the National Natural Science Foundation of China (No. 51672029 and 51372271).

Institutional Review Board Statement

Written informed consent has been obtained from the patient(s) to publish this paper.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xie, J.F.; Zhou, Z.; Wang, Y.B. Metal-CO2 Batteries at the Crossroad to Practical Energy Storage and CO2 Recycle. Adv. Funct. Mater. 2020, 30, 1908285. [Google Scholar] [CrossRef]
  2. Wang, F.; Li, Y.; Xia, X.H.; Cai, W.; Chen, Q.G.; Chen, M.H. Metal-CO2 Electrochemistry: From CO2 Recycling to Energy Storage. Adv. Energy Mater. 2021, 11, 2100667. [Google Scholar] [CrossRef]
  3. Xia, C.; Black, R.; Fernandes, R.; Adams, B.; Nazar, L.F. The critical role of phase-transfer catalysis in aprotic sodium oxygen batteries. Nat. Chem. 2015, 7, 496–501. [Google Scholar] [CrossRef] [PubMed]
  4. Palomares, V.; Serras, P.; Villaluenga, I.; Hueso, K.B.; Carretero-González, J.; Rojo, T. Na-ion batteries, recent advances and present challenges to become low cost energy storage systems. Energy Environ. Sci. 2012, 5, 5884–5901. [Google Scholar] [CrossRef]
  5. Li, W.J.; Han, C.; Wang, W.L.; Gebert, F.; Chou, S.L.; Liu, H.K.; Zhang, X.H.; Dou, S.X. Commercial Prospects of Existing Cathode Materials for Sodium Ion Storage. Adv. Energy Mater. 2017, 7, 1700274. [Google Scholar] [CrossRef]
  6. Hu, X.F.; Sun, J.C.; Li, Z.F.; Zhao, Q.; Chen, C.C.; Chen, J. Rechargeable Room-Temperature Na-CO2 Batteries. Angew. Chem. Int. Ed. 2016, 55, 6482–6486. [Google Scholar] [CrossRef]
  7. Jena, A.; Tong, Z.Z.; Chang, H.; Hu, S.F.; Liu, R.S. Capturing carbon dioxide in Na-CO2 batteries: A route for green energy. J. Chin. Chem. Soc. 2020, 68, 421–428. [Google Scholar] [CrossRef]
  8. Pham, T.A.; Kweon, K.E.; Samanta, A.; Lordi, V.; Pask, J.E. Solvation and Dynamics of Sodium and Potassium in Ethylene Carbonate from ab Initio Molecular Dynamics Simulations. J. Phys. Chem. C 2017, 121, 21913–21920. [Google Scholar] [CrossRef]
  9. Mu, X.W.; Pan, H.; He, P.; Zhou, H.S. Li-CO2 and Na-CO2 Batteries: Toward Greener and Sustainable Electrical Energy Storage. Adv. Mater. 2020, 37, e1903790. [Google Scholar]
  10. Xie, Z.J.; Zhang, X.; Zhang, Z.; Zhou, Z. Metal-CO2 Batteries on the Road: CO2 from Contamination Gas to Energy Source. Adv. Mater. 2017, 29, 1605891. [Google Scholar] [CrossRef]
  11. Zheng, Z.; Wu, C.; Gu, Q.; Konstantinov, K.; Wang, J. Research Progress and Future Perspectives on Rechargeable Na-O2 and Na-CO2 Batteries. Energy Environ. Mater. 2021, 4, 158–177. [Google Scholar] [CrossRef]
  12. Takechi, K.; Shiga, T.; Asaoka, T. A Li-O2/CO2 battery. Chem. Commun. 2011, 47, 3463–3465. [Google Scholar] [CrossRef]
  13. Das, S.K.; Xu, S.M.; Archer, L.A. Carbon dioxide assist for non-aqueous sodium-oxygen batteries. Electrochem. Commun. 2013, 27, 59–62. [Google Scholar] [CrossRef]
  14. Liu, B.; Sun, Y.L.; Liu, L.Y.; Chen, J.T.; Yang, B.J.; Xu, S.; Yan, X.B. Recent advances in understanding Li-CO2 electrochemistry. Energy Environ. Sci. 2019, 12, 887–922. [Google Scholar] [CrossRef]
  15. Xu, S.M.; Lu, Y.Y.; Wang, H.S.; Abruña, H.D.; Archer, L.A. A rechargeable Na-CO2/O2 battery enabled by stable nanoparticle hybrid electrolytes. J. Mater. Chem. A 2014, 2, 17723–17729. [Google Scholar] [CrossRef]
  16. Guo, L.N.; Li, B.; Thirumal, V.; Song, J.X. Advanced rechargeable Na-CO2 batteries enabled by a ruthenium@porous carbon composite cathode with enhanced Na2CO3 reversibility. Chem. Commun. 2019, 55, 7946–7949. [Google Scholar] [CrossRef]
  17. Sui, D.; Chang, M.J.; Wang, H.Y.; Qian, H.; Yang, Y.L.; Li, S.; Zhang, Y.S.; Song, Y.Z. A Brief Review of Catalytic Cathode Materials for Na-CO2 Batteries. Catalysts 2021, 11, 603. [Google Scholar] [CrossRef]
  18. Zhou, J.W.; Li, X.L.; Yang, C.; Li, Y.C.; Guo, K.K.; Cheng, J.L.; Yuan, D.W.; Song, C.H.; Lu, J.; Wang, B. A Quasi-Solid-State Flexible Fiber-Shaped Li-CO2 Battery with Low Overpotential and High Energy Efficiency. Adv. Mater. 2019, 31, e1804439. [Google Scholar] [CrossRef]
  19. Yang, C.; Guo, K.K.; Yuan, D.W.; Cheng, J.L.; Wang, B. Unraveling Reaction Mechanisms of Mo2C as Cathode Catalyst in a Li-CO2 Battery. J. Am. Chem. Soc. 2020, 142, 6983–6990. [Google Scholar] [CrossRef]
  20. Zhang, Z.; Zhang, Q.; Chen, Y.N.; Bao, J.; Zhou, X.L.; Xie, Z.J.; Wei, J.P.; Zhou, Z. The First Introduction of Graphene to Rechargeable Li-CO2 Batteries. Angew. Chem. 2015, 127, 6550–6653. [Google Scholar] [CrossRef]
  21. Zhang, Z.; Wang, X.G.; Zhang, X.; Xie, Z.J.; Chen, Y.N.; Ma, L.P.; Peng, Z.Q.; Zhou, Z. Verifying the Rechargeability of Li-CO2 Batteries on Working Cathodes of Ni Nanoparticles Highly Dispersed on N-Doped Graphene. Adv. Sci. 2017, 5, 1700567. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, X.; Zhang, Q.; Zhang, Z.; Chen, Y.N.; Xie, Z.J.; Wei, J.P.; Zhou, Z. Rechargeable Li-CO2 batteries with carbon nanotubes as air cathodes. Chem. Commun. 2015, 51, 14636–14639. [Google Scholar] [CrossRef] [PubMed]
  23. Sun, J.Y.; Zhao, N.; Li, Y.Q.; Guo, X.X.; Feng, X.F.; Liu, X.S.; Liu, Z.; Cui, G.L.; Zheng, H.; Gu, L.; et al. A Rechargeable Li-Air Fuel Cell Battery Based on Garnet Solid Electrolytes. Sci. Rep. 2017, 7, 41217. [Google Scholar] [CrossRef] [Green Version]
  24. Li, S.W.; Dong, Y.; Zhou, J.W.; Liu, Y.; Wang, J.M.; Gao, X.; Han, Y.Z.; Qi, P.F.; Wang, B. Carbon dioxide in the cage: Manganese metal–organic frameworks for high performance CO2 electrodes in Li-CO2 batteries. Energy Environ. Sci. 2018, 11, 1318–1325. [Google Scholar] [CrossRef]
  25. Hu, X.F.; Joo, P.H.; Matios, E.; Wang, C.L.; Luo, J.M.; Yang, K.S.; Li, W.Y. Designing an All-Solid-State Sodium-Carbon Dioxide Battery Enabled by Nitrogen-Doped Nanocarbon. Nano Lett. 2020, 20, 3620–3626. [Google Scholar] [CrossRef] [PubMed]
  26. Xu, S.M.; Das, S.K.; Archer, L.A. The Li-CO2 battery: A novel method for CO2 capture and utilization. RSC Adv. 2013, 3, 6656–6660. [Google Scholar] [CrossRef]
  27. Yang, S.X.; Qiao, Y.; He, P.; Liu, Y.J.; Cheng, Z.; Zhu, J.J.; Zhou, H.S. A reversible lithium-CO2 battery with Ru nanoparticles as a cathode catalyst. Energy Environ. Sci. 2017, 10, 972–978. [Google Scholar] [CrossRef]
  28. Ha, T.A.; Pozo-Gonzalo, C.; Nairn, K.; MacFarlane, D.R.; Forsyth, M.; Howlett, P.C. An investigation of commercial carbon air cathode structure in ionic liquid based sodium oxygen batteries. Sci. Rep. 2020, 10, 7123. [Google Scholar] [CrossRef]
  29. Lu, Y.; Cai, Y.C.; Zhang, Q.; Liu, L.J.; Niu, Z.Q.; Chen, J. A compatible anode/succinonitrile-based electrolyte interface in all-solid-state Na-CO2 batteries. Chem. Sci. 2019, 10, 4306–4312. [Google Scholar] [CrossRef] [Green Version]
  30. Wang, X.C.; Zhang, X.J.; Lu, Y.; Yan, Z.H.; Tao, Z.L.; Jia, D.Z.; Chen, J. Flexible and Tailorable Na-CO2 Batteries Based on an All-Solid-State Polymer Electrolyte. ChemElectroChem 2018, 5, 3628–3632. [Google Scholar] [CrossRef]
  31. Li, Y.L.; Wang, J.J.; Li, X.F.; Geng, D.S.; Banis, M.N.; Li, R.Y.; Sun, X.L. Nitrogen-doped graphene nanosheets as cathode materials with excellent electrocatalytic activity for high capacity lithium-oxygen batteries. Electrochem. Commun. 2012, 18, 12–15. [Google Scholar] [CrossRef]
  32. Ci, L.J.; Song, L.; Jin, C.H.; Jariwala, D.; Wu, D.X.; Li, Y.J.; Srivastava, A.; Wang, Z.F.; Storr, K.; Balicas, L.; et al. Atomic layers of hybridized boron nitride and graphene domains. Nat. Mater. 2010, 9, 430–435. [Google Scholar] [CrossRef] [PubMed]
  33. Sui, D.; Xu, L.Q.; Zhang, H.T.; Sun, Z.H.; Kan, B.; Ma, Y.F.; Chen, Y.S. A 3D cross-linked graphene-based honeycomb carbon composite with excellent confinement effect of organic cathode material for lithium-ion batteries. Carbon 2020, 157, 656–662. [Google Scholar] [CrossRef]
  34. Li, Y.; Chen, M.H.; Liu, B.; Zhang, Y.; Liang, X.Q.; Xia, X.H. Heteroatom Doping: An Effective Way to Boost Sodium Ion Storage. Adv. Energy Mater. 2020, 10, 2000927. [Google Scholar] [CrossRef]
  35. Wang, H.; Jia, J.; Song, P.; Wang, Q.; Li, D.; Min, S.; Qian, C.; Wang, L.; Li, Y.; Ma, C.; et al. Efficient Electrocatalytic Reduction of CO2 by Nitrogen-Doped Nanoporous Carbon/Carbon Nanotube Membranes: A Step Towards the Electrochemical CO2 Refinery. Angew. Chem. Int. Ed. 2012, 18, 12–15. [Google Scholar]
  36. Xu, C.F.; Zhang, K.W.; Zhang, D.; Chang, S.L.; Liang, F.; Yan, P.F.; Yao, Y.C.; Qu, T.; Zhan, J.; Ma, W.H.; et al. Reversible hybrid sodium-CO2 batteries with low charging voltage and long-life. Nano Energy 2020, 68, 104318. [Google Scholar] [CrossRef]
  37. Li, S.W.; Liu, Y.; Zhou, J.W.; Hong, S.S.; Dong, Y.; Wang, J.M.; Gao, X.; Qi, P.F.; Han, Y.; Wang, B. Monodispersed MnO nanoparticles in graphene-an interconnected N-doped 3D carbon framework as a highly efficient gas cathode in Li-CO2 batteries. Energy Environ. Sci. 2019, 12, 1046–1054. [Google Scholar] [CrossRef]
  38. Dong, H.Y.; Jin, C.; Gao, Y.C.; Xiao, X.L.; Li, K.; Tang, P.P.; Li, X.N.; Yang, S.T. Nitrogen and sulfur co-doped three-dimensional graphene@NiO composite as cathode catalyst for the Li-O2 and Li-CO2 batteries. Mater. Res. Express 2019, 6, 115616. [Google Scholar] [CrossRef]
  39. Sun, Z.M.; Wang, D.; Lin, L.; Liu, Y.H.; Yuan, M.W.; Nan, C.Y.; Li, H.F.; Sun, G.B.; Yang, X.J. Ultrathin hexagonal boron nitride as a van der Waals’ force initiator activated graphene for engineering efficient non-metal electrocatalysts of Li-CO2 battery. Nano Res. 2022, 15, 1171–1177. [Google Scholar] [CrossRef]
  40. Chen, B.; Wang, D.S.; Tan, J.Y.; Liu, Y.Q.; Jiao, M.L.; Liu, B.L.; Zhao, N.Q.; Zou, X.L.; Zhou, G.M.; Cheng, H.M. Designing Electrophilic and Nucleophilic Dual Centers in the ReS2 Plane toward Efficient Bifunctional Catalysts for Li-CO2 Batteries. J. Am. Chem. Soc. 2022, 144, 3106–3116. [Google Scholar] [CrossRef]
  41. Shen, J.R.; Wu, H.T.; Sun, W.; Wu, Q.B.; Zhen, S.Y.; Wang, Z.H.; Sun, K.N. Biomass-derived hierarchically porous carbon skeletons with in situ decorated IrCo nanoparticles as high-performance cathode catalysts for Li-O2 batteries. J. Mater. Chem. A 2019, 7, 10662–10671. [Google Scholar] [CrossRef]
  42. Tong, Z.Z.; Wang, S.B.; Fang, M.H.; Lin, Y.T.; Tsai, K.T.; Tsai, S.Y.; Yin, L.C.; Hu, S.F.; Liu, R.S. Na-CO2 battery with NASICON-structured solid-state electrolyte. Nano Energy 2021, 85, 105972. [Google Scholar] [CrossRef]
  43. Thoka, S.; Tong, Z.Z.; Jena, A.; Hung, T.F.; Wu, C.C.; Chang, W.S.; Wang, F.M.; Wang, X.C.; Yin, L.C.; Chang, H.; et al. High-performance Na-CO2 batteries with ZnCo2O4@CNT as the cathode catalyst. J. Mater. Chem. A 2020, 8, 23974–23982. [Google Scholar] [CrossRef]
  44. Zhu, Y.; Feng, S.; Zhang, P.; Guo, M.; Wang, Q.; Wu, D.; Zhang, L.; Li, H.; Wang, H.; Chen, L.; et al. Probing the electrochemical evolutions of Na-CO2 nanobatteries on Pt@NCNT cathodes using in-situ environmental TEM. Energy Storage Mater. 2020, 33, 88–94. [Google Scholar] [CrossRef]
  45. Xu, C.; Zhan, J.; Wang, Z.; Fang, X.; Chen, J.; Liang, F.; Zhao, H.; Lei, Y. Biomass-derived highly dispersed Co/Co9S8 nanoparticles encapsulated in S, N-co-doped hierarchically porous carbon as an efficient catalyst for hybrid Na-CO2 batteries. Mater. Today Energy 2021, 19, 100594. [Google Scholar] [CrossRef]
  46. Jin, T.; Han, Q.Q.; Jiao, L.F. Binder-Free Electrodes for Advanced Sodium-Ion Batteries. Adv. Mater. 2020, 32, 1806304. [Google Scholar] [CrossRef]
  47. Zhai, Y.J.; Han, P.; Yun, Q.B.; Ge, Y.Y.; Zhang, X.; Chen, Y.; Zhang, H. Phase engineering of metal nanocatalysts for electrochemical CO2 reduction. eScience 2022, 2, 467–485. [Google Scholar] [CrossRef]
  48. Lu, J.; Lee, Y.J.; Luo, X.; Lau, K.C.; Asadi, M.; Wang, H.H.; Brombosz, S.; Wen, J.; Zhai, D.; Chen, Z.; et al. A lithium-oxygen battery based on lithium superoxide. Nature 2016, 529, 377–382. [Google Scholar] [CrossRef] [Green Version]
  49. Qiao, Y.; Xu, S.M.; Liu, Y.; Dai, J.Q.; Xie, H.; Yao, Y.G.; Mu, X.W.; Chen, C.J.; Kline, D.J.; Hitz, E.M.; et al. Transient, in situ synthesis of ultrafine ruthenium nanoparticles for a high-rate Li-CO2 battery. Energy Environ. Sci. 2019, 12, 1100–1107. [Google Scholar] [CrossRef]
  50. Zhang, S.; Wen, Z.; Rui, K.; Shen, C.; Lu, Y.; Yang, J. Graphene nanosheets loaded with Pt nanoparticles with enhanced electrochemical performance for sodium-oxygen batteries. J. Mater. Chem. A 2015, 3, 2568–2571. [Google Scholar] [CrossRef]
  51. Xing, Y.; Yang, Y.; Li, D.; Luo, M.; Chen, N.; Ye, Y.; Qian, J.; Li, L.; Yang, D.; Wu, F.; et al. Crumpled Ir Nanosheets Fully Covered on Porous Carbon Nanofibers for Long-Life Rechargeable Lithium-CO2 Batteries. Adv. Mater. 2018, 30, e1803124. [Google Scholar] [CrossRef] [PubMed]
  52. Wu, G.; Li, X.; Zhang, Z.; Dong, P.; Xu, M.; Peng, H.; Zeng, X.; Zhang, Y.; Liao, S. Design of ultralong-life Li-CO2 batteries with IrO2 nanoparticles highly dispersed on nitrogen-doped carbon nanotubes. J. Mater. Chem. A 2020, 8, 3763–3770. [Google Scholar] [CrossRef]
  53. Wang, X.; Xie, J.; Ghausi, M.A.; Lv, J.; Huang, Y.; Wu, M.; Wang, Y.; Yao, J. Rechargeable Zn-CO2 Electrochemical Cells Mimicking Two-Step Photosynthesis. Adv. Mater. 2019, 31, e1807807. [Google Scholar] [CrossRef]
  54. Ma, W.; Liu, X.; Li, C.; Yin, H.; Xi, W.; Liu, R.; He, G.; Zhao, X.; Luo, J.; Ding, Y. Rechargeable Al-CO2 Batteries for Reversible Utilization of CO2. Adv. Mater. 2018, 30, e1801152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Fang, C.; Luo, J.; Jin, C.; Yuan, H.; Sheng, O.; Huang, H.; Gan, Y.; Xia, Y.; Liang, C.; Zhang, J.; et al. Enhancing Catalyzed Decomposition of Na2CO3 with Co2MnOx Nanowire-Decorated Carbon Fibers for Advanced Na-CO2 Batteries. ACS Appl. Mater. Interfaces 2018, 10, 17240–17248. [Google Scholar] [CrossRef]
  56. Dong, L.Z.; Zhang, Y.; Lu, Y.F.; Zhang, L.; Huang, X.; Wang, J.H.; Liu, J.; Li, S.L.; Lan, Y.Q. A well-defined dual Mn-site based Metal-organic framework to promote CO2 reduction/evolution in Li-CO2 batteries. Chem. Commun. 2021, 57, 8937–8940. [Google Scholar] [CrossRef]
  57. Zhang, Z.; Zhang, Z.; Liu, P.; Xie, Y.; Cao, K.; Zhou, Z. Identification of cathode stability in Li-CO2 batteries with Cu nanoparticles highly dispersed on N-doped graphene. J. Mater. Chem. A 2018, 6, 3218–3223. [Google Scholar] [CrossRef]
  58. Xu, C.; Zhan, J.; Wang, H.; Kang, Y.; Liang, F. Dense binary Fe-Cu sites promoting CO2 utilization enable highly reversible hybrid Na-CO2 batteries. J. Mater. Chem. A 2021, 9, 22114–22128. [Google Scholar] [CrossRef]
  59. Zhang, P.; Zhang, J.; Sheng, T.; Lu, Y.; Yin, Z.; Li, Y.; Peng, X.; Zhou, Y.; Li, J.; Wu, Y.; et al. Synergetic Effect of Ru and NiO in the Electrocatalytic Decomposition of Li2CO3 to Enhance the Performance of a Li-CO2/O2 Battery. ACS Catalysis 2019, 10, 1640–1651. [Google Scholar] [CrossRef]
  60. Xiao, X.; Zhang, Z.; Yu, W.; Shang, W.; Ma, Y.; Zhu, X.; Tan, P. Ultrafine Co-Doped NiO Nanoparticles Decorated on Carbon Nanotubes Improving the Electrochemical Performance and Cycling Stability of Li-CO2 Batteries. ACS Appl. Energy Mater. 2021, 4, 11858–11866. [Google Scholar] [CrossRef]
  61. Ge, B.; Sun, Y.; Guo, J.; Yan, X.; Fernandez, C.; Peng, Q. A Co-Doped MnO2 Catalyst for Li-O2 Batteries with Low Overpotential and Ultrahigh Cyclability. Small 2019, 15, e1902220. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, W.X.; Xiong, F.Y.; Zhu, S.H.; Chen, J.H.; Xie, J.; An, Q.Y. Defect engineering in molybdenum-based electrode materials for energy storage. eScience 2022, 2, 278–294. [Google Scholar] [CrossRef]
  63. Kwak, W.; Lau, K.; Shin, C.; Amine, K.; Curtiss, L.; Sun, Y. A Mo2C/Carbon Nanotube Composite Cathode for Lithium-Oxygen Batteries with High Energy Efficiency and Long Cycle Life. ACS Nano 2015, 9, 4129–4137. [Google Scholar] [CrossRef] [PubMed]
  64. Hou, Y.; Wang, J.; Liu, L.; Liu, Y.; Chou, S.; Shi, D.; Liu, H.; Wu, Y.; Zhang, W.; Chen, J. Mo2C/CNT: An Efficient Catalyst for Rechargeable Li-CO2 Batteries. Adv. Funct. Mater. 2017, 27, 1700564. [Google Scholar] [CrossRef] [Green Version]
  65. Asadi, M.; Sayahpour, B.; Abbasi, P.; Ngo, A.T.; Karis, K.; Jokisaari, J.R.; Liu, C.; Narayanan, B.; Gerard, M.; Yasaei, P.; et al. A lithium-oxygen battery with a long cycle life in an air-like atmosphere. Nature 2018, 555, 502–506. [Google Scholar] [CrossRef]
  66. Senthil Kumar, R.; Senthil Kumar, S.; Anbu Kulandainathan, M. Highly selective electrochemical reduction of carbon dioxide using Cu based metal organic framework as an electrocatalyst. Electrochem. Commun. 2012, 25, 70–73. [Google Scholar] [CrossRef]
  67. Millward, A.R.; Yaghi, O.M. Metal-Organic Frameworks with Exceptionally High Capacity for Storage of Carbon Dioxide at Room Temperature. J. Am. Chem. Soc. 2005, 127, 17998–17999. [Google Scholar] [CrossRef]
  68. Wu, D.; Guo, Z.; Yin, X.; Pang, Q.; Tu, B.; Zhang, L.; Wang, Y.G.; Li, Q. Metal-organic frameworks as cathode materials for Li-O2 batteries. Adv. Mater. 2014, 26, 3258–3262. [Google Scholar] [CrossRef]
  69. Chen, Y.; Ji, S.; Chen, C.; Peng, Q.; Wang, D.; Li, Y. Single-Atom Catalysts: Synthetic Strategies and Electrochemical Applications. Joule 2018, 2, 1242–1264. [Google Scholar] [CrossRef] [Green Version]
  70. Qiao, B.; Wang, A.; Yang, X.; Allard, L.F.; Jiang, Z.; Cui, Y.; Liu, J.; Li, J.; Zhang, T. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat. Chem. 2011, 3, 634–641. [Google Scholar] [CrossRef]
  71. Wang, Y.; Liu, Y.; Liu, W.; Wu, J.; Li, Q.; Feng, Q.; Chen, Z.; Xiong, X.; Wang, D.; Lei, Y. Regulating the coordination structure of metal single atoms for efficient electrocatalytic CO2 reduction. Energy Environ. Sci. 2020, 13, 4609–4624. [Google Scholar] [CrossRef]
  72. Cheng, Y.; Yang, S.; Jiang, S.P.; Wang, S. Supported Single Atoms as New Class of Catalysts for Electrochemical Reduction of Carbon Dioxide. Small Methods 2019, 3, 1800440. [Google Scholar] [CrossRef]
  73. Cheng, Y.; Zhao, S.; Johannessen, B.; Veder, J.P.; Saunders, M.; Rowles, M.R.; Cheng, M.; Liu, C.; Chisholm, M.F.; De Marco, R.; et al. Atomically Dispersed Transition Metals on Carbon Nanotubes with Ultrahigh Loading for Selective Electrochemical Carbon Dioxide Reduction. Adv. Mater. 2018, 30, e1706287. [Google Scholar] [CrossRef]
  74. Shah, S.S.A.; Najam, T.; Bashir, M.S.; Peng, L.; Nazir, M.A.; Javed, M.S. Single-atom catalysts for next-generation rechargeable batteries and fuel cells. Energy Storage Mater. 2022, 45, 301–322. [Google Scholar] [CrossRef]
  75. Xi, J.; Sun, H.; Wang, D.; Zhang, Z.; Duan, X.; Xiao, J.; Xiao, F.; Liu, L.; Wang, S. Confined-interface-directed synthesis of Palladium single-atom catalysts on graphene/amorphous carbon. Appl. Catal. B Environ. 2018, 225, 291–297. [Google Scholar] [CrossRef] [Green Version]
  76. Liu, D.; Wu, C.; Chen, S.; Ding, S.; Xie, Y.; Wang, C.; Wang, T.; Haleem, Y.A.; ur Rehman, Z.; Sang, Y.; et al. In situ trapped high-density single metal atoms within graphene: Iron-containing hybrids as representatives for efficient oxygen reduction. Nano Res. 2018, 11, 2217–2228. [Google Scholar] [CrossRef]
  77. Wang, T.; Zhao, Q.; Fu, Y.; Lei, C.; Yang, B.; Li, Z.; Lei, L.; Wu, G.; Hou, Y. Carbon-Rich Nonprecious Metal Single Atom Electrocatalysts for CO2 Reduction and Hydrogen Evolution. Small Methods 2019, 3, 1900210. [Google Scholar] [CrossRef]
  78. Jiang, K.; Liu, B.; Luo, M.; Ning, S.; Peng, M.; Zhao, Y.; Lu, Y.R.; Chan, T.S.; de Groot, F.M.F.; Tan, Y. Single platinum atoms embedded in nanoporous cobalt selenide as electrocatalyst for accelerating hydrogen evolution reaction. Nat. Commun. 2019, 10, 1743. [Google Scholar] [CrossRef] [Green Version]
  79. Sun, S.; Zhang, G.; Gauquelin, N.; Chen, N.; Zhou, J.; Yang, S.; Chen, W.; Meng, X.; Geng, D.; Banis, M.N.; et al. Single-atom Catalysis Using Pt/Graphene Achieved through Atomic Layer Deposition. Sci. Rep. 2013, 3, 1775. [Google Scholar] [CrossRef] [Green Version]
  80. Liu, X.; Jia, S.; Yang, M.; Tang, Y.; Wen, Y.; Chu, S.; Wang, J.; Shan, B.; Chen, R. Activation of subnanometric Pt on Cu-modified CeO2 via redox-coupled atomic layer deposition for CO oxidation. Nat. Commun. 2020, 11, 4240. [Google Scholar] [CrossRef]
  81. Wei, H.; Liu, X.; Wang, A.; Zhang, L.; Qiao, B.; Yang, X.; Huang, Y.; Miao, S.; Liu, J.; Zhang, T. FeOx-supported platinum single-atom and pseudo-single-atom catalysts for chemoselective hydrogenation of functionalized nitroarenes. Nat. Commun. 2014, 5, 5634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Sun, X.; Chen, C.; Liu, S.; Hong, S.; Zhu, Q.; Qian, Q.; Han, B.; Zhang, J.; Zheng, L. Aqueous CO2 Reduction with High Efficiency Using alpha-Co(OH)2-Supported Atomic Ir Electrocatalysts. Angew. Chem. Int. Ed. 2019, 58, 4669–4673. [Google Scholar] [CrossRef] [PubMed]
  83. Xu, L.; Chen, X.; Jing, H.; Wang, L.; Wei, J.; Han, Y. Design and performance of Ag nanoparticle-modified graphene/SnAgCu lead-free solders. Mater. Sci. Eng. A 2016, 667, 87–96. [Google Scholar] [CrossRef]
  84. Yan, D.; Chen, J.; Jia, H. Temperature-Induced Structure Reconstruction to Prepare a Thermally Stable Single-Atom Platinum Catalyst. Angew. Chem. Int. Ed. 2020, 59, 13562–13567. [Google Scholar] [CrossRef]
  85. Zhao, C.; Wang, Y.; Li, Z.; Chen, W.; Xu, Q.; He, D.; Xi, D.; Zhang, Q.; Yuan, T.; Qu, Y.; et al. Solid-Diffusion Synthesis of Single-Atom Catalysts Directly from Bulk Metal for Efficient CO2 Reduction. Joule 2019, 3, 584–594. [Google Scholar] [CrossRef] [Green Version]
  86. Wen, X.; Zhang, Q.; Guan, J. Applications of metal-organic framework-derived materials in fuel cells and metal-air batteries. Coord. Chem. Rev. 2020, 409, 213214. [Google Scholar] [CrossRef]
  87. Han, A.; Chen, W.; Zhang, S.; Zhang, M.; Han, Y.; Zhang, J.; Ji, S.; Zheng, L.; Wang, Y.; Gu, L.; et al. A Polymer Encapsulation Strategy to Synthesize Porous Nitrogen-Doped Carbon-Nanosphere-Supported Metal Isolated-Single-Atomic-Site Catalysts. Adv. Mater. 2018, 30, e1706508. [Google Scholar] [CrossRef]
  88. Wang, J.; Huang, Z.; Liu, W.; Chang, C.; Tang, H.; Li, Z.; Chen, W.; Jia, C.; Yao, T.; Wei, S.; et al. Design of N-Coordinated Dual-Metal Sites: A Stable and Active Pt-Free Catalyst for Acidic Oxygen Reduction Reaction. J. Am. Chem. Soc. 2017, 139, 17281–17284. [Google Scholar] [CrossRef]
  89. Liu, J.J. Advanced Electron Microscopy of Metal-Support Interactions in Supported Metal Catalysts. ChemCatChem 2011, 3, 934–948. [Google Scholar] [CrossRef]
  90. Wang, X.; Pi, M.; Zhang, D.; Li, H.; Feng, J.; Chen, S.; Li, J. Insight into the Superior Electrocatalytic Performance of a Ternary Nickel Iron Poly-Phosphide Nanosheet Array: An X-ray Absorption Study. ACS Appl. Mater. Interfaces 2019, 11, 14059–14065. [Google Scholar] [CrossRef]
  91. Cheng, N.; Zhang, L.; Doyle-Davis, K.; Sun, X. Single-Atom Catalysts: From Design to Application. Electrochem. Energy Rev. 2019, 2, 539–573. [Google Scholar] [CrossRef] [Green Version]
  92. Zhu, Y.; Wang, J.; Chu, H.; Chu, Y.; Chen, H.M. In Situ/Operando Studies for Designing Next-Generation Electrocatalysts. ACS Energy Lett. 2020, 5, 1281–1291. [Google Scholar] [CrossRef]
  93. Hui, Y.; Liu, Y.F.; Liu, X.L.; Wang, X.K.; Tian, H.; Waterhouse, G.I.; Kruger, P.E.; Telfer, S.G.; Ma, S.Q. Large-scale synthesis of N-doped carbon capsules supporting atomically dispersed iron for efficient oxygen reduction reaction electrocatalysis. eScience 2022, 2, 227–234. [Google Scholar]
  94. Rao, P.; Wu, D.X.; Wang, T.J.; Li, J.; Deng, P.L.; Chen, Q.; Shen, Y.J.; Chen, Y.; Tian, X.L. Single atomic cobalt electrocatalyst for efficient oxygen reduction reaction. eScience 2022, 2, 399–404. [Google Scholar] [CrossRef]
  95. Zheng, C.Y.; Zhang, X.; Zhou, Z.; Hu, Z.P. A first-principles study on the electrochemical reaction activity of 3d transition metal single-atom catalysts in nitrogen-doped graphene: Trends and hints. eScience 2022, 2, 219–226. [Google Scholar] [CrossRef]
  96. Thangadurai, V.; Narayanan, S.; Pinzaru, D. Garnet-type solid-state fast Li ion conductors for Li batteries: Critical review. Chem. Soc. Rev. 2014, 43, 4714–4727. [Google Scholar] [CrossRef] [PubMed]
  97. Lu, Y.; Meng, X.; Alonso, J.A.; Fernandez-Diaz, M.T.; Sun, C. Effects of Fluorine Doping on Structural and Electrochemical Properties of Li6.25Ga0.25La3Zr2O12 as Electrolytes for Solid-State Lithium Batteries. ACS Appl. Mater. Interfaces 2019, 11, 2042–2049. [Google Scholar] [CrossRef]
  98. Thangadurai, V.; Pinzaru, D.; Narayanan, S.; Baral, A.K. Fast Solid-State Li Ion Conducting Garnet-Type Structure Metal Oxides for Energy Storage. J. Phys. Chem. Lett. 2015, 6, 292–299. [Google Scholar] [CrossRef]
  99. Wang, Y.; Song, S.; Xu, C.; Hu, N.; Molenda, J.; Lu, L. Development of solid-state electrolytes for sodium-ion battery-A short review. Nano Mater. Sci. 2019, 1, 91–100. [Google Scholar] [CrossRef]
  100. Lu, Y.; Li, L.; Zhang, Q.; Niu, Z.; Chen, J. Electrolyte and Interface Engineering for Solid-State Sodium Batteries. Joule 2018, 2, 1747–1770. [Google Scholar] [CrossRef] [Green Version]
  101. Bachman, J.; Muy, S.; Grimaud, A.; Chang, H.; Pour, N.; Lux, S.; Paschos, O.; Maglia, F.; Lupart, S.; Lamp, P.; et al. Inorganic Solid-State Electrolytes for Lithium Batteries: Mechanisms and Properties Governing Ion Conduction. Chem. Rev. 2016, 116, 140–162. [Google Scholar] [CrossRef] [PubMed]
  102. Famprikis, T.; Canepa, P.; Dawson, J.A.; Islam, M.S.; Masquelier, C. Fundamentals of inorganic solid-state electrolytes for batteries. Nat. Mater. 2019, 18, 1278–1291. [Google Scholar] [CrossRef] [PubMed]
  103. Zhang, Z.; Zhang, Q.; Shi, J.; Chu, Y.S.; Yu, X.; Xu, K.; Ge, M.; Yan, H.; Li, W.; Gu, L.; et al. A Self-Forming Composite Electrolyte for Solid-State Sodium Battery with Ultralong Cycle Life. Adv. Energy Mater. 2017, 7, 1601196. [Google Scholar] [CrossRef]
  104. Yue, L.; Ma, J.; Zhang, J.; Zhao, J.; Dong, S.; Liu, Z.; Cui, G.; Chen, L. All solid-state polymer electrolytes for high-performance lithium ion batteries. Energy Storage Mater. 2016, 5, 139–164. [Google Scholar] [CrossRef]
  105. Song, S.; Duong, H.M.; Korsunsky, A.M.; Hu, N.; Lu, L. A Na(+) Superionic Conductor for Room-Temperature Sodium Batteries. Sci. Rep. 2016, 6, 32330. [Google Scholar] [CrossRef]
  106. Zhu, Z.; Chu, I.-H.; Deng, Z.; Ong, S.P. Role of Na+ Interstitials and Dopants in Enhancing the Na+ Conductivity of the Cubic Na3PS4 Superionic Conductor. Chem. Mater. 2015, 27, 8318–8325. [Google Scholar] [CrossRef] [Green Version]
  107. Park, H.; Jung, K.; Nezafati, M.; Kim, C.S.; Kang, B. Sodium Ion Diffusion in Nasicon (Na3Zr2Si2PO12) Solid Electrolytes: Effects of Excess Sodium. ACS Appl. Mater. Interfaces 2016, 8, 27814–27824. [Google Scholar] [CrossRef]
  108. Ma, Q.; Guin, M.; Naqash, S.; Tsai, C.-L.; Tietz, F.; Guillon, O. Scandium-Substituted Na3Zr2(SiO4)2(PO4) Prepared by a Solution-Assisted Solid-State Reaction Method as Sodium-Ion Conductors. Chem. Mater. 2016, 28, 4821–4828. [Google Scholar] [CrossRef]
  109. Ruan, Y.; Song, S.; Liu, J.; Liu, P.; Cheng, B.; Song, X.; Battaglia, V. Improved structural stability and ionic conductivity of Na3Zr2Si2PO12 solid electrolyte by rare earth metal substitutions. Ceram. Int. 2017, 43, 7810–7815. [Google Scholar] [CrossRef]
  110. Goodenough, J.B.; Hong, H.; Kafalas, J. Fast Na+-ion transport in skeleton structures. Mater. Res. Bull. 1976, 11, 203–220. [Google Scholar] [CrossRef]
  111. Hong, Y. Crystal structures and crystal chemistry in the system Na1+xZr2SixP3−xO1. Mater. Res. Bull. 1976, 11, 173–182. [Google Scholar] [CrossRef]
  112. Im, E.; Ryu, J.H.; Baek, K.; Moon, G.D.; Kang, S.J. “Water-in-salt” and NASICON Electrolyte-Based Na-CO2 Battery. Energy Storage Mater. 2021, 37, 424–432. [Google Scholar] [CrossRef]
  113. Lu, L.; Sun, C.; Hao, J.; Wang, Z.; Mayer, S.F.; Fernández-Díaz, M.T.; Alonso, J.A.; Zou, B. A High-Performance Solid-State Na-CO2 Battery with Poly(Vinyliden Fluoride-co-Hexafluoropropylene)-Na3.2Zr1.9Mg0.1Si2PO12 Electrolyte. Energy Environ. Mater. 2022, 1–9. [Google Scholar]
  114. Jung, J.I.; Kim, D.; Kim, H.; Jo, Y.N.; Park, J.S.; Kim, Y. Progressive Assessment on the Decomposition Reaction of Na Superionic Conducting Ceramics. ACS Appl. Mater. Interfaces 2017, 9, 304–310. [Google Scholar] [CrossRef] [PubMed]
  115. Sun, C.; Liu, J.; Gong, Y.; Wilkinson, D.P.; Zhang, J. Recent advances in all-solid-state rechargeable lithium batteries. Nano Energy 2017, 33, 363–386. [Google Scholar] [CrossRef] [Green Version]
  116. Lu, L.; Lu, Y.; Alonso, J.A.; Lopez, C.A.; Fernandez-Diaz, M.T.; Zou, B.; Sun, C. A Monolithic Solid-State Sodium-Sulfur Battery with Al-Doped Na3.4Zr2(Si0.8P0.2O4)3 Electrolyte. ACS Appl. Mater. Interfaces 2021, 13, 42927–42934. [Google Scholar] [CrossRef]
  117. Shen, L.; Yang, J.; Liu, G.; Avdeev, M.; Yao, X. High ionic conductivity and dendrite-resistant NASICON solid electrolyte for all-solid-state sodium batteries. Mater. Today Energy 2021, 20, 100691. [Google Scholar] [CrossRef]
  118. Ihlefeld, J.F.; Gurniak, E.; Jones, B.H.; Wheeler, D.R.; Rodriguez, M.A.; McDaniel, A.H.; Dunn, B. Scaling Effects in Sodium Zirconium Silicate Phosphate (Na1+xZr2SixP3−xO12) Ion-Conducting Thin Films. J. Am. Ceram. Soc. 2016, 99, 2729–2736. [Google Scholar] [CrossRef]
  119. Yao, Y.; Kummer, J. Ion exchange properties of and rates of ionic diffusion in beta-alumina. J. Inorg. Nucl. Chem. 1967, 29, 2453–2475. [Google Scholar]
  120. Hueso, K.B.; Palomares, V.; Armand, M.; Rojo, T. Challenges and perspectives on high and intermediate-temperature sodium batteries. Nano Res. 2017, 10, 4082–4114. [Google Scholar] [CrossRef]
  121. Lu, X.; Xia, G.; Lemmon, J.P.; Yang, Z. Advanced Materials for sodium-beta alumina batteries: Status, challenges and perspectives. J. Power Source 2010, 195, 2431–2442. [Google Scholar] [CrossRef]
  122. Baffier, N.; Badot, J.; Colomban, P. Conductivity of ion rich β and β″ alumina: Sodium and potassium compounds. Mater. Res. Bull. 1981, 16, 259–265. [Google Scholar] [CrossRef]
  123. Wu, Y.; Zhuo, L.; Ming, J.; Yu, Y.; Zhao, F. Coating of Al2O3 on layered Li(Mn1/3Ni1/3Co1/3)O2 using CO2 as green precipitant and their improved electrochemical performance for lithium ion batteries. J. Energy Chem. 2013, 22, 468–476. [Google Scholar] [CrossRef]
  124. Ghadbeigi, L.; Szendrei, A.; Moreno, P.; Sparks, T.D.; Virkar, A.V. Synthesis of iron-doped Na-β’’-alumina+yttria-stabilized zirconia composite electrolytes by a vapor phase process. Solid State Ion. 2016, 290, 77–82. [Google Scholar] [CrossRef] [Green Version]
  125. Virkar, A.V.; Gordon, R.S. Fracture Properties of Polycrystalline Lithia-Stabilized β’’ Alumina. J. Am. Ceram. Soc. 1977, 60, 58–61. [Google Scholar] [CrossRef]
  126. Wang, Y.; Wang, Z.X.; Wu, D.X.; Niu, Q.H.; Lu, P.S.; Ma, T.H.; Su, Y.B.; Chen, L.Q.; Li, H.; Wu, F. Stable Ni-rich layered oxide cathode for sulfide-based all-solid-state lithium battery. eScience 2022, 2, 537–545. [Google Scholar] [CrossRef]
  127. Hayashi, A.; Noi, K.; Sakuda, A.; Tatsumisago, M. Superionic glass-ceramic electrolytes for room-temperature rechargeable sodium batteries. Nat. Commun. 2012, 3, 856. [Google Scholar] [CrossRef] [Green Version]
  128. Shang, S.L.; Yu, Z.X.; Wang, Y.; Wang, D.H.; Liu, Z.K. Origin of Outstanding Phase and Moisture Stability in a Na3P1–xAsxS4 Superionic Conductor. ACS Appl. Mater. Interfaces 2017, 9, 16261–16269. [Google Scholar] [CrossRef]
  129. Tian, Y.; Shi, T.; Richards, W.D.; Li, J.; Kim, J.C.; Bo, S.-H.; Ceder, G. Compatibility issues between electrodes and electrolytes in solid-state batteries. Energy Environ. Sci. 2017, 10, 1150–1166. [Google Scholar] [CrossRef] [Green Version]
  130. Zhang, L.; Yang, K.; Mi, J.; Lu, L.; Zhao, L.; Wang, L.; Li, Y.; Zeng, H. Na3PSe4: A Novel Chalcogenide Solid Electrolyte with High Ionic Conductivity. Adv. Energy Mater. 2015, 5, 1501294. [Google Scholar] [CrossRef]
  131. Nishimura, S.; Tanibata, N.; Hayashi, A.; Tatsumisago, M.; Yamada, A. The crystal structure and sodium disorder of high-temperature polymorph β-Na3PS4. J. Mater. Chem. A 2017, 5, 25025–25030. [Google Scholar] [CrossRef]
  132. Jansen, M.; Henseler, U. Synthesis, Structure Determination, and Ionic Conductivity of Sodium Tetrathiophosphate. J. Solid State Chem. 1992, 99, 110–119. [Google Scholar] [CrossRef]
  133. Yu, Z.; Shang, S.L.; Seo, J.H.; Wang, D.; Luo, X.; Huang, Q.; Chen, S.; Lu, J.; Li, X.; Liu, Z.K.; et al. Exceptionally High Ionic Conductivity in Na3P0.62As0.38S4 with Improved Moisture Stability for Solid-State Sodium-Ion Batteries. Adv. Mater. 2017, 29, 1605561. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, L.; Zhang, D.; Yang, K.; Yan, X.; Wang, L.; Mi, J.; Xu, B.; Li, Y. Vacancy-Contained Tetragonal Na3SbS4 Superionic Conductor. Adv. Sci. 2016, 3, 1600089. [Google Scholar] [CrossRef]
  135. Wang, Y.; Yang, N.; Shuai, Y.; Chen, K. Low-Cost Raw Materials Synthesized Na2.9375PS3.9375Cl0.0625 Solid Electrolyte. Energy Technol. 2020, 8, 2000606. [Google Scholar] [CrossRef]
  136. de Klerk, N.J.J.; Wagemaker, M. Diffusion Mechanism of the Sodium-Ion Solid Electrolyte Na3PS4 and Potential Improvements of Halogen Doping. Chem. Mater. 2016, 28, 3122–3130. [Google Scholar] [CrossRef] [Green Version]
  137. Krauskopf, T.; Pompe, C.; Kraft, M.A.; Zeier, W.G. Influence of Lattice Dynamics on Na+ Transport in the Solid Electrolyte Na3PS4–xSex. Chem. Mater. 2017, 29, 8859–8869. [Google Scholar] [CrossRef]
  138. Wang, H.; Chen, Y.; Hood, Z.D.; Sahu, G.; Pandian, A.S.; Keum, J.K.; An, K.; Liang, C. An Air-Stable Na3SbS4 Superionic Conductor Prepared by a Rapid and Economic Synthetic Procedure. Angew. Chem. Int. Ed. Engl. 2016, 55, 8551–8555. [Google Scholar] [CrossRef]
  139. Zhang, Z.; Zhang, J.; Sun, Y.; Jia, H.; Peng, L.; Zhang, Y.; Xie, J. Li4−xSbxSn1−xS4 solid solutions for air-stable solid electrolytes. J. Energy Chem. 2020, 41, 171–176. [Google Scholar] [CrossRef] [Green Version]
  140. Oguchi, H.; Matsuo, M.; Kuromoto, S.; Kuwano, H.; Orimo, S. Sodium-ion conduction in complex hydrides NaAlH4 and Na3AlH6. J. Appl. Phys. 2012, 111, 036102. [Google Scholar] [CrossRef]
  141. Matsuo, M.; Oguchi, H.; Sato, T.; Takamura, H.; Tsuchida, E.; Ikeshoji, T.; Orimo, S.-I. Sodium and magnesium ionic conduction in complex hydrides. J. Alloys Compd. 2013, 580, S98–S101. [Google Scholar] [CrossRef]
  142. Udovic, T.J.; Matsuo, M.; Unemoto, A.; Verdal, N.; Stavila, V.; Skripov, A.V.; Rush, J.J.; Takamura, H.; Orimo, S. Sodium superionic conduction in Na2B12H12. Chem. Commun. 2014, 50, 3750–3752. [Google Scholar] [CrossRef] [PubMed]
  143. Verdal, N.; Udovic, T.J.; Stavila, V.; Tang, W.S.; Rush, J.J.; Skripov, A.V. Anion Reorientations in the Superionic Conducting Phase of Na2B12H12. J. Phys. Chem. C 2014, 118, 17483–17489. [Google Scholar] [CrossRef]
  144. Tang, W.S.; Yoshida, K.; Soloninin, A.V.; Skoryunov, R.V.; Babanova, O.A.; Skripov, A.V.; Dimitrievska, M.; Stavila, V.; Orimo, S.-I.; Udovic, T.J. Stabilizing Superionic-Conducting Structures via Mixed-Anion Solid Solutions of Monocarba-closo-borate Salts. ACS Energy Lett. 2016, 1, 659–664. [Google Scholar] [CrossRef]
  145. Tang, W.S.; Matsuo, M.; Wu, H.; Stavila, V.; Zhou, W.; Talin, A.A.; Soloninin, A.V.; Skoryunov, R.V.; Babanova, O.A.; Skripov, A.V.; et al. Liquid-Like Ionic Conduction in Solid Lithium and Sodium Monocarba-closo-Decaborates Near or at Room Temperature. Adv. Energy Mater. 2016, 6, 1502237. [Google Scholar] [CrossRef]
  146. Duchêne, L.; Kühnel, R.S.; Rentsch, D.; Remhof, A.; Hagemann, H.; Battaglia, C. A highly stable sodium solid-state electrolyte based on a dodeca/deca-borate equimolar mixture. Chem. Commun. 2017, 53, 4195–4198. [Google Scholar] [CrossRef] [Green Version]
  147. Duchêne, L.; Kühnel, R.S.; Stilp, E.; Cuervo Reyes, E.; Remhof, A.; Hagemann, H.; Battaglia, C. A stable 3 V all-solid-state sodium–ion battery based on a closo-borate electrolyte. Energy Environ. Sci. 2017, 10, 2609–2615. [Google Scholar] [CrossRef]
  148. Udovic, T.J.; Matsuo, M.; Tang, W.S.; Wu, H.; Stavila, V.; Soloninin, A.V.; Skoryunov, R.V.; Babanova, O.A.; Skripov, A.V.; Rush, J.J.; et al. Exceptional superionic conductivity in disordered sodium decahydro-closo-decaborate. Adv. Mater. 2014, 26, 7622–7626. [Google Scholar] [CrossRef]
  149. Sun, Y.; Wang, Y.; Liang, X.; Xia, Y.; Peng, L.; Jia, H.; Li, H.; Bai, L.; Feng, J.; Jiang, H.; et al. Rotational Cluster Anion Enabling Superionic Conductivity in Sodium-Rich Antiperovskite Na3OBH4. J. Am. Chem. Soc. 2019, 141, 5640–5644. [Google Scholar] [CrossRef]
  150. Dimitrievska, M.; Shea, P.; Kweon, K.E.; Bercx, M.; Varley, J.B.; Tang, W.S.; Skripov, A.V.; Stavila, V.; Udovic, T.J.; Wood, B.C. Carbon Incorporation and Anion Dynamics as Synergistic Drivers for Ultrafast Diffusion in Superionic LiCB11H12 and NaCB11H12. Adv. Energy Mater. 2018, 8, 1703422. [Google Scholar] [CrossRef]
  151. Tang, W.S.; Matsuo, M.; Wu, H.; Stavila, V.; Unemoto, A.; Orimo, S.-I.; Udovic, T.J. Stabilizing lithium and sodium fast-ion conduction in solid polyhedral-borate salts at device-relevant temperatures. Energy Storage Mater. 2016, 4, 79–83. [Google Scholar] [CrossRef] [Green Version]
  152. Fenton, D.E.; Parker, J.M.; Wright, P.V. Complexes of alkali metal ions with poly(ethylene oxide). Polymer 1973, 14, 589. [Google Scholar] [CrossRef]
  153. Feuillade, G.; Perche, P. Ion-conductive macromolecular gels and membranes for solid lithium cells. J. Appl. Electrochem. 1975, 5, 63–69. [Google Scholar] [CrossRef]
  154. Chen, R.; Qu, W.; Guo, X.; Li, L.; Wu, F. The pursuit of solid-state electrolytes for lithium batteries: From comprehensive insight to emerging horizons. Mater. Horiz. 2016, 3, 487–516. [Google Scholar] [CrossRef]
  155. Xue, Z.; He, D.; Xie, X. Poly(ethylene oxide)-based electrolytes for lithium-ion batteries. J. Mater. Chem. A 2015, 3, 19218–19253. [Google Scholar] [CrossRef]
  156. Ratner, M.A.; Johansson, P.; Shriver, D.F. Polymer Electrolytes: Ionic Transport Mechanisms and Relaxation Coupling. MRS Bull. 2000, 3, 31–37. [Google Scholar] [CrossRef]
  157. Zhang, Q.; Liu, K.; Ding, F.; Liu, X. Recent advances in solid polymer electrolytes for lithium batteries. Nano Res. 2017, 10, 4139–4174. [Google Scholar] [CrossRef]
  158. Hou, H.D.; Xu, Q.K.; Pang, Y.K.; Li, L.; Wang, J.L.; Zhang, C.; Sun, C.W. Efficient Storing Energy Harvested by Triboelectric Nanogenerators Using a Safe and Durable All-solid-state Sodium-ion Battery. Adv. Sci. 2017, 4, 1700072. [Google Scholar] [CrossRef]
  159. Hu, X.; Li, Z.; Zhao, Y.; Sun, J.; Zhao, Q.; Wang, J.; Tao, Z.; Chen, J. Quasi-solid state rechargeable Na-CO2 batteries with reduced graphene oxide Na anodes. Sci. Adv. 2017, 3, e1602396. [Google Scholar] [CrossRef] [Green Version]
  160. Nagasubramanian, G.; Stefano, S.D. 12-Crown-4 Ether-Assisted Enhancement of Ionic Conductivity and Interfacial Kinetics in Polyethylene Oxide Electrolytes. J. Electrochem. Soc. 1990, 12, 3830. [Google Scholar] [CrossRef]
  161. Song, J.; Wang, Y.; Wan, C. Review of gel-type polymer electrolytes for lithium-ion batteries. J. Power Source 1999, 2, 183–197. [Google Scholar] [CrossRef]
  162. Yu, Q.; Lu, Q.; Qi, X.; Zhao, S.; He, Y.-B.; Liu, L.; Li, J.; Zhou, D.; Hu, Y.-S.; Yang, Q.-H.; et al. Liquid electrolyte immobilized in compact polymer matrix for stable sodium metal anodes. Energy Storage Mater. 2019, 23, 610–616. [Google Scholar] [CrossRef]
  163. Zhou, D.; He, Y.-B.; Cai, Q.; Qin, X.; Li, B.; Du, H.; Yang, Q.-H.; Kang, F. Investigation of cyano resin-based gel polymer electrolyte: In situ gelation mechanism and electrode-electrolyte interfacial fabrication in lithium-ion battery. J. Mater. Chem. A 2014, 2, 20059–20066. [Google Scholar] [CrossRef]
  164. Zhang, W.; Yi, Q.; Li, S.; Sun, C. An ion-conductive Li7La3Zr2O12-based composite membrane for dendrite-free lithium metal batteries. J. Power Source 2020, 450, 227710. [Google Scholar] [CrossRef]
  165. Skaarup, S.; West, K.; Zachau-Christiansen, B. Mixed phase solid electrolytes. Solid State Ion. 1988, 28–30, 975–978. [Google Scholar] [CrossRef]
  166. Wieczorek, W. Modifications of crystalline structure of peo polymer electrolytes with ceramic additives. Solid State Ion. 1989, 3–4, 255–257. [Google Scholar] [CrossRef]
  167. Liu, L.; Qi, X.; Yin, S.; Zhang, Q.; Liu, X.; Suo, L.; Li, H.; Chen, L.; Hu, Y.-S. In Situ Formation of a Stable Interface in Solid-State Batteries. ACS Energy Lett. 2019, 4, 1650–1657. [Google Scholar] [CrossRef]
  168. Ni’mah, Y.L.; Cheng, M.-Y.; Cheng, J.H.; Rick, J.; Hwang, B.-J. Solid-state polymer nanocomposite electrolyte of TiO2/PEO/NaClO4 for sodium ion batteries. J. Power Source 2015, 278, 375–381. [Google Scholar] [CrossRef]
  169. Thakur, A.K.; Upadhyaya, H.M.; Hashmi, S.A. Polyethylene oxide based sodium ion conducting composite polymer electrolytes dispersed with Na2SiO3. Indian J. Pure Appl. Phys. 1999, 37, 302–305. [Google Scholar]
  170. Zhang, Z.; Zhang, Q.; Ren, C.; Luo, F.; Ma, Q.; Hu, Y.-S.; Zhou, Z.; Li, H.; Huang, X.; Chen, L. A ceramic/polymer composite solid electrolyte for sodium batteries. J. Mater. Chem. A 2016, 4, 15823–15828. [Google Scholar] [CrossRef]
  171. Zhang, Q.Q.; Su, X.; Lu, Y.X.; Hu, Y.S. A Composite Solid-state Polymer Electrolyte for Solid-state Sodium Batteries. J. Chin. Ceram. Soc. 2020, 48, 939–946. [Google Scholar]
  172. Colò, F.; Bella, F.; Nair, J.R.; Destro, M.; Gerbaldi, C. Cellulose-based novel hybrid polymer electrolytes for green and efficient Na-ion batteries. Electrochim. Acta 2015, 174, 185–190. [Google Scholar] [CrossRef]
  173. Gao, H.; Guo, B.; Song, J.; Park, K.; Goodenough, J.B. A Composite Gel-Polymer/Glass-Fiber Electrolyte for Sodium-Ion Batteries. Adv. Energy Mater. 2015, 5, 1402235. [Google Scholar] [CrossRef]
  174. Villaluenga, I.; Bogle, X.; Greenbaum, S.; Gil de Muro, I.; Rojo, T.; Armand, M. Cation only conduction in new polymer-SiO2 nanohybrids: Na+ electrolytes. J. Mater. Chem. A 2013, 1, 8348–8352. [Google Scholar] [CrossRef]
  175. Lu, Y.; Alonso, J.A.; Yi, Q.; Lu, L.; Wang, Z.L.; Sun, C.W. A High-performance Monolithic Solid-State Sodium Battery with Ca2+ Doped Na3Zr2Si2PO12 Electrolyte. Adv. Energy Mater. 2019, 9, 1901205. [Google Scholar] [CrossRef]
  176. Zhang, X.; Sun, C.W. Recent Advances in Dendrite-free Lithium Metal Anode for High-Performance Batteries. Phys. Chem. Chem. Phys. 2022, 24, 19996–20011. [Google Scholar] [CrossRef] [PubMed]
  177. Yang, G.P.; Li, N.W.; Sun, C.W. High-performance sodium metal batteries with sodium-bismuth alloy anode. ACS Appl. Energy Mater. 2020, 3, 12607–12612. [Google Scholar] [CrossRef]
  178. Lu, Y.; Lu, L.; Qiu, G.R.; Sun, C.W. Flexible Quasi-Solid-State Sodium Battery for Storing Pulse Electricity Harvested by Triboelectric Nanogenerators. ACS Appl. Mater. Interfaces 2020, 12, 39342–39351. [Google Scholar] [CrossRef]
  179. Zhou, W.; Li, Y.; Xin, S.; Goodenough, J.B. Rechargeable Sodium All-Solid-State Battery. ACS Cent. Sci. 2017, 3, 52–57. [Google Scholar] [CrossRef]
  180. Yi, Q.; Lu, Y.; Sun, X.R.; Zhang, H.; Yu, H.L.; Sun, C.W. Fluorinated ether based electrolyte with low flammability enabling sodium-metal batteries with exceptional cycling stability. ACS Appl. Mater. Interfaces 2019, 11, 46965–46972. [Google Scholar] [CrossRef]
  181. Chen, D.; Zhu, M.; Kang, P.; Zhu, T.; Yuan, H.; Lan, J.; Yang, X.; Sui, G. Self-Enhancing Gel Polymer Electrolyte by In Situ Construction for Enabling Safe Lithium Metal Battery. Adv. Sci. 2022, 9, e2103663. [Google Scholar] [CrossRef] [PubMed]
  182. Wang, J.N.; Ma, Q.Y.; Sun, S.Y.; Yang, K.; Cai, Q.; Olsson, E.; Chen, X.; Wang, Z.; Abdelkader, A.M.; Li, Y.S.; et al. Highly aligned lithiophilic electrospun nanofiber membrane for the multiscale suppression of Li dendrite growth. eScience 2022, 2, 655–665. [Google Scholar] [CrossRef]
  183. Yi, Q.; Zhang, W.Q.; Li, S.Q.; Sun, C.W. A durable sodium battery with a flexible Na3Zr2Si2PO12-PVDF-HFP composite electrolyte and sodium/carbon cloth anode. ACS Appl. Mater. Interfaces 2018, 10, 35039–35046. [Google Scholar] [CrossRef] [PubMed]
  184. Zhao, Q.; Stalin, S.; Zhao, C.Z.; Archer, L.A. Designing solid-state electrolytes for safe, energy-dense batteries. Nat. Rev. Mater. 2020, 5, 229–252. [Google Scholar] [CrossRef]
  185. Feng, L.L.; Yu, T.S.; Cheng, D.H.; Sun, C.W. Progress of Sodium Battery Failure Research. Sci. Sin. Chim. 2020, 50, 1801–1815. [Google Scholar]
  186. Eng, S.Y.S.; Soni, C.B.; Lum, Y.; Khoo, E.; Yao, Z.P.; Vineeth, S.K.; Kumar, V.; Lu, J.; Johnson, C.S.; Wolverton, C.; et al. Theory-guided experimental design in battery materials research. Sci. Adv. 2022, 8, eabm2422. [Google Scholar] [CrossRef]
  187. Liu, Q.; Tang, Y.; Sun, H.; Yang, T.; Sun, Y.; Du, C.; Jia, P.; Ye, H.; Chen, J.; Peng, Q.; et al. In Situ Electrochemical Study of Na-O2/CO2 Batteries in an Environmental Transmission Electron Microscope. ACS Nano 2020, 14, 13232–13245. [Google Scholar] [CrossRef]
  188. Yang, M.; Bi, R.; Wang, J.; Yu, R.; Wang, D. Decoding lithium batteries through advanced in situ characterization techniques. Int. J. Miner. Metall. Mater. 2022, 29, 965–989. [Google Scholar] [CrossRef]
  189. Dixit, M.B.; Park, J.-S.; Kenesei, P.; Almer, J.; Hatzell, K.B. Status and prospect of in situ and operando characterization of solid-state batteries. Energy Environ. Sci. 2021, 14, 4672–4711. [Google Scholar] [CrossRef]
  190. Hou, D.; Xia, D.; Gabriel, E.; Russell, J.A.; Graff, K.; Ren, Y.; Sun, C.-J.; Lin, F.; Liu, Y.; Xiong, H. Spatial and Temporal Analysis of Sodium-Ion Batteries. ACS Energy Letters 2021, 6, 4023–4054. [Google Scholar] [CrossRef]
  191. Liu, D.; Shadike, Z.; Lin, R.; Qian, K.; Li, H.; Li, K.; Wang, S.; Yu, Q.; Liu, M.; Ganapathy, S.; et al. Review of Recent Development of In Situ/Operando Characterization Techniques for Lithium Battery Research. Adv. Mater. 2019, 31, e1806620. [Google Scholar] [CrossRef] [PubMed]
  192. Gao, F.; Tian, X.D.; Lin, J.S.; Dong, J.C.; Lin, X.M.; Li, J.F. In situ Raman, FTIR, and XRD spectroscopic studies in fuel cells and rechargeable batteries. Nano Res. 2022. [Google Scholar] [CrossRef]
  193. Bak, S.-M.; Shadike, Z.; Lin, R.; Yu, X.; Yang, X.-Q. In situ/operando synchrotron-based X-ray techniques for lithium-ion battery research. NPG Asia Mater. 2018, 10, 563–580. [Google Scholar] [CrossRef] [Green Version]
  194. Bragg, W. The structure of some crystals as indicated by their diffraction of X-rays. Proc. R. Soc. Lond. Ser. A Contain. Pap. A Math. Phys. Character 1997, 89, 248–277. [Google Scholar]
  195. Ameh, E.S. A review of basic crystallography and x-ray diffraction applications. Int. J. Adv. Manuf. Technol. 2019, 105, 3289–3302. [Google Scholar] [CrossRef]
  196. Xia, M.; Liu, T.; Peng, N.; Zheng, R.; Cheng, X.; Zhu, H.; Yu, H.; Shui, M.; Shu, J. Lab-Scale In Situ X-ray Diffraction Technique for Different Battery Systems: Designs, Applications, and Perspectives. Small Methods 2019, 3, 1900119. [Google Scholar] [CrossRef]
  197. Yu, W.; Yang, W.; Liu, R.; Qin, L.; Lei, Y.; Liu, L.; Zhai, D.; Li, B.; Kang, F. A soluble phenolic mediator contributing to enhanced discharge capacity and low charge overpotential for lithium-oxygen batteries. Electrochem. Commun. 2017, 79, 68–72. [Google Scholar] [CrossRef]
  198. Sheng, C.; Yu, F.; Wu, Y.; Peng, Z.; Chen, Y. Disproportionation of Sodium Superoxide in Metal-air Batteries. Angew. Chem. Int. Ed. Engl. 2018, 57, 9906–9910. [Google Scholar] [CrossRef]
  199. Ganapathy, S.; Adams, B.D.; Stenou, G.; Anastasaki, M.S.; Goubitz, K.; Miao, X.F.; Nazar, L.F.; Wagemaker, M. Nature of Li2O2 oxidation in a Li-O2 battery revealed by operando X-ray diffraction. J. Am. Chem. Soc. 2014, 136, 16335–16344. [Google Scholar] [CrossRef]
  200. Lin, F.; Liu, Y.; Yu, X.; Cheng, L.; Singer, A.; Shpyrko, O.G.; Xin, H.L.; Tamura, N.; Tian, C.; Weng, T.C.; et al. Synchrotron X-ray Analytical Techniques for Studying Materials Electrochemistry in Rechargeable Batteries. Chem. Rev. 2017, 117, 13123–13186. [Google Scholar] [CrossRef]
  201. Sun, C.W.; López, C.A.; Alonso, J.A. Elucidating the Diffusion Pathway of Protons in Ammonium Polyphosphate: A potential Electrolyte for Intermediate Temperature Fuel Cells. J. Mater. Chem. A 2017, 5, 7839–7844. [Google Scholar] [CrossRef] [Green Version]
  202. Sun, C.W.; Chen, L.L.; Shi, S.Q.; Reeb, B.; López, C.A.; Alonso, J.A.; Stimming, U. Visualization of the Diffusion Pathway of Protons in the Ammonium Polyphosphate for Intermediate Temperature Fuel Cells. Inorg. Chem. 2018, 57, 676–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Li, S.Q.; Meng, X.Y.; Alonso, J.A.; Fernández-Díaz, M.T.; Sun, C.W.; Wang, Z.L. Structural and Electrochemical Properties of LiMn0.6Fe0.4PO4 as a Cathode Material for Flexible Lithium-ion Batteries and Self-charging Power Pack. Nano Energy 2018, 52, 510–516. [Google Scholar] [CrossRef]
  204. Lu, Y.; López, C.A.; Wang, J.; Alonso, J.A.; Sun, C.W. Insight into the Structure and Functional Application of Mg doped Na0.5Bi0.5TiO3 electrolyte for Solid Oxide Fuel Cells. J. Alloys Compd. 2018, 752, 213–219. [Google Scholar] [CrossRef]
  205. Gong, Y.D.; Sun, C.W.; Huang, Q.A.; Alonso, J.A.; Fernández-Díaz, M.T.; Chen, L.Q. Dynamic octahedral breathing in oxygen-deficient Ba0.9Co0.7Fe0.2Nb0.1O3−δ perovskite performing as a cathode for intermediate-temperature solid oxide fuel cells. Inorg. Chem. 2016, 55, 3091–3097. [Google Scholar] [CrossRef]
  206. Yang, W.; Zhang, H.R.; Sun, C.W.; Liu, L.L.; Alonso, J.A.; Fernández-Díaz, M.T.; Chen, L.Q. Insight into the structure and functional application of Sr0.95Ce0.05CoO3−δ cathode for solid oxide fuel cells. Inorg. Chem. 2015, 54, 3477–3485. [Google Scholar] [CrossRef]
  207. Ma, Z.H.; Wang, Y.S.; Sun, C.W.; Alonso, J.A.; Fernández-Díaz, M.T.; Chen, L.Q. Experimental visualization of the diffusion pathway of sodium ions in the Na3[Ti2P2O10F] anode for sodium-ion battery. Sci. Rep. 2014, 4, 7231. [Google Scholar] [CrossRef] [Green Version]
  208. Wang, J.; Sun, C.W.; Gong, Y.D.; Zhang, H.R.; Alonso, J.A.; Fernández-Díaz, M.T.; Wang, Z.L.; Goodenough, J.B. Imaging of the diffusion pathway of Al3+ ion in NASICON-type (Al0.2Zr0.8)20/19Nb(PO4)3 electrolyte for high-temperature solid-state Al batteries. Chin. Phys. B 2018, 27, 128201. [Google Scholar] [CrossRef]
  209. Rong, X.; Liu, J.; Hu, E.; Liu, Y.; Wang, Y.; Wu, J.; Yu, X.; Page, K.; Hu, Y.-S.; Yang, W.; et al. Structure-Induced Reversible Anionic Redox Activity in Na Layered Oxide Cathode. Joule 2018, 2, 125–140. [Google Scholar] [CrossRef]
  210. Castellanos, M.M.; McAuley, A.; Curtis, J.E. Investigating Structure and Dynamics of Proteins in Amorphous Phases Using Neutron Scattering. Comput. Struct. Biotechnol. J. 2017, 15, 117–130. [Google Scholar] [CrossRef]
  211. Matthey, J. In the Lab: Synchrotron Radiation Based Solid Phase Characterisation of Industrial Catalysts and Materials. Johnson Matthey Technol. Rev. 2016, 60, 158–160. [Google Scholar]
  212. Wang, X.; Tan, S.; Yang, X.-Q.; Hu, E. Pair distribution function analysis: Fundamentals and application to battery materials. Chin. Phys. B 2020, 29, 028802. [Google Scholar] [CrossRef]
  213. Castillo-Blas, C.; Moreno, J.M.; Romero-Muniz, I.; Platero-Prats, A.E. Applications of pair distribution function analyses to the emerging field of non-ideal Metal-organic framework materials. Nanoscale 2020, 12, 15577–15587. [Google Scholar] [CrossRef] [PubMed]
  214. Yadav, N.; Folastre, N.; Bolomont, M.; Jamali, A.; Morcrette, M.; Davoisne, C. Understanding the Battery Degradation Mechanism in All-solid-state Batteries via In-situ SEM. Microsc. Microanal. 2021, 27, 105–106. [Google Scholar] [CrossRef]
  215. Han, S.; Cai, C.; Yang, F.; Zhu, Y.; Sun, Q.; Zhu, Y.G.; Li, H.; Wang, H.; Shao-Horn, Y.; Sun, X.; et al. Interrogation of the Reaction Mechanism in a Na-O2 Battery Using In Situ Transmission Electron Microscopy. ACS Nano 2020, 14, 3669–3677. [Google Scholar] [CrossRef]
  216. Zhang, L.; Tang, Y.; Liu, Q.; Yang, T.; Du, C.; Jia, P.; Wang, Z.; Tang, Y.; Li, Y.; Shen, T.; et al. Probing the charging and discharging behavior of K-CO2 nanobatteries in an aberration corrected environmental transmission electron microscope. Nano Energy 2018, 53, 544–549. [Google Scholar] [CrossRef]
  217. Liang, Z.; Zou, Q.; Wang, Y.; Lu, Y.C. Recent Progress in Applying In Situ/Operando Characterization Techniques to Probe the Solid/Liquid/Gas Interfaces of Li-O2 Batteries. Small Methods 2017, 1, 1700150. [Google Scholar] [CrossRef]
  218. Cortes, H.A.; Corti, H.R. In-situ characterization of discharge products of lithium-oxygen battery using Flow Electrochemical Atomic Force Microscopy. Ultramicroscopy 2021, 230, 113369. [Google Scholar] [CrossRef]
  219. Wang, Y.; Wang, W.; Xie, J.; Wang, C.H.; Yang, Y.W.; Lu, Y.-C. Electrochemical reduction of CO2 in ionic liquid: Mechanistic study of Li–CO2 batteries via in situ ambient pressure X-ray photoelectron spectroscopy. Nano Energy 2021, 83, 105830. [Google Scholar] [CrossRef]
  220. Wang, M.; Árnadóttir, L.; Xu, Z.J.; Feng, Z. In Situ X-ray Absorption Spectroscopy Studies of Nanoscale Electrocatalysts. Nano-Micro Lett. 2019, 11, 47. [Google Scholar] [CrossRef]
  221. Kimura, Y.; Fakkao, M.; Nakamura, T.; Okumura, T.; Ishiguro, N.; Sekizawa, O.; Nitta, K.; Uruga, T.; Tada, M.; Uchimoto, Y.; et al. Influence of Active Material Loading on Electrochemical Reactions in Composite Solid-State Battery Electrodes Revealed by Operando 3D CT-XANES Imaging. ACS Appl. Energy Mater. 2020, 3, 7782–7793. [Google Scholar] [CrossRef]
  222. Gao, R.; Zhou, D.; Ning, D.; Zhang, W.; Huang, L.; Sun, F.; Schuck, G.; Schumacher, G.; Hu, Z.; Liu, X. Probing the Self-Boosting Catalysis of LiCoO2 in Li-O2 Battery with Multiple In Situ/Operando Techniques. Adv. Funct. Mater. 2020, 30, 2002223. [Google Scholar] [CrossRef]
  223. Ding, S.; Yu, X.; Ma, Z.-F.; Yuan, X. A review of rechargeable aprotic lithium–oxygen batteries based on theoretical and computational investigations. J. Mater. Chem. A 2021, 9, 8160–8194. [Google Scholar] [CrossRef]
  224. Deng, Q.; Yang, Y.; Qu, S.; Wang, W.; Zhang, Y.; Ma, X.; Yan, W.; Zhang, Y. Electron structure and reaction pathway regulation on porous cobalt-doped CeO2/graphene aerogel: A free-standing cathode for flexible and advanced Li-CO2 batteries. Energy Storage Mater. 2021, 42, 484–492. [Google Scholar] [CrossRef]
  225. Jiang, H.R.; Tan, P.; Liu, M.; Zeng, Y.K.; Zhao, T.S. Unraveling the Positive Roles of Point Defects on Carbon Surfaces in Nonaqueous Lithium-Oxygen Batteries. J. Phys. Chem. C 2016, 120, 18394–18402. [Google Scholar] [CrossRef]
  226. Yuan, M.W.; Sun, Z.M.; Yang, H.; Wang, D.; Liu, Q.M.; Nan, C.Y.; Li, H.F.; Sun, G.B.; Chen, S.W. Self-Catalyzed Rechargeable Lithium-Air Battery by in-situ Metal Ion Doping of Discharge Products: A Combined Theoretical and Experimental Study. Energy Environ. Mater. 2021, 1–9. [Google Scholar] [CrossRef]
  227. Liu, Y.K.; Li, J.; Shen, Q.Y.; Zhang, J.; He, P.G.; Qu, X.H.; Liu, Y.C. Advanced characterizations and measurements for sodium-ion batteries with NASICON-type cathode materials. eScience 2022, 2, 10–31. [Google Scholar] [CrossRef]
  228. Jain, A.; Hautier, G.; Moore, C.J.; Ping Ong, S.; Fischer, C.C.; Mueller, T.; Persson, K.A.; Ceder, G. A high-throughput infrastructure for density functional theory calculations. Comp. Mater. Sci. 2011, 50, 2295–2310. [Google Scholar] [CrossRef]
  229. Wang, H.S.; Ji, Y.J.; Li, Y.Y. Simulation and design of energy materials accelerated by machine learning. WIREs Comput. Mol. Sci. 2019, 10, e1421. [Google Scholar] [CrossRef]
Figure 1. (a) Constant current discharge curves of Na-CO2/O2 battery based on IL electrolyte with mixed O2/CO2 supply. (b) Constant current discharge curves of Na-CO2/O2 battery with TEGDME-based electrolyte. (c) Relative capacity as a function of CO2 concentration. Reproduced with permission from [13]. Copyright 2013, Elsevier Ltd. (d) Schematic of Na-O2/CO2 reaction routes.
Figure 1. (a) Constant current discharge curves of Na-CO2/O2 battery based on IL electrolyte with mixed O2/CO2 supply. (b) Constant current discharge curves of Na-CO2/O2 battery with TEGDME-based electrolyte. (c) Relative capacity as a function of CO2 concentration. Reproduced with permission from [13]. Copyright 2013, Elsevier Ltd. (d) Schematic of Na-O2/CO2 reaction routes.
Batteries 09 00036 g001
Figure 2. Schematic of a Na-CO2 battery.
Figure 2. Schematic of a Na-CO2 battery.
Batteries 09 00036 g002
Figure 3. (a) SEM image of t-MWCNT cathode at room temperature, from top to side view. (b) TEM image. (c) HRTEM image. (d) In situ Raman spectra and corresponding discharge/charge product distributions at 11 selected points. (e) XPS of Ag wire cathode in different states. (f) EELS. (g) Complete discharge/charge curves at 1 A g−1. Reproduced with permission [6]. Copyright 2016, John Wiley and Sons. (h) IR spectra of pristine and activated MWCNTS. Reproduced with permission [29]. Copyright 2019, Royal Society of Chemistry.
Figure 3. (a) SEM image of t-MWCNT cathode at room temperature, from top to side view. (b) TEM image. (c) HRTEM image. (d) In situ Raman spectra and corresponding discharge/charge product distributions at 11 selected points. (e) XPS of Ag wire cathode in different states. (f) EELS. (g) Complete discharge/charge curves at 1 A g−1. Reproduced with permission [6]. Copyright 2016, John Wiley and Sons. (h) IR spectra of pristine and activated MWCNTS. Reproduced with permission [29]. Copyright 2019, Royal Society of Chemistry.
Batteries 09 00036 g003
Figure 6. Schematic representation of surface free energy and specific activity of supported catalysts as a function of particle size.
Figure 6. Schematic representation of surface free energy and specific activity of supported catalysts as a function of particle size.
Batteries 09 00036 g006
Figure 7. STEM-HAADF images of Pt@NCNT at low magnification (a) and high magnification (b). Red, circle in 7b shows the presence of single Pt atoms. (c) Schematic diagram of Na-CO2 nanobatteries constructed in ETEM. (d) Schematic diagram of the discharge/charge electrochemical processes in the Na-CO2 nanobattery. Reproduced with permission from [44]. Copyright 2020, Elsevier Ltd.
Figure 7. STEM-HAADF images of Pt@NCNT at low magnification (a) and high magnification (b). Red, circle in 7b shows the presence of single Pt atoms. (c) Schematic diagram of Na-CO2 nanobatteries constructed in ETEM. (d) Schematic diagram of the discharge/charge electrochemical processes in the Na-CO2 nanobattery. Reproduced with permission from [44]. Copyright 2020, Elsevier Ltd.
Batteries 09 00036 g007
Figure 13. Optimized structures (top), electron density profiles (middle) and free energy diagrams (bottom) of (a) graphite (0001) surface. (b) SV defects. (c) DV5-8-5 defects. (d) DV555-777 defects. (e) DV5555-6-7777 defects and (f) SW defects. The brown balls represent carbon atoms and the electron density is in |e| Bohr−3. Reproduced with permission from [225]. Copyright 2016, American Chemistry Society. (g) Free energy diagrams of co-doped CeO2 (110) surface CO2 to Li2CO3 were obtained based on DFT calculations. Asterisk * represents activated reactive intermediates. Reproduced with permission from [224]. Copyright 2021, Elsevier Ltd.
Figure 13. Optimized structures (top), electron density profiles (middle) and free energy diagrams (bottom) of (a) graphite (0001) surface. (b) SV defects. (c) DV5-8-5 defects. (d) DV555-777 defects. (e) DV5555-6-7777 defects and (f) SW defects. The brown balls represent carbon atoms and the electron density is in |e| Bohr−3. Reproduced with permission from [225]. Copyright 2016, American Chemistry Society. (g) Free energy diagrams of co-doped CeO2 (110) surface CO2 to Li2CO3 were obtained based on DFT calculations. Asterisk * represents activated reactive intermediates. Reproduced with permission from [224]. Copyright 2021, Elsevier Ltd.
Batteries 09 00036 g013
Figure 14. (a) Schematics of the cross−section of a molten Na battery and NASICON structures (bare sample and Na-excess sample); (b) measured total ionic conductivity for bare and Na excess sample. Reproduced with permission from [107]. Copyright 2016, American Chemical Society.
Figure 14. (a) Schematics of the cross−section of a molten Na battery and NASICON structures (bare sample and Na-excess sample); (b) measured total ionic conductivity for bare and Na excess sample. Reproduced with permission from [107]. Copyright 2016, American Chemical Society.
Batteries 09 00036 g014
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, Z.; Sun, C.; Lu, L.; Jiao, L. Recent Progress and Perspectives of Solid State Na-CO2 Batteries. Batteries 2023, 9, 36. https://doi.org/10.3390/batteries9010036

AMA Style

Wang Z, Sun C, Lu L, Jiao L. Recent Progress and Perspectives of Solid State Na-CO2 Batteries. Batteries. 2023; 9(1):36. https://doi.org/10.3390/batteries9010036

Chicago/Turabian Style

Wang, Zelin, Chunwen Sun, Liang Lu, and Lifang Jiao. 2023. "Recent Progress and Perspectives of Solid State Na-CO2 Batteries" Batteries 9, no. 1: 36. https://doi.org/10.3390/batteries9010036

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop