Next Article in Journal
Determination of Liberation Degree of Mechanically Processed Waste Printed Circuit Boards by Using the Digital Microscope and SEM-EDS Analysis
Next Article in Special Issue
Technical and Economic Optimal Solutions for Utility-Scale Solar Photovoltaic Parks
Previous Article in Journal
Modified Technique of Parameter Identification of a Permanent Magnet Synchronous Motor with PWM Inverter in the Presence of Dead-Time Effect and Measurement Noise
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Sodium Batteries: A Review on Sodium-Sulfur and Sodium-Air Batteries

1
Environmental Health and Safety Department, Carnegie Mellon University, Pittsburgh, PA 15213, USA
2
Mechanical and Materials Engineering Department, Florida International University, Miami, FL 33174, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Electronics 2019, 8(10), 1201; https://doi.org/10.3390/electronics8101201
Submission received: 28 September 2019 / Revised: 16 October 2019 / Accepted: 21 October 2019 / Published: 22 October 2019
(This article belongs to the Special Issue Advanced Renewable Energy Technologies)

Abstract

:
Lithium-ion batteries are currently used for various applications since they are lightweight, stable, and flexible. With the increased demand for portable electronics and electric vehicles, it has become necessary to develop newer, smaller, and lighter batteries with increased cycle life, high energy density, and overall better battery performance. Since the sources of lithium are limited and also because of the high cost of the metal, it is necessary to find alternatives. Sodium batteries have shown great potential, and hence several researchers are working on improving the battery performance of the various sodium batteries. This paper is a brief review of the current research in sodium-sulfur and sodium-air batteries.

1. Introduction

Energy storage has become one of the major global concerns as the world is preparing to find an alternative to fossil fuels for storage and energy generation. Lithium-ion based batteries are already established as an alternative energy source to fossil fuels. They are the most common type of rechargeable batteries that can be found in almost all the portable devices to electric vehicles (EVs) [1,2,3,4]. Lithium metal is the lightest metal with a small ionic radius and has a very low redox potential of −3.04 V against hydrogen electrode. The above properties enable batteries with higher voltage and energy densities. Furthermore, the small ionic radius helps ions to diffuse in solids. Higher capacity, longer cycle life, and higher rate capability make lithium-ion batteries a perfect fit for portable electronics and electric vehicles.
The demand for lithium-ion batteries is increasing rapidly because of the increased use of portable electronics and the growing future use of EVs. Several developed countries are going to ban fossil fuel-driven vehicles, so the demand for lithium metal will increase rapidly, and the metal is not abundant in nature, so there is a possibility of the prices to go up sharply in the next few decades [5,6]. Hence research has been going on to find an alternative to lithium for future generation energy storage devices. Sodium is an alternate choice for lithium as the metal is abundant in nature, which will make batteries cheaper, is the second lightest material next to lithium, and its redox potential is −2.7 V against hydrogen electrode, which is slightly higher than lithium.
The application of RT Na/S battery, however, faces some critical challenges of low cyclic stability and rapid capacity fading in the first few cycles, which is similar to Li-S battery systems because of the polysulfide shuttle effect during the charge-discharge process. Unfortunately, the problem in Na-S battery is more serious than the Li-S system. In this review article, we discuss the recent development beyond sodium-ion batteries, focusing on room temperature sodium-sulfur (RT Na-S) and sodium-air/O2 batteries. The article first introduces the principles of charge/discharge mechanisms of RT Na-S and Na-air/O2 batteries, followed by a summary of the recent progress in research.

2. Sodium Sulfur Battery

Sulfur as cathode materials possesses a high discharge capacity of 1675 mAh g−1 which is one order of magnitudes compared to the insertion-cathode system. This high capacity makes this material a serious candidate for the future generation battery system. In 2006, room temperate sodium-sulfur (Na-S) battery was reported for the first time [7]. Since then several researches have been conducted to improve the electrochemical utilization of the active sulfur material and its cyclic performance for longer cycles [8]. The principle mechanism and the selection of Na-S battery electrodes are similar to Li-S battery technology. The most common Na-S cell consists of Na metal as anode, sulfur-carbon composite as cathode, and organic liquids as electrolyte. During the discharge process the anode sodium metal gets oxidized and forms sodium ions and electrons. The ion moves through the electrolyte and separator to reach the sulfur composite cathode, while the electrons move to the cathode through the external circuit and generate electricity. Meanwhile, on the cathode, sulfur gets reduced by accepting the sodium ions, and the electrons thus form complex sodium polysulfides. The typical charge-discharge cut-off voltage for Na-S cell is from 1 to 3 V. The charge-discharge of Na-S cell shows a series of complicated formation of long-chain (Na2Sn, 4 ≤ n ≤ 8) and short-chain (Na2Sn, 1 ≤ n < 4) polysulfides similar to Li-S cells [9,10]. Figure 1b shows a typical discharge voltage profile of a Na-S battery. The figure shows two plateaus at ~2.2 V and ~1.65 V during discharge, and at charging the plateaus are at ~1.75 V and ~2.4 V. The discharge curve in the figure is divided into four stages. On the first stage, elemental sulfur gets reduced to long-chain sodium polysulfides, where a solid to liquid transition occurs at ≈2.2 V:
S 8 + 2 Na + + 2 e Na 2 S 8
Stage 2 represents the sloping line in the voltage range from 2.2 V to 1.65 V, where a liquid to liquid transition from Na2S8 to Na2S4 occurs:
Na 2 S 8 + 2 Na + + 2 e 2 Na 2 S 4
In stage 3, a low-voltage plateau at ≈1.65 V can be seen because of a liquid to solid transition where Na2S4 is reduced to short-chain insoluble Na2S3, Na2S2, or Na2S:
Na 2 S 4 + 2 3 Na + + 2 3 e 4 3 Na 2 S 3
Na 2 S 4 + 2 Na + + 2 e 2 Na 2 S 2
Na 2 S 4 + 6 Na + + 6 e 2 Na 2 S
In the last stage, a second sloping line in the voltage range from 1.65 V to 1.2 V corresponding to a solid-solid reaction by the reduction reaction of Na2S2 to Na2S:
Na 2 S 2 + 2 Na + + 2 e 2 Na 2 S
Stage 2 is the most complicated reaction step compared to the other stages as it is dominated by the presence of multiple polysulfide species in the solution. In stage 3, the reaction competition between the reactions 3 to 5 determines the discharge voltage and the capacity of the stage. Stage 4 is kinetically the slowest among the others and shows high polarization because of the non-conductive nature of the short-chain Na2S2 and Na2S formation.

2.1. Technical Challenges

Although sulfur cathode-based batteries possess high theoretical capacity, there are some technical challenges for Na-S batteries coming from the materials chemistry and the system of the battery [12,13]. The common problems are the same as of Li-S batteries. The extremely low conductivity of sulfur (≈10−30 S cm−1), the low reactivity between sulfur and sodium, the formation of intermediate polysulfide during the discharge process makes unstable electrochemical contact, and the gradual increase in impedances because of the formation of passivation layers on the electrodes during cycling makes it difficult to reach high theoretical capacity. Although RT Na-S batteries have similar problems like Li-S batteries, the concerns are greater for Na-S batteries because of the sluggish nature of the electrolyte and rapid polysulfide migration during cycling, which makes the application of RT Na-S batteries more challenging.
So far, several types of research have been going on to overcome the challenges for RT Na-S batteries. Most of the researches are focusing on the sulfur-containing cathode to improve the reaction kinetics of the cathode; lower the shuttle effect, and improve the performance of the RT Na-S batteries at room temperature. Other researches also focused on the application of solid-state electrolyte, which improved the interfacial contact between electrode-electrolyte interfaces, thus reduced the Na dendrite formation. The following sections will highlight the recent progress on the electrode and electrolyte materials that are used to improve the performance of RT Na-S batteries.

2.2. Cathodes

As it is discussed earlier that the two major drawbacks in RT-Na-S battery system are the insulating nature of the active material sulfur and the shuttle effect because of the formation of polysulfides during cycling, which gives poor cyclic stability of the battery system. A large number of researches are going on to improve the conductivity of the cathode system by mixing or embedding or encapsulating sulfur in a conducting agent like carbon or polymers. The list of the research in RT Na-S battery is shown in Table 1, which summarizes the cathode, electrolyte materials that are used in the battery, and the cyclic performances at the first cycle and at the longest cycles at a certain current density.
At the early stages of research, sulfur was mixed with carbon in a ball mill for a proper mixing of sulfur and carbon, which created a conductive framework to enhance the utilization of active sulfur. Acetylene carbon black and super P were mostly used as a source of conductive agents. Park et al., for the first time reported this improvement by mixing 70 wt% S with 20 wt% C in 10 wt% polyethylene oxide (PEO) binder and a solid electrolyte with an initial discharge capacity of 505 mAh g−1 at 90 °C. However, a rapid capacity drop was observed in just after 10 cycles to only 166 mAh g−1 [14]. An increase in carbon content to 40 wt% was reported by Wenzel et al. with an initial discharge capacity of ~475 mAhg−1 and enhanced cyclic stability to 40 cycles at room temperature using beta alumina solid electrolyte [15]. Although the mechanical mixing of sulfur, carbon, and binder is an easier approach to resolve the conductivity issue of the electrode system, more research was needed to improve the cyclic stability. Low sulfur content creates poor contact between S and C, while high sulfur content creates sulfur agglomerates on the carbon particles and results in poor electrochemical performances.
To overcome the above-mentioned problems, different approaches were made to prepare sulfur-based composite cathode which improved the contact between the sulfur and carbon and also the prevention of agglomeration. Wang et al. for the first time reported sulfurized composite cathode with conductive polyacrylonitrile (PAN) and achieved 654 mAhg−1 at the initial cycles while maintaining stability to 500 mAh.g−1 after 18 cycles [12]. In recent times the application of sulfurized RT Na-S batteries has been reported to have high initial discharge capacity and long cycle life [15,16,17,18,19], which is due to the stable atomic configuration of sulfur in the PAN matrix which helped to reduce polysulfide formation and improved kinetics of the reaction. Wu et al. recently reported improved results using PAN/S composite cathode, as shown in Figure 2.
To increase the active sulfur loading, sulfur has been introduced in various forms of carbonaceous matrix framework. The most frequently used forms of carbon are CNTs, graphene, microporous, mesoporous, macroporous, and hollow carbon spheres. These types of carbons help to increase the active sulfur loading in the carbon matrix. This process also improves intimate contact between the sulfur and carbon which improves sodium ion transfer and electrolyte transport between the electrodes and thus improves the capacity and cycle life. Recently, Ghosh et al. reported 631 mAh.g−1 initial specific capacity using a three-dimensional ternary composite porous cathode, which is made of reduced graphene oxide (rGO), sulfur, and MnxOy incorporated within a Na alginate/polyaniline hybrid matrix as shown in Figure 3. They have concluded that the addition of MnxOy forms complex polythionates and acts as a mediator to anchor higher-order polysulfides and prevents polysulfide dissolution [22].

2.3. Electrolytes

Electrolyte soaked with a separator helps to move ions to transfer between the electrodes and prevents the short circuit by creating a barrier between the electrodes to come in contact with each other. So far various salts containing sodium and solvents based on carbonates, ethers, and ionic liquid-based have been reported in RT-Na/S battery. It has been found that the electrolyte also plays an important part in the performance of the battery alongside the cathodes. Carbonate solvents with NaClO4 salts are mostly used in RT-Na/S batteries [11,12,19,20]. However, the problem with this type of electrolyte is that it cannot protect the Na metal from forming dendrites, so at longer cycles and high C-rates these dendrites grow and short circuit the battery. To protect the Na metal, Zhang et al. added fluoroethylene carbonate (FEC) in the carbonate electrolyte which forms an SEI layer on the metal and improved the cyclic life of the battery. Xu et al. reported a high initial discharge capacity of 1635 mAh g−1 at 0.1 C rate at room temperature using multiporous carbon fibers (MPCFs), super-P conductive agents, and sodium carboxymethyl cellulose (CMCNa) binder [11]. They have also reported high cyclic stability of 1134 mAh g−1 after 50 cycles, which they attributed as the addition of InI3 additive in the carbonate electrolyte. According to their findings, InI3 acted as a redox mediator where the Na metal was protected by the formation of an indium (In) layer. This layer is formed by the In3+ redox reaction and a fluorine (F)-rich stable SEI film. It also helped to oxidize the irreversible Na2S during the charging process which improved the Coulombic efficiency.
Various ether-based electrolytes have been investigated in RT-Na/S batteries to overcome the shuttling effect. Long-chain ethers such as glyme-based tetraethylene glycol dimethyl ether (TEGDME) showed much-improved performance in this type of battery system because of its better solvation of oxygen atoms in the glyme structure which gives higher dissociation of sodium salts and polysulfide compounds. It has also been reported that the addition of 0.1M NaNO3 in the TEGDME and NaClO4 electrolyte forms a protective SEI layer on the Na metal and improves the cyclic stability of the battery [28]. In a recent study, it has been reported that 0.1M NaNO3 is not sufficient to protect the Na metal completely, while the addition of 0.2M NaNO3 shows complete protection of the metal and shows improved performance [22].
Room-temperature ionic liquids (RTIL) are salts in a molten condition where the cations and anions are bonded as a weak van der Waals force. The major advantage of RTILs is its low flammability, negligible vapor pressure, high thermal stability, moderate ionic conductivity, and wide electrochemical potential window [29,30]. The above-mentioned advantages make it a serious candidate in lithium-ion and sulfur battery system. It has been widely used as an electrolyte as a replacement for the conventional carbonate-based electrolyte in lithium battery applications [31,32,33,34]. RTILs have also been reported in Li-S battery application, however the capacity and cycle life was poor because of the viscous nature of the electrolyte [29,35,36,37,38]. The application of RTIL in RT-Na/S battery is very rare. So far IL added as an additive in carbonate-based electrolyte, and improved capacity and cycle life have been observed [27].
Solid polymer or gel polymer electrolyte(GPE) is another recent approach which has shown to have good interfacial contact with the electrodes and prevents dendrite growth in the metal. Recently, Yu et al. reported Nasicon (Na3Zr2Si2PO12)-based solid electrolyte which acted as a polysulfide shield by enhancing the interfacial properties, as shown in Figure 4 [17].

3. Sodium-Air Battery

3.1. Introduction

Metal-air batteries have been under investigation for many years because of their higher theoretical capacities. Most of the research has been focused on lithium-air batteries because they have the highest energy density of about 3458 Wh.kg−1 and an overpotential of 2.96 V [31,39,40,41,42,43,44,45,46,47]. However, because of limitations including poor cyclability, large overpotential, SEI (solid electrolyte interface) layer formation [39], blockage of cathode reaction sites by electrolyte decomposition [42,44], electrolyte degradation [39], and high lithium metal cost, several other battery chemistries are being studied simultaneously. Recently, research is being conducted on sodium air, aluminum air, magnesium air, zinc-air, iron air, and tin air batteries.
Sodium air batteries have reported promising results as an alternative metal-air battery and have gained attention as the next generation batteries because of their high specific densities and capacities. Sodium is the sixth most abundant element found on earth with an energy density of about 1600 Wh.kg−1 [48]. The theoretical energy density and specific capacity of sodium are 3164 Wh.kg−1 and 1166 mAh.g−1, respectively. The energy cost of rechargeable Li-air batteries is $300–$500 kWh−1 whereas for Na-air batteries, it is $100–150 kWh−1 [48]. Even though lithium and sodium are chemically close, they react very differently with oxygen. In lithium-air batteries, lithium reacts with oxygen to form LiO2 (lithium oxide) as an intermediate species before forming Li2O2 (lithium peroxide). Li2O2 deposition on the cathodes results in lower reversal of discharge products during charging. Whereas for Na-air batteries, sodium combines with oxygen to form NaO2 (sodium superoxide) which is more stable, does not decompose, and hence assists during the reversal of discharge products while charging.
Peled et al. first investigated Na-air batteries in 2011 [49]. They operated the Na-air battery at 105 °C using liquid sodium, ETEK cathode, and polymer electrolyte. First room temperature Na-air battery was reported by Sun et al. in 2012, with 20 cycles, the discharge capacity of 1058 mAh.g−1, and a coulombic efficiency of 85% [50]. They confirmed the formation of the major discharge product, Na2O2.
A typical Na-air battery as shown in Figure 5, consists of a sodium metal anode and a porous cathode separated by a separator in the electrolyte. The batteries undergo discharge and charge processes. During discharging, sodium gets oxidized and forms sodium ions (Na+). The cation then moves through the electrolyte to combine with the electron from the external circuit and reduces the adsorbed oxygen, thus forming NaO2 that is reversibly reduced to O2 and Na+. Na-air batteries can be classified as aprotic and aqueous/hybrid Na-air batteries [51,52]. In aprotic batteries, the insoluble and non-conductive NaO2 and Na2O2 (discharge products) are deposited on the porous cathode thus blocking the reaction sites. Whereas in hybrid or aqueous Na-air batteries, a Na-ion conducting NASICON-type (Na3Zr2Si2PO12) separator has been used that assists in reducing the clogging of cathode pores with the discharge products [51,53,54]. Figure 6 shows the structure of non-aqueous and aqueous Na-air batteries with nanoporous gold electrode and their discharge-charge behavior.

3.2. Electrochemical Reactions in the Na-Air Battery

The following reactions occur during the discharge–charge process of aqueous Na-air batteries [55]:
At cathode:O2 + 2H2O + 4e ↔ 4 OH;E0 = +0.40 V (i)
At anode:Na ↔ Na+ + e;E0 = +2.71 V (ii)
Overall:4 Na + O2 + 2 H2O ↔ 4 NaOH;E0 = 3.11 V (iii)
The standard cell voltage based on the above equations is 3.11 V.
However, for aprotic Na-air batteries, the reactions are given as follows [56]:
At Cathode:O2 + e ↔ O2;E0 = −0.44 V (iv)
OrO2 + 2e ↔ O22−;E0 = 0.38 V (v)
Anode:Na ↔ Na+ + e;E0 = +2.71 V (vi)
Overall:Na + O2 ↔ NaO2;E0 = 2.27 V (vii)
Or2Na + O2 ↔ Na2O2;E0 = 2.33 V (viii)

3.3. Components of the Na-Air Battery

3.3.1. Electrolytes

To enhance Na-air battery performance, the electrolyte must be highly stable against oxygen for longer cycles and reversibility. Na-air batteries have been classified as aqueous and non-aqueous based on the electrolytes. Aqueous electrolytes assist in the prevention of deactivation of catalyst by discharge products. Anode electrolytes are commonly formed with sodium salts, namely NaCF3SO3 and NaClO4, in 0.1 M–1 M of organic solvents such as ethylene carbonate/dimethyl carbonate (EC/DMC) and tetra-ethylene glycol dimethyl ether (TEGDME) [58]. The electrolyte used significantly affects the formation of SEI film thus reducing the battery performance if the SEI film is unstable or improving the battery performance by the stable SEI film formation [59]. The influence of various sodium salts including TFSi, PF6, OTf, and ClO4 on Na-air batteries was studied by Lutz et al. [59]. Figure 7 shows the effect of the different salts on the SEI formation.

3.3.2. Separator

Separators are inactive components used in aqueous Na-air batteries to separate the anode and cathode to prevent short circuits, to provide a path for ionic conduction, and prevent mixing of electrolytes. NASICON, sodium-ion conducting oxide ceramic with the chemical formula Na1+xZr2SixP3−xO12, 0 < x < 3 [60] is the most commonly used separator for Na-air batteries.

3.3.3. Cathodes and Catalysts

The cathodes play a vital role in the electrochemical performance of the batteries. Porous carbon is a typical cathode for most of the Na-air batteries. However, the discharge products get deposited in the porous structure of the cathode, thus blocking the reaction sites which results in battery failure. Catalysts optimize battery performance by providing better-dispersed reaction sites. The comparison and performance of Na-air batteries using various cathodes, catalysts, separators, and electrolytes are shown in Table 2. Lithium metal anode was used for all the references mentioned in the table below.

4. Conclusions

Although sodium-ion batteries are promising candidates for energy storage purposes, they still suffer from limited cyclability and are under development. The main purpose of this article was to summarize the work done in RT Na-S and Na-air/O2 batteries to implement the technology in application. However, technology is still in its infancy and much more research is needed to overcome the major hurdles like poor capacity retention and cycle life because of the poor conductivity and polysulfide formation of sodium during cycling. Newer sodium ion-conducting separator may be considered to improve the battery performance further. Modification or synthesis of novel catalysts should also be considered to suppress the sodium dendrite formation that eventually results in battery death. By addressing these issues, better batteries can be developed for use in automobiles and other electronic devices.

Author Contributions

N.C. and M.S. have contributed equally to the review paper.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goodenough, J.B. Evolution of Strategies for Modern Rechargeable Batteries. Acc. Chem. Res. 2012, 46, 1053–1061. [Google Scholar] [CrossRef] [PubMed]
  2. Etacheri, V.; Marom, R.; Elazari, R.; Salitra, G.; Aurbach, D. Challenges in the Development of Advanced Li-Ion Batteries: A Review. Energy Environ. Sci. 2011, 4, 3243–3262. [Google Scholar] [CrossRef]
  3. Scrosati, B.; Garche, J. Lithium Batteries: Status, Prospects and Future. J. Power Sources 2010, 195, 2419–2430. [Google Scholar] [CrossRef]
  4. Manthiram, A. Materials Challenges and Opportunities of Lithium Ion Batteries. J. Phys. Chem. Lett. 2011, 2, 176–184. [Google Scholar] [CrossRef]
  5. Albertus, P.; Babinec, S.; Litzelman, S.; Newman, A. Status and Challenges in Enabling the Lithium Metal Electrode for High-Energy and Low-Cost Rechargeable Batteries. Nat. Energy 2017, 3, 16–21. [Google Scholar] [CrossRef]
  6. McCloskey, B.D. Attainable Gravimetric and Volumetric Energy Density of Li–S and Li Ion Battery Cells with Solid Separator-Protected Li Metal Anodes. J. Phys. Chem. Lett. 2015, 6, 4581–4588. [Google Scholar] [CrossRef]
  7. Park, C.W.; Ahn, J.H.; Ryu, H.S.; Kim, K.W.; Ahn, H.J. Room-Temperature Solid-State Sodium/Sulfur Battery. Electrochem. Solid State Lett. 2006, 9, A123–A125. [Google Scholar] [CrossRef]
  8. Yu, X.; Manthiram, A. Capacity Enhancement and Discharge Mechanisms of Room-Temperature Sodium-Sulfur Batteries. ChemElectroChem 2014, 1, 1275–1280. [Google Scholar] [CrossRef]
  9. Zhang, B.; Lai, W.; Xu, Y.; Wang, Y.X.; Chou, S.L.; Liu, H.K.; Dou, S.X.; Wang, Y.; Chou, S.; Liu, H.; et al. Room-Temperature Sodium-Sulfur Batteries: A Comprehensive Review on Research Progress and Cell Chemistry. Adv. Energy Mater. 2017, 7, 1602829. [Google Scholar]
  10. Manthiram, A.; Yu, X. Ambient Temperature Sodium-Sulfur Batteries. Small 2015, 11, 2108–2114. [Google Scholar] [CrossRef]
  11. Xu, X.; Zhou, D.; Qin, X.; Lin, K.; Kang, F.; Li, B.; Shanmukaraj, D.; Rojo, T.; Armand, M.; Wang, G. A Room-Temperature Sodium–Sulfur Battery with High Capacity and Stable Cycling Performance. Nat. Commun. 2018, 9, 3870. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, J.; Yang, J.; Nuli, Y.; Holze, R. Room Temperature Na/S Batteries with Sulfur Composite Cathode Materials. Electrochem. Commun. 2007, 9, 31–34. [Google Scholar] [CrossRef]
  13. Kim, J.S.; Ahn, H.J.; Kim, I.P.; Kim, K.W.; Ahn, J.H.; Park, C.W.; Ryu, H.S. The Short-Term Cycling Properties of Na/PVdF/S Battery at Ambient Temperature. J. Solid State Electrochem. 2008, 12, 861–865. [Google Scholar] [CrossRef]
  14. Park, C.W.; Ryu, H.S.; Kim, K.W.; Ahn, J.H.; Lee, J.Y.; Ahn, H.J. Discharge Properties of All-Solid Sodium–Sulfur Battery Using Poly (Ethylene Oxide) Electrolyte. J. Power Sources 2007, 165, 450–454. [Google Scholar] [CrossRef]
  15. Wenzel, S.; Metelmann, H.; Raiis, C.; Durr, A.K.; Janek, J.; Adelhelm, P. Thermodynamics and Cell Chemistry of Room Temperature Sodium/Sulfur Cells with Liquid and Liquid/Solid Electrolyte. J. Power Sources 2013, 243, 758–765. [Google Scholar] [CrossRef]
  16. Kim, I.; Kim, C.H.; Choi, S.H.; Ahn, J.P.; Ahn, J.H.; Kim, K.W.; Cairns, E.J.; Ahn, H.J. A Singular Flexible Cathode for Room Temperature Sodium/Sulfur Battery. J. Power Sources 2016, 307, 31–37. [Google Scholar] [CrossRef]
  17. Yu, X.; Manthiram, A. Sodium-Sulfur Batteries with a Polymer-Coated NASICON-Type Sodium-Ion Solid Electrolyte. Matter 2019, 1, 439–451. [Google Scholar] [CrossRef]
  18. Wu, J.; Liu, J.; Lu, Z.; Lin, K.; Lyu, Y.Q.; Li, B.; Ciucci, F.; Kim, J.K. Non-Flammable Electrolyte for Dendrite-Free Sodium-Sulfur Battery. Energy Storage Mater. 2019. [Google Scholar] [CrossRef]
  19. Hwang, T.H.; Jung, D.S.; Kim, J.S.; Kim, B.G.; Choi, J.W. One-Dimensional Carbon–Sulfur Composite Fibers for Na–S Rechargeable Batteries Operating at Room Temperature. Nano Lett. 2013, 13, 4532–4538. [Google Scholar] [CrossRef]
  20. Zhang, B.W.; Sheng, T.; Liu, Y.D.; Wang, Y.X.; Zhang, L.; Lai, W.H.; Wang, L.; Yang, J.; Gu, Q.F.; Chou, S.L.; et al. Atomic Cobalt as an Efficient Electrocatalyst in Sulfur Cathodes for Superior Room-Temperature Sodium-Sulfur Batteries. Nat. Commun. 2018, 9, 4082. [Google Scholar] [CrossRef]
  21. Lu, Q.; Wang, X.; Cao, J.; Chen, C.; Chen, K.; Zhao, Z.; Niu, Z.; Chen, J. Freestanding Carbon Fiber Cloth/Sulfur Composites for Flexible Room-Temperature Sodium-Sulfur Batteries. Energy Storage Mater. 2017, 8, 77–84. [Google Scholar] [CrossRef]
  22. Ghosh, A.; Kumar, A.; Roy, A.; Panda, M.R.; Kar, M.; Macfarlane, D.R.; Mitra, S. Three-Dimensionally Reinforced Freestanding Cathode for High-Energy Room-Temperature Sodium–Sulfur Batteries. ACS Appl. Mater. Interfaces 2019, 11, 14101–14109. [Google Scholar] [CrossRef] [PubMed]
  23. Xin, S.; Gu, L.; Zhao, N.H.; Yin, Y.X.; Zhou, L.J.; Guo, Y.G.; Wan, L.J. Smaller Sulfur Molecules Promise Better Lithium–Sulfur Batteries. J. Am. Chem. Soc. 2012, 134, 18510–18513. [Google Scholar] [CrossRef] [PubMed]
  24. Lee, D.J.; Park, J.W.; Hasa, I.; Sun, Y.K.; Scrosati, B.; Hassoun, J. Alternative Materials for Sodium Ion–Sulphur Batteries. J. Mater. Chem. A 2013, 1, 5256. [Google Scholar] [CrossRef]
  25. Wang, Y.X.; Yang, J.; Lai, W.; Chou, S.L.; Gu, Q.F.; Liu, H.K.; Zhao, D.; Dou, S.X. Achieving High-Performance Room-Temperature Sodium–Sulfur Batteries With S@Interconnected Mesoporous Carbon Hollow Nanospheres. J. Am. Chem. Soc. 2016, 138, 16576–16579. [Google Scholar] [CrossRef]
  26. Zheng, S.; Wen, Y.; Zhu, Y.; Han, Z.; Wang, J.; Yang, J.; Wang, C. In Situ Sulfur Reduction and Intercalation of Graphite Oxides for Li-S Battery Cathodes. Adv. Energy Mater. 2014, 4, 1400482. [Google Scholar] [CrossRef]
  27. Wei, S.; Xu, S.; Agrawral, A.; Choudhury, S.; Lu, Y.; Tu, Z.; Ma, L.; Archer, L.A. A Stable Room-Temperature Sodium–Sulfur Battery. Nat. Commun. 2016, 7, 11722. [Google Scholar] [CrossRef]
  28. Zhang, S.S.; Zhang, S. Role of LiNO3 in Rechargeable Lithium/Sulfur Battery. Electrochim. Acta 2012, 70, 344–348. [Google Scholar] [CrossRef]
  29. Safa, M.; Hao, Y.; Chamaani, A.; Adelowo, E.; Chawla, N.; Wang, C.; El Zahab, B. Capacity Fading Mechanism in Lithium-Sulfur Battery Using Poly(Ionic Liquid) Gel Electrolyte. Electrochim. Acta 2017, 258, 1284–1292. [Google Scholar] [CrossRef]
  30. Safa, M.; Adelowo, E.; Chamaani, A.; Chawla, N.; Baboukani, A.R.; Herndon, M.; Wang, C.; El Zahab, B. Poly(Ionic Liquid)-Based Composite Gel Electrolyte for Lithium Batteries. ChemElectroChem 2019, 6, 3319–3326. [Google Scholar] [CrossRef]
  31. Safa, M.; Chamaani, A.; Chawla, N.; El Zahab, B. Polymeric Ionic Liquid Gel Electrolyte for Room Temperature Lithium Battery Applications. Electrochim. Acta 2016, 213, 587–593. [Google Scholar] [CrossRef]
  32. Xu, K. Electrolytes and Interphases in Li-Ion Batteries and Beyond. Chem. Rev. 2014, 114, 11503–11618. [Google Scholar] [CrossRef] [PubMed]
  33. Kalhoff, J.; Eshetu, G.G.; Bresser, D.; Passerini, S. Safer Electrolytes for Lithium-Ion Batteries: State of the Art and Perspectives. ChemSusChem 2015, 8, 2154–2175. [Google Scholar] [CrossRef] [PubMed]
  34. Seki, S.; Watanabe, M. Polymer and Ionic Liquid Electrolytes for Advanced Lithium Batteries. In Nanoscale Technology for Advanced Lithium Batteries; Springer: Berlin, Germany, 2014; pp. 51–61. [Google Scholar]
  35. Watanabe, M.; Thomas, M.L.; Zhang, S.; Ueno, K.; Yasuda, T.; Dokko, K. Application of Ionic Liquids to Energy Storage and Conversion Materials and Devices. Chem. Rev. 2017, 117, 7190–7239. [Google Scholar] [CrossRef] [Green Version]
  36. Barghamadi, M.; Best, A.S.; Bhatt, A.I.; Hollenkamp, A.F.; Musameh, M.; Rees, R.J.; Ruther, T. Lithium–Sulfur Batteries—The Solution is in the Electrolyte, but is the Electrolyte a Solution? Energy Environ. Sci. 2014, 7, 3902–3920. [Google Scholar] [CrossRef]
  37. Li, G.; Li, Z.; Zhang, B.; Lin, Z. Developments of Electrolyte Systems for Lithium–Sulfur Batteries: A Review. Energy Storage 2015, 3, 5. [Google Scholar]
  38. Yuan, L.; Feng, J.; Ai, X.; Cao, Y.; Chen, S.; Yang, H. Improved Dischargeability and Reversibility of Sulfur Cathode in a Novel Ionic Liquid Electrolyte. Electrochem. Commun. 2006, 8, 610–614. [Google Scholar] [CrossRef]
  39. Chamaani, A.; Safa, M.; Chawla, N.; El Zahab, B. Composite Gel Polymer Electrolyte for Improved Cyclability in Lithium–Oxygen Batteries. ACS Appl. Mater. Interfaces 2017, 9, 33819–33826. [Google Scholar] [CrossRef]
  40. Chawla, N.; Bharti, N.; Singh, S. Recent advances in non-flammable electrolytes for safer lithium-ion batteries. Batteries 2019, 5, 19. [Google Scholar]
  41. Chawla, N. Nanocatalysts in Lithium Oxygen Batteries. Nano Sci. Nano Technol. An Indian J. 2018, 12, 128. [Google Scholar]
  42. Chawla, N.; Chamaani, A.; Safa, M.; El Zahab, B. Palladium-Filled Carbon Nanotubes Cathode for Improved Electrolyte Stability and Cyclability Performance of Li-O2 Batteries. J. Electrochem. Soc. 2017, 164, A6303–A6307. [Google Scholar] [CrossRef]
  43. Chawla, N.; Chamaani, A.; Safa, M.; Herndon, M.; El Zahab, B. Mechanism of Ionic Impedance Growth for Palladium-Containing CNT Electrodes in Lithium-Oxygen Battery Electrodes and its Contribution to Battery Failure. Batteries 2019, 5, 15. [Google Scholar] [CrossRef]
  44. Viswanathan, V.; GirishKumar, G.; McCloskey, B.D.; Hummelshoj, J.S.; Norskov, J.K.; Thygesen, K.S.; Luntz, A.C. Electrical Conductivity in Li2O2 and its Role in Determining Capacity Limitations in Non-Aqueous Li-O2 Batteries. J. Chem. Phys. 2011, 135, 214704. [Google Scholar] [CrossRef] [PubMed]
  45. GirishKumar, G.; McCloskey, B.; Luntz, A.C.; Swanson, S.; Wilcke, W. Lithium−Air Battery: Promise and Challenges. J. Phys. Chem. Lett. 2010, 1, 2193–2203. [Google Scholar] [CrossRef]
  46. Ogasawara, T.; Debart, A.; Holzapfel, M.; Novak, P.; Bruce, P.G. Rechargeable Li2O2 Electrode for Lithium Batteries. J. Am. Chem. Soc. 2006, 37, 1390–1393. [Google Scholar] [CrossRef]
  47. Chawla, N. Recent Advances in Air-Battery Chemistries. Mater. Today Chem. 2019, 12, 324–331. [Google Scholar] [CrossRef]
  48. Yin, W.W.; Fu, Z.W. The Potential of Na–Air Batteries. ChemCatChem 2017, 9, 1545–1553. [Google Scholar] [CrossRef]
  49. Peled, E.; Golodnitsky, D.; Mazor, H.; Goor, M.; Avshalomov, S. Parameter Analysis of a Practical Lithium- and Sodium-Air Electric Vehicle Battery. J. Power Sources 2011, 196, 6835–6840. [Google Scholar] [CrossRef]
  50. Sun, Q.; Yang, Y.; Fu, Z.W. Electrochemical Properties of Room Temperature Sodium–Air Batteries with Non-Aqueous Electrolyte. Electrochem. Commun. 2012, 16, 22–25. [Google Scholar] [CrossRef]
  51. Senthilkumar, B.; Khan, Z.; Park, S.; Seo, I.; Ko, H.; Kim, Y. Exploration of Cobalt Phosphate as a Potential Catalyst for Rechargeable Aqueous Sodium-Air Battery. J. Power Sources 2016, 311, 29–34. [Google Scholar] [CrossRef]
  52. Liang, F.; Qiu, X.; Zhang, Q.; Kang, Y.; Koo, A.; Hayashi, K.; Chen, K.; Xue, D.; Hui, K.N.; Yadegari, H.; et al. A Liquid Anode for Rechargeable Sodium-Air Batteries with Low Voltage Gap and High Safety. Nano Energy 2018, 49, 574–579. [Google Scholar] [CrossRef]
  53. Zhao, Y.; Goncharova, L.V.; Zhang, Q.; Kaghazchi, P.; Sun, Q.; Lushington, A.; Wang, B.; Li, R.; Sun, X. Inorganic–Organic Coating Via Molecular Layer Deposition Enables Long Life Sodium Metal Anode. Nano Lett. 2017, 17, 5653–5659. [Google Scholar] [CrossRef] [PubMed]
  54. Yu, J.; Hu, Y.S.; Pan, F.; Zhang, Z.; Wang, Q.; Li, H.; Huang, X.; Chen, L. A Class of Liquid Anode for Rechargeable Batteries with Ultralong Cycle Life. Nat. Commun. 2017, 8, 14629. [Google Scholar] [CrossRef] [PubMed]
  55. Sahgong, S.H.; Senthilkumar, S.; Kim, K.; Hwang, S.M.; Kim, Y. Rechargeable Aqueous Na–Air Batteries: Highly Improved Voltage Efficiency by Use of Catalysts. Electrochem. Commun. 2015, 61, 53–56. [Google Scholar] [CrossRef]
  56. Faktorovich Simon, E.; Natan, A.; Peled, E.; Golodnitsky, D. Oxygen Redox Processes in PEGDME-Based Electrolytes for the Na-Air Battery. J. Solid State Electrochem. 2018, 22, 1015–1022. [Google Scholar] [CrossRef]
  57. Hashimoto, T.; Hayashi, K. Aqueous and Nonaqueous Sodium-Air Cells with Nanoporous Gold Cathode. Electrochim. Acta 2015, 182, 809–814. [Google Scholar] [CrossRef]
  58. Kubisiak, P.; Eilmes, A. Molecular Dynamics Simulations of Ionic Liquid Based Electrolytes for Na-Ion Batteries: Effects of Force Field. J. Phys. Chem. B 2017, 121, 9957–9968. [Google Scholar] [CrossRef]
  59. Lutz, L.; Corte, D.A.D.; Tang, M.; Salager, E.; Deschamps, M.; Grimaud, A.; Johnson, L.; Bruce, P.G.; Tarascon, J.M. Role of Electrolyte Anions in the Na–O2 Battery: Implications for NaO2 Solvation and the Stability of the Sodium Solid Electrolyte Interphase in Glyme Ethers. Chem. Mater. 2017, 29, 6066–6075. [Google Scholar] [CrossRef]
  60. Goodenough, J.B.; Hong, H.Y.P.; Kafalas, J.A. Fast Na+-Ion Transport in Skeleton Structures. Mater. Res. Bull. 1976, 11, 203–220. [Google Scholar] [CrossRef]
  61. Kang, Y.; Zou, D.; Zhang, J.; Liang, F.; Hayashi, K.; Wang, H.; Xue, D.; Chen, K.; Adair, K.R.; Sun, X. Dual–Phase Spinel MnCo 2 O 4 Nanocrystals with Nitrogen-Doped Reduced Graphene Oxide as Potential Catalyst for Hybrid Na–Air Batteries. Electrochim. Acta 2017, 244, 222–229. [Google Scholar] [CrossRef]
  62. Khan, Z.; Parveen, N.; Ali, S.; Senthilkumar, S.T.; Park, S.; Kim, Y.; Hwan, M.; Ko, H. Electrochimica Acta Three-dimensional SnS 2 Nanopetals for Hybrid Sodium-Air Batteries. Electrochim. Acta 2017, 257, 328–334. [Google Scholar] [CrossRef]
  63. Sun, Q.; Liu, J.; Li, X.; Wang, B.; Yadegari, H.; Lushington, A.; Banis, M.N.; Zhao, Y.; Xiao, W.; Chen, N.; et al. Atomic Layer Deposited Non-Noble Metal Oxide Catalyst for Sodium-Air Batteries: Tuning the Morphologies and Compositions of Discharge Product. Adv. Funct. Mater. 2017, 27, 1606662. [Google Scholar] [CrossRef]
  64. Kim, M.; Ju, H.; Kim, J. Dihydrogen Phosphate Ion Functionalized Nanocrystalline Thallium Ruthenium Oxide Pyrochlore as a Bifunctional Electrocatalyst for Aqueous Na-Air Batteries. Appl. Catal. B Environ. 2019, 245, 29–39. [Google Scholar] [CrossRef]
  65. Su, F.; Qiu, X.; Liang, F.; Tanaka, M.; Qu, T.; Yao, Y.; Ma, W.; Yang, B.; Dai, Y.; Hayashi, K.; et al. Preparation of Nickel Nanoparticles by Direct Current Arc Discharge Method and Their Catalytic Application in Hybrid Na-Air Battery. Nanomaterials 2018, 8, 684. [Google Scholar] [CrossRef] [PubMed]
  66. Kim, M.; Ju, H.; Kim, J. Single Crystalline Bi2Ru2O7 Pyrochlore Oxide Nanoparticles as Efficient Bifunctional Oxygen Electrocatalyst for Hybrid Na-Air Batteries. Chem. Eng. J. 2019, 358, 11–19. [Google Scholar] [CrossRef]
  67. Wu, Y.; Qiu, X.; Liang, F.; Zhang, Q.; Koo, A.; Dai, Y.; Lei, Y.; Sun, X. A Metal-Organic Framework-Derived Bifunctional Catalyst for Hybrid Sodium-Air Batteries. Appl. Catal. B Environ. 2019, 241, 407–414. [Google Scholar] [CrossRef]
  68. Kim, M.; Ju, H.; Kim, J. Single Crystalline Thallium Rhodium Oxide Pyrochlore for Highly Improved Round Trip Efficiency of Hybrid Na–Air Batteries. Dalton Trans. 2018, 47, 15217–15225. [Google Scholar] [CrossRef]
  69. Yin, W.W.; Fu, Z.W. A Highly Efficient Bifunctional Heterogeneous Catalyst for Morphological Control of Discharged Products in Na–Air Batteries. Chem. Commun. 2017, 53, 1522–1525. [Google Scholar] [CrossRef]
  70. Shang, M.; Liu, Y.; Xia, J.; Zhang, S.; Yang, J. Synthesis and Characterization of MnCo2O4 Microspheres Based Air Electrode for Rechargeable Sodium-Air Batteries. Ceram. Int. 2017, 43, 3218–3223. [Google Scholar] [CrossRef]
  71. Abirami, M.; Hwang, S.M.; Yang, J.; Senthilkumar, S.T.; Kim, J.; Go, W.S.; Senthilkumar, B.; Song, H.K.; Kim, Y. A Metal–Organic Framework Derived Porous Cobalt Manganese Oxide Bifunctional Electrocatalyst for Hybrid Na–Air/Seawater Batteries. ACS Appl. Mater. Interfaces 2016, 8, 32778–32787. [Google Scholar] [CrossRef]
  72. Cheon, J.Y.; Kim, K.; Sa, Y.J.; Sahgong, S.H.; Hong, Y.; Woo, J.; Yim, S.; Jeong, H.Y.; Kim, Y.; Joo, S.H. Graphitic Nanoshell/Mesoporous Carbon Nanohybrids as Highly Efficient and Stable Bifunctional Oxygen Electrocatalysts for Rechargeable Aqueous Na-Air Batteries. Adv. Energy Mater. 2016, 6, 1501794. [Google Scholar] [CrossRef]
  73. Sang, L.; Yin, W.W.; Yue, J.L.; Cao, M.H.; Liu, W.; Ding, J.J.; Fu, Z.W. Dual Catalytic Behavior of a Soluble Ferrocene as an Electrocatalyst and in the Electrochemistry for Na–Air Batteries. J. Mater. Chem. A 2015, 3, 19027–19032. [Google Scholar]
  74. Sun, Q.; Yadegari, H.; Banis, M.N.; Liu, J.; Xiao, B.; Wang, B.; Lawes, S.; Li, X.; Li, R.; Sun, X. Self-Stacked Nitrogen-Doped Carbon Nanotubes as Long-Life Air Electrode for Sodium-Air Batteries: Elucidating the Evolution of Discharge Product Morphology. Nano Energy 2015, 12, 698–708. [Google Scholar] [CrossRef]
  75. Shadike, Z.; Sang, L.; Yin, W.W.; Yang, Y.; Ding, F.; Li, H.; Fu, Z.W. A Long-Life Na–Air Battery Based on a Soluble NaI Catalyst. Chem. Commun. 2015, 51, 2324–2327. [Google Scholar]
  76. Yadegari, H.; Li, Y.; Banis, M.N.; Li, X.; Wang, B.; Sun, Q.; Li, R.; Sham, T.K.; Cui, X.; Sun, X. On Rechargeability and Reaction Kinetics of Sodium–Air Batteries. Energy Environ. Sci. 2014, 7, 3747–3757. [Google Scholar] [CrossRef]
  77. Liu, W.M.; Yin, W.W.; Ding, F.; Sang, L.; Fu, Z.W. NiCo2O4 Nanosheets Supported on Ni Foam for Rechargeable Nonaqueous Sodium–Air Batteries. Electrochem. Commun. 2014, 45, 87–90. [Google Scholar] [CrossRef]
  78. Li, Y.; Yadegari, H.; Li, X.; Banis, M.N.; Li, R.; Sun, X. Superior Catalytic Activity of Nitrogen-Doped Graphene Cathodes for High Energy Capacity Sodium–Air Batteries. Chem. Commun. 2013, 49, 11731–11733. [Google Scholar] [CrossRef]
  79. Liu, W.; Sun, Q.; Yang, Y.; Xie, J.Y.; Fu, Z.W. An Enhanced Electrochemical Performance of a Sodium–Air Battery with Graphene Nanosheets as Air Electrode Catalysts. Chem. Commun. 2013, 49, 1951–1953. [Google Scholar] [CrossRef]
  80. Xu, S.; Lu, Y.; Wang, H.; Abruna, 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]
  81. Khan, Z.; Senthilkumar, B.; O Park, S.; Park, S.; Yang, J.; Lee, J.H.; Song, H.K.; Kim, Y.; Kwak, S.K.; Ko, H. Carambola-Shaped VO2 Nanostructures: A Binder-Free Air Electrode for an Aqueous Na–Air Battery. J. Mater. Chem. A 2017, 5, 2037–2044. [Google Scholar] [CrossRef]
  82. Hu, Y.; Han, X.; Zhao, Q.; Du, J.; Cheng, F.; Chen, J. Porous Perovskite Calcium–Manganese Oxide Microspheres as an Efficient Catalyst for Rechargeable Sodium–Oxygen batteries. J. Mater. Chem. A 2015, 3, 3320–3324. [Google Scholar] [CrossRef]
  83. Khan, Z.; Park, S.; Hwang, S.M.; Yang, J.; Lee, Y.; Song, H.K.; Kim, Y.; Ko, H. Hierarchical Urchin-Shaped α-MnO2 on Graphene-Coated Carbon Microfibers: A Binder-Free Electrode for Rechargeable Aqueous Na–Air Battery. NPG Asia Mater. 2016, 8, e294. [Google Scholar] [CrossRef]
  84. Wen, Z.; Rui, K.; Shen, C.; Zhang, S.; 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]
Figure 1. (a) Schematic illustration of RT -Na/S batteries [11] (b) Reaction stages during discharge for RT-Na/S batteries [8]. Reprinted with permission from Wiley, copyright 2019.
Figure 1. (a) Schematic illustration of RT -Na/S batteries [11] (b) Reaction stages during discharge for RT-Na/S batteries [8]. Reprinted with permission from Wiley, copyright 2019.
Electronics 08 01201 g001
Figure 2. (a) Schematic illustration of the synthesis procedure; (b) cyclic performance of Na/Span cells with 2M NaTFSI/TMP electrolytes at 1C [18]. Reprinted with permission from Wu, J.; Liu, J.; Lu, Z.; Lin, K.; Lyu, Y.Q.; Li, B.; Ciucci, F.; Kim, J.-K.; Non-flammable electrolyte for dendrite-free sodium-sulfur battery, Elsevier, copyright 2019.
Figure 2. (a) Schematic illustration of the synthesis procedure; (b) cyclic performance of Na/Span cells with 2M NaTFSI/TMP electrolytes at 1C [18]. Reprinted with permission from Wu, J.; Liu, J.; Lu, Z.; Lin, K.; Lyu, Y.Q.; Li, B.; Ciucci, F.; Kim, J.-K.; Non-flammable electrolyte for dendrite-free sodium-sulfur battery, Elsevier, copyright 2019.
Electronics 08 01201 g002
Figure 3. Schematic illustration of freestanding cathode synthesis and the role of the binder. (a) Schematic illustration of the free-standing cathode preparation. (b) Schematic of the proposed role of the polysaccharide binder to counteract volume expansion of sulfur during sodium-ion insertion [22]. Reprinted with permission from Ghosh, A.; Kumar, A.; Roy, A.; Panda, M.R.; Kar, M.; MacFarlane, D.R.; Mitra, S. Three-Dimensionally Reinforced Freestanding Cathode for High-Energy Room-Temperature Sodium–Sulfur Batteries. Copyright 2019, American Chemical Society.
Figure 3. Schematic illustration of freestanding cathode synthesis and the role of the binder. (a) Schematic illustration of the free-standing cathode preparation. (b) Schematic of the proposed role of the polysaccharide binder to counteract volume expansion of sulfur during sodium-ion insertion [22]. Reprinted with permission from Ghosh, A.; Kumar, A.; Roy, A.; Panda, M.R.; Kar, M.; MacFarlane, D.R.; Mitra, S. Three-Dimensionally Reinforced Freestanding Cathode for High-Energy Room-Temperature Sodium–Sulfur Batteries. Copyright 2019, American Chemical Society.
Electronics 08 01201 g003
Figure 4. Illustration of the NASICON structure of Na3Zr2Si2PO12 material and molecular structure of a polymer with intrinsic porosity (PIN) and specific discharge capacities and Coulombic efficiencies versus cycling number of a Na || PIN [17]. Reprinted with permission from Yu, X.; Manthiram, A. Sodium-Sulfur Batteries with a Polymer-Coated NASICON-type Sodium-Ion Solid Electrolyte; copyright 2019 from Elsevier.
Figure 4. Illustration of the NASICON structure of Na3Zr2Si2PO12 material and molecular structure of a polymer with intrinsic porosity (PIN) and specific discharge capacities and Coulombic efficiencies versus cycling number of a Na || PIN [17]. Reprinted with permission from Yu, X.; Manthiram, A. Sodium-Sulfur Batteries with a Polymer-Coated NASICON-type Sodium-Ion Solid Electrolyte; copyright 2019 from Elsevier.
Electronics 08 01201 g004
Figure 5. Schematic representation of a typical sodium-air battery.
Figure 5. Schematic representation of a typical sodium-air battery.
Electronics 08 01201 g005
Figure 6. Schematic structure of (a) non-aqueous and (b) aqueous Na-air batteries with nanoporous gold electrode. (c) First discharge charge curves of non-aqueous and aqueous batteries at 0.5 mA·cm−2 up to 7 mAh·g−1NPG [57]. Reprinted from Aqueous and Nonaqueous Sodium-Air Cells with Nanoporous Gold Cathode, 182; Taijyu Hashimoto, Katsuro Hayashi; Copyright 2019, with permission from Elsevier.
Figure 6. Schematic structure of (a) non-aqueous and (b) aqueous Na-air batteries with nanoporous gold electrode. (c) First discharge charge curves of non-aqueous and aqueous batteries at 0.5 mA·cm−2 up to 7 mAh·g−1NPG [57]. Reprinted from Aqueous and Nonaqueous Sodium-Air Cells with Nanoporous Gold Cathode, 182; Taijyu Hashimoto, Katsuro Hayashi; Copyright 2019, with permission from Elsevier.
Electronics 08 01201 g006
Figure 7. Illustration of the suggested sodium solid electrolyte interface (SEI) formation mechanism and its composition in different DME (1,2-dimethoxyethane) electrolytes [59]. Reprinted with permission from ACS, Copyright 2017.
Figure 7. Illustration of the suggested sodium solid electrolyte interface (SEI) formation mechanism and its composition in different DME (1,2-dimethoxyethane) electrolytes [59]. Reprinted with permission from ACS, Copyright 2017.
Electronics 08 01201 g007
Table 1. Literature review of components and performance of various Na-sulfur batteries.
Table 1. Literature review of components and performance of various Na-sulfur batteries.
CathodeElectrolyteCurrent RateCapacity at cycle 1/mAh g−1Capacity/mAh g−1 after (n) cyclesRef
Sulfurized PAN (S-PAN)1M NaClO4 in EC-DMC0.1 mA cm−2654.8500 (18)[12]
Sulfurized PAN nanofiber Sulfur 41wt%1M NaPF6 in 1:1 EC:DMC0.1C756648 (200)[16]
Sulfurized CNFPolymer with intrinsic porosity(PIN) coated on Na3Zr2Si2PO12C/5~650550 (100)[17]
Polyacrylonitrile (PAN): S 1:82M NaTFSI/TMP in FEC1C~1000788 (300)[18]
C-PANS: Super P: PVDF0.8M NaClO4 in EC/EMC220 mA g−1364150 (500)[19]
S@Con-HC: Carbon Black: Carboxymethyl cellulose (CMC) 70:10:201M NaClO4 in PC/EC + 5wt% FEC100 mA g−11081508 (600)[20]
Sulfur nanoparticles: Super P: PVDF 60:30:10 soaked carbon fiber composites (CFC)1.5M NaClO4 & 0.2M NaNO3 in TEGDME0.1 C390120 (300)[21]
Multiporous carbon fibers (MPCFs): Super-P: Sodium carboxymethyl cellulose (CMCNa) 80:10:102M NaTFSI in TEGDME and PC:FEC (1:1 vol) as co-solvent0.1C16351134 (50)[11]
rGO/S/MnxOy@SA-PANI (Na Alginate/polyaniline)1M NaClO4 with 0.2M NaNO3 in TEGDME0.2 Ag−1631535 (50)[22]
S/CNT@MPC: Super P: PVDF1M NaClO4 in PC/EC3.34 Ag−11610500 (200)[23]
S-hollow C: Carbon: PEONaTFSI in TEGDME167 mA g−11000600 (20)[24]
S@interconnected mesoporous C hollow spheres (iMCHS): CB: CMC1M NaClO4 in EC/PC + FEC100 mA g−1328.4292 (200)[25]
S/Cu-decorated mesoporous C: C: CMC1M NaClO4 in EC/DMC50 mA g−1718641 (110)[26]
S/microporous C polyhedron composites(MCPS): CB: PVDF1M NaClO4 in EC/PC84 mA g−1689354 (100)[27]
Table 2. Literature review of components and performance of various Na-air batteries. Lithium metal was the anode in all cases.
Table 2. Literature review of components and performance of various Na-air batteries. Lithium metal was the anode in all cases.
CathodeCatalystSeparatorElectrolyteCurrent DensityCyclesOver-Potential (V)Round Trip Efficiency (%)Ref
Catalyst on nickel meshDual-phase spinel MnCo2O4 with nitrogen-doped reduced graphene oxide hybrid (dp-MnCo2O4/N-RGO)NASICON1 M NaClO4 in EC/DMC (1:1) with 1 vol% FEC as anolyte. 1 M NaOH as catholyte0.13 mA.cm−2250.3987.6[61]
Carbon blackPt/CNASICON1 M NaClO4 in EC/DMC (1:1) with 1 vol% FEC as anolyte. 1 M NaOH as catholyte0.13 mA.cm−20.62--[61]
Catalyst on nickel meshNitrogen-doped reduced graphene oxide hybrid (N-RGO)NASICON1 M NaClO4 in EC/DMC (1:1) with 1 vol% FEC as anolyte. 1 M NaOH as catholyte0.13 mA.cm−20.6-[61]
Catalyst on nickel meshA dual-phase spinel MnCo2O4 (DP- MnCo2O4)NASICON1 M NaClO4 in EC/DMC (1:1) with 1 vol % FEC as anolyte. 1 M NaOH as catholyte0.13 mA.cm−20.49-[61]
Carbon blackNASICON1 M NaClO4 in EC/DMC (1:1) with 1 vol% FEC as anolyte. 1 M NaOH as catholyte0.13 mA.cm−21.34-[61]
3D flower-shaped sns2 nanostructures, called as nanopetalsNASICON0.1 M NaOH electrolyte5 mA.g−1400.5283[62]
Co3O4 decorated carbon nanotubes (CNT@Co3O4)Nanostructured cobalt oxide (Co3O4)Polypropylene (PP)-based polymerNaCF3SO3/diethylene glycol dimethyl ether (DEGDME)150 mA.g−112 cycles at a cut off of 300 mA.h.g−1 capacity0.4-[63]
CNT/Co3O4 nanocomposite powders prepared by grinding Co3O4 nanoparticles with CNTsCo3O4Polypropylene (PP)-based polymerNaCF3SO3/diethylene glycol dimethyl ether (DEGDME)150 mA.g−112 cycles at a cutoff of 300 mA.h.g−1 capacity0.4-[63]
CNT-Polypropylene (PP)-based polymerNaCF3SO3/diethylene glycol dimethyl ether ((DEGDME)150 mA.g−113 cycles at a cutoff of 300 mA.h.g−1 capacity0.6-[63]
Catalyst on glassy carbon electrodeTl2Ru2O7 functionalized using dihydrogen phosphate ion (P-Tl2Ru2O7)NASICON1M NaCF3SO3/TEGDME, 0.1 M KOH0.01 mA.cm−2500.1695[64]
Catalyst on glassy carbon electrodeTl2Ru2O7NASICON1M NaCF3SO3/TEGDME, 0.1 M KOH0.01 mA.cm−2-[64]
20 wt% Pt on Vulcan carbon blackPlatinum (Pt)NASICON1M NaCF3SO3/TEGDME, 0.1 M KOH0.01 mA.cm−2-0.25~90[64]
20 wt% Ir on VulcanIrradium (Ir)NASICON1M NaCF3SO3/TEGDME, 0.1 M KOH0.01 mA.cm−2-0.3~90[64]
Catalyst on carbon paperCarbon blackSolid conductor with Na3Zr2Si2PO121 M NaClO4 in EC/DMC (1:1) with 1 vol% of fluoroethylene carbonate as anolyte, 1 M NaOH as catholyte0.1 mA.cm−2-1.05[65]
Catalyst on carbon paperSilver nanoparticlesSolid conductor with Na3Zr2Si2PO121 M NaClO4 in EC/DMC (1:1) with 1 vol% of fluoroethylene carbonate as anolyte, 1 M NaOH as catholyte0.1 mA.cm−2-0.75[65]
Catalyst on carbon paperAir electrode containing nickel nanoparticles as catalystsSolid conductor with Na3Zr2Si2PO121 M NaClO4 in EC/DMC (1:1) with 1 vol% of fluoroethylene carbonate as anolyte, 1 M NaOH as catholyte0.1 mA.cm−21000.5780.0% to 69.3%[65]
Bi2Ru2O7 (90 wt%) with polyvinylidene fluoride (10 wt%) as a binder on one side of the Teflon-treated carbon paperNovel nanocrystalline bismuth ruthenate pyrochlore oxide (Bi2Ru2O7) as an effective oxygen electrocatalystNASICON1M NaCF3SO3/TEGDME0.01 mA cm−2500.2193.58[66]
Catalyst on conductive carbon powderMetal-organic framework-derived N-doped carbon nanotubes (MOF-N-CNTs)NASICON1M NaCF3SO3/TEGDME0.1 mA.cm−2350.387[67]
CarbonPt/C (Platinum on carbon)NASICON1M NaCF3SO3/TEGDME0.1 mA.cm−2-0.5[67]
CarbonRuthenium oxide (RuO2)NASICON1M NaCF3SO3/TEGDME0.1 mA.cm−2-0.5[67]
CNTCo-CNTs (Cobalt on carbon nanotubes)NASICON1M NaCF3SO3/TEGDME0.1 mA.cm−2-0.67[67]
N-CNTs (nitrogen-doped carbon nanotubes)NASICON1M NaCF3SO3/TEGDME0.1 mA.cm−2-0.77[67]
Multi-walled carbon nanotubes (MWCNTs)NASICON1M NaCF3SO3/TEGDME0.1 mA.cm−2-0.9[67]
Carbon paperNASICON1M NaCF3SO3/TEGDME0.1 mA.cm−2-1.18[67]
Ti2Rh2O7 (90 wt%) with polyvinylidene fluoride (10 wt%) binder on Teflon-treated carbon paperSingle crystalline thallium rhodium pyrochlore oxide (Tl2Rh2O7) as a highly efficient bifunctional electrocatalystNASICON1 M NaCF3SO3/TEGDME120 mA.g−1500.20893.65[68]
CNT@Co3O4 or CNT/Co3O4Co3O4 is decorated on carbon nanotubes by ALD (CNT@Co3O4)Polypropylene (PP)-based polymerNaCF3SO3/diethylene glycol dimethyl ether (DEGDME)150 mA.g−113–140.4[63]
Co3O4 nanowires/CHeterogeneous catalysts with Co3O4 and liquid redox mediatorsFerrocene added to 0.05M Fe(C5H5)2 in 1M NaClO4/1,2-dimethoxyethane (DME)500 mA.g−15700.6[69]
MnCo2O4/C air electrode with MnCo2O4, Super P and PVDFNanosized porous MnCo2O4 microspheresGlass fiber1.0 M NaClO4 in propylene carbonate (PC)0.1 mA cm−21301.7[70]
Carbon black with nickel nanoparticle catalystsAir electrode containing nickel nanoparticles as catalystsSolid conductor with Na3Zr2Si2PO121 M NaClO4 in EC/DMC (1:1) with 1 vol % of fluoroethylene carbonate as anolyte, 1 M NaOH as catholyte0.1 mA cm−21000.57[65]
Porous cobalt manganese oxide (CMO) nanocubes CMO nanoparticles, carbon black Super-P and PVDFSpinel-type porous cobalt manganese oxide (CoxMn3−xO4, CMO) nanocubes as a noble-metal-free electrocatalystNASICON ceramic electrolyteused as Na+ ion conductive membrane to separate the anode from the cathodeNASICON ceramic electrolytes; 1 M NaCF3SO3 in TEGDME0.01 mA cm−21000.53∼85%[71]
Cobalt phosphate, conductive carbon black Super-P, PVDFCo3(PO4)2 nanostructureNASICON separator1 M NaCF3SO3 in tetra ethylene glycol dimethyl ether0.01 mA cm−2500.23~93%[51]
Mesoporous carbonGraphitic nanoshell/mesoporous carbon (GNS/MC) nanohybridsNa super ionic conducting (NASICON) solid electrolyte1.0 M NaCF3SO3/TEGDME (nonaqueous electrolyte) and 0.1 M NaOH aqueous solution as an anolyte and a catholyte respectively60 mA.g−1100.11596.28[72]
Pt/C (Platinum on carbon)Na super ionic conducting (NASICON) solid electrolyte1.0 M NaCF3SO3/TEGDME (nonaqueous electrolyte) and 0.1 M NaOH aqueous solution were used as an anolyte and a catholyte60 mA.g−1-0.17994.41[72]
Ir/CNa super ionic conducting (NASICON) solid electrolyte1.0 M NaCF3SO3/TEGDME (nonaqueous electrolyte) and 0.1 M NaOH aqueous solution were used as an anolyte and a catholyte60 mA.g−10.36488.77[72]
Carbon paper w/o catalystN/ANa super ionic conducting (NASICON) solid electrolyte1.0 M NaCF3SO3/TEGDME (nonaqueous electrolyte) and 0.1 M NaOH aqueous solution were used as an anolyte and a catholyte60 mA.g−10.69879.47[72]
Highly graphitic nanoshells embedded in mesoporous carbon (GNS/MC) sprayed on Teflon-treated carbon paper used as the gas diffusion layerHighly graphitic nanoshells embedded in mesoporous carbon (GNS/MC)NASICON1.0 M NaCF3SO3/TEGDME0.01 mA.cm−2100.72[72]
Vulcan XC72RN/ANASICON3 M NaCF3SO3/TEGDME0.01 mA.cm−218~0.678[55]
Pt/C-coated carbon paperPt/CNASICON3 M NaCF3SO3/TEGDME0.01 mA.cm−218~0.472.4[55]
Nanoporous gold (NPG) air electrode-NASICON1 M NaClO4 and 1 vol% fluoroethylene carbonate as an additive in EC/DMC with a ratio of 1:1 for the organic electrolyte. 1 M NaOH solution as aqueous electrolyte.0.5 mA cm−2~10~1.0[57]
Carbon nanotube (CNT) @Ni electrodeN/ANASICONNaPF6/DME electrolyte200 mA.g−110 cycles at cutoff capacity of 1200 mAh.g−1~1.8[73]
Self-stacked nitrogen-doped carbon nanotubes (ncnts) on commercial porous polypropylene substratesN/ACelgard35010.5 M sodium triflate (NaSO3CF3) dissolved in DEGDME75 mA.g−110 cycles at cutoff capacity of 300 mAh.g−1~2.0[74]
CNT/Ni foamN/A1.0 M NaClO4 in 1,2-dimethoxyethane (DME)150 mA.g−131.2[75]
N/A-0.5 M solutions of NaCF3SO3 in DEGDME, TEGDME, and DME as ether-based electrolytes100 mA.g−120 cycles at cutoff capacity 0f 500 mAh.g−1~1.8-[76]
NiCo2O4 nanosheets were directly deposited on the Ni foamN/A-0.05 M NaI containing DME (1.2-dimethoxyethane) electrolyte500 mA.g−1150 cycles with a capacity limit of 1000 mA.h.g−1~0.8-[77]
Nitrogen-doped graphene nanosheets (N-GNSs) displayedN/ACelgard 35010.5 M sodium triflate (NaSO3CF3) dissolved in diethylene glycol diethyl ether (DEGDME)130 mA.g−110--[78]
Graphene nanosheetsN/A-0.05 M Fe(C5H5)2-containing electrolyte in 1.0 M NaClO4 in a nonaqueous solution of 1,2-dimethoxyethane (DME)500 mA.g−1230 cycles with a high capacity of 1000 mA.h.g−1~1.0-[79]
Carbon black on Al meshGlass fiber separator1 M solutions of NaTFSI in propylene carbonate (PC) electrolyte200 mA.g−120 cycles at cutoff of 800 mAh.g−1-47[80]
Carambola-shaped vanadium oxide (VO2) nanostructures on carbon paperVanadium oxide (VO2)NASICON1 M NaCF3SO3 in TEGDME4 mA. g−1500.6481[81]
Porous CaMnO3/CCalcium–manganese oxide microspheresNASICONNaSO3CF3/TEGDME100 mA g−180 cycles at cut off of 1000 mAh.g−1~1.055[82]
Urchin-shaped MnO2 on the RGO-coated carbon microfiberUrchin-shaped MnO2 on RGONASICON0.1 M NaOH aqueous catholyte. NaCF3SO3 in TEGDME as anolyte15 mA.g−1200.781[83]
Pt@graphene nanosheetsPlatinumNASICON1 M NaClO4/PC0.1 mA.cm−210 cycles at a cut off of 1000 mAh.g−1~1.568[84]

Share and Cite

MDPI and ACS Style

Chawla, N.; Safa, M. Sodium Batteries: A Review on Sodium-Sulfur and Sodium-Air Batteries. Electronics 2019, 8, 1201. https://doi.org/10.3390/electronics8101201

AMA Style

Chawla N, Safa M. Sodium Batteries: A Review on Sodium-Sulfur and Sodium-Air Batteries. Electronics. 2019; 8(10):1201. https://doi.org/10.3390/electronics8101201

Chicago/Turabian Style

Chawla, Neha, and Meer Safa. 2019. "Sodium Batteries: A Review on Sodium-Sulfur and Sodium-Air Batteries" Electronics 8, no. 10: 1201. https://doi.org/10.3390/electronics8101201

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