Phosphate-based polyanionic insertion materials for oxygen electrocatalysis

Ritambhara Gond *a, Jiefang Zhu a and Prabeer Barpanda *b
aDepartment of Chemistry – Ångström Laboratory, Uppsala University, Box 538, 75121 Uppsala, Sweden. E-mail: ritambhara.gond@kemi.uu.se
bFaraday Materials Laboratory (FaMaL), Materials Research Centre, Indian Institute of Science, Bangalore 560012, India. E-mail: prabeer@iisc.ac.in

Received 8th October 2023 , Accepted 9th January 2024

First published on 9th January 2024


Abstract

Electrocatalyst-based energy storage technologies such as alkali metal–air batteries, fuel cells, and water splitting devices are the new holy grail in the next-generation energy storage landscape as they deliver higher energy densities than Li-ion/Na-ion batteries (LIBs/SIBs). The new chemistries of energy storage such as metal–air batteries under aqueous or non-aqueous conditions will complement existing LIBs/SIBs owing to the increasing requirement for batteries with high energy density in the present era. Phosphate-based polyanionic frameworks have long been known for their ability to (de)intercalate alkali metal ions. Because of their innate oxygen electrocatalytic activity, these insertion cathode materials have lately emerged as air electrodes in metal–air battery systems. In this review, the present status of phosphate-based polyanionic insertion materials for oxygen reduction and oxygen evolution reaction (ORR and OER) electrocatalysis is summarized. Factors influencing electrocatalytic activity in these materials, such as the presence of different types of alkali metal cations, transition metals, and the type of ligand/mixed anion as well as coordination around the transition metals are discussed. Finally, the development of metal–air batteries derived from phosphate-based polyanionic insertion materials as air electrodes is discussed.


1. Introduction

Energy storage from lithium-ion batteries (LIBs) needs no introduction and validation in the present era as they are ubiquitous in the modern world. The ever-growing consumption of LIBs has triggered the pressure to overcome the existing supply limitations and has opened more markets and industries for other analogues of lithium-based energy storage devices. In this era of post-Li-ion batteries, Li–O2 batteries form a potential candidate as they exhibit the highest energy density, capable of achieving the ultimate goal of batteries for long-range electric vehicles (EVs). For this application, O2 gas should ideally be replaced with air from Earth's ambient atmosphere.1 In 1996, Abraham et al. introduced the first Li–O2 battery employing an organic polymer electrolyte as an O2 permeable membrane.2 Since then, numerous advancements have been made at the architecture level to explore efficient Li–air batteries. However, Li–O2/air battery technologies are still in their infancy because of poor cyclability, large overpotential during the charging process, and the existence of other side reactions forming parasitic products owing to either unstable electrolytes or inefficient electrocatalysts.1,3–8

Practically, perovskite oxide-based electrodes are the most active and widely studied catalysts to perform oxygen electrocatalytic reactions (oxygen reduction reaction (ORR) and oxygen evolution reaction (OER)).9–11 Exhibiting efficient O2 electrocatalysis, the transition metal phosphate (PO4)3− class of polyanionic materials has also emerged as a potential candidate for possible application as cathodes in metal–air batteries.12 Despite respectable current densities and promising overpotentials delivered by these (PO4)3−-based catalysts, these materials suffer from poor stability in highly alkaline aqueous electrolytes. To the best of our knowledge, there is no literature covering different types of transition metal (PO4)3− based polyanionic materials showing intercalation chemistry and acting as catalysts for oxygen electrocatalysis. In light of this, this review offers a thorough summary of the research work to date on transition metal phosphate-based polyanionic materials for electrocatalytic oxygen reduction, oxygen evolution, and metal–air battery applications.

Electrocatalysts are known to lower the activation energy of oxygen reduction and water-splitting reactions. In principle, ORR/OER processes show different reaction pathways in acidic and alkaline aqueous environments. These pathways are represented as follows:13

The oxygen half-cell reaction at alkaline pH:

image file: d3qm01088k-t1.tif
The oxygen half-cell reaction at acidic pH:
image file: d3qm01088k-t2.tif
Based on linear free energy relationships (LFERs) phenomenologically employed by ab initio computations to forecast catalytic activity trends of reaction on metal surfaces, the four-proton coupled electron transfer steps for ORRs/OERs at the surface site (marked by * symbol) can be expressed as follows:

Reactions at alkaline pH:

image file: d3qm01088k-t3.tif

image file: d3qm01088k-t4.tif

image file: d3qm01088k-t5.tif

image file: d3qm01088k-t6.tif
Reactions at acidic pH:
image file: d3qm01088k-t7.tif

image file: d3qm01088k-t8.tif

image file: d3qm01088k-t9.tif

image file: d3qm01088k-t10.tif
Here, OOH*, O*, and OH* are the reaction intermediates undergoing desorption/adsorption. Due to the intrinsic stability of phosphates against temperature and pH, the electrocatalytic properties of most of the PO4-based polyanionic insertion materials investigated in an alkaline medium make the reaction pathways under alkaline pH more important for discussion in the later sections.

On another note, to realise a sustainable circular energy economy, it is highly important to recycle present batteries such as EV batteries (containing LiFePO4, LiCoO2, etc., as cathodes) when they reach their end of life. Recycling the electrode materials in the present battery market is of utmost importance. Instead of subjecting the coated electrode materials to toxic acid treatment to extract the remnant precursor for further synthesis, the recycled materials can directly be used in various applications. In this spirit, this review will also briefly highlight a different approach of using the cycled cathodes from the dead batteries towards the application in energy storage devices based on ORRs and OERs. This review outlines various oxide and polyanionic insertion materials demonstrating electrocatalytic activity with emphasis on phosphate-based polyanionic battery insertion compounds. It will begin with the discussion on environmental benignity, materials economy, and efficient scalable synthesis of PO4-based materials. Highlighting oxygen electrocatalysis chemistry (ORR and OER) at alkaline pH, the specific discharge capacity of a variety of PO4-based polyanionic insertion materials will also be stressed. The electrocatalytic activity (ORR and OER) is found to be greatly influenced by the type of transition metal (Mn, Fe, Co and Ni) present as well as the neighbouring alkali cation and PO4 units. We also incorporated fluorophosphates, which have an F-atom bonded to the transition metal centre, showing superior electrochemical activity. Eventually, all these catalysts (oxides or phosphates) cause electrodeposition of metal-oxides/oxy(hydroxide)/phosphate at the surface that cover the electrocatalyst. The underlying mechanism was correlated with the molecular orbital (MO) concepts, eg filling theory and metal-oxygen covalency followed by transition metal coordination and its connection with PO4-units (corner-shared or edge-shared).9

Phosphate-based materials with bifunctional electrocatalytic activities are further extended to the application encompassing hybrid Na–air batteries and Zn–air batteries. These batteries employ PO4 materials as air cathodes and function based on the underlying ORR and OER mechanisms to store energy. These materials can open up a hitherto untapped route for advancement in the research and development of Li–O2/Li–air batteries, which are still in their infancy marred with several challenges.14 Similar to hybrid Na–air batteries and Zn–air batteries, in the case of Li–O2/Li–air batteries, the cathode reaction overpotentials are much higher than those of the anode reaction.15,16 Various groups have reported the use of carbonaceous materials, precious metals/metal-oxides and composite materials as cathodes for non-aqueous Li–O2/Li–air batteries, which may require sophisticated synthesis. In this scenario, the PO4-based insertion materials, with their attractive material/process economy, can work as potential cathodes.9,11–19

2. Why phosphate-based polyanion materials?

Mother Nature has already considered and proven the major significance of phosphates in the vital events of life. We are all aware of the critical role that phosphorous plays in every biological or physiological event of energy gain or loss.20 The maxim by Alexander Todd, who won the Nobel Prize in 1958, “Where there is life, there is phosphorus” is unquestionably true. The vital roles of phosphates, enlisted by Bowler et al., also showed phosphate as the most important entity from genome stability and energetics to regulation and signalling.21 Inspired by the article by Knouse et al. and highly encouraged to choose phosphates, P(5+) oxidation state by chemists, the next-generation energy-driven materials can also be phosphate-based systems offering ample room for exploration targeting continuously growing demands of energy storage devices.22,23 Since the recognition of LiFePO4 as a cathode material, it has attracted the attention of electric vehicle (EV) industries due to its cheapness and thermal stability. In the beginning, major car industries chose LiCoO2/layered oxides as cathodes like major portable electronics sector, causing significant price rise for Co-precursors. Owing to the abundance of Fe-based resources, currently LiFePO4 has enabled low-cost consumer electronics and electric vehicles.24 An imaginary scheme of energy storage in a biological system is compared with the physical world's energy storage employing phosphate-based cathode materials that are synthesized in industries propelling several energy storage devices (Fig. 1).
image file: d3qm01088k-f1.tif
Fig. 1 Design of phosphate-based materials for energy storage and hereditary carriers inside living beings at the biological level and industrial-scale man-made cathode materials such as LiFePO4 for sustainable batteries at the industrial level.25,26

Perovskites (ABX3, A = cation+, B = transition metal+, X = anion) are known to catalyse various chemical and electrochemical reactions such as oxidation of CO, hydrocarbons and NOx, reduction of O2, N2, and CO2 and photo/electrochemical splitting of H2O.9 One of the key difficulties is to rationalise the design of nanostructured perovskite-type oxide-electrocatalysts from non-precious transition metals with least overpotentials for ORR and OER processes.13 The central challenge for oxygen electrochemical technologies is to achieve descent long-term stability of catalysts due to the dissolution of the catalyst surface during the ORR or OER. Here, PO4-based materials can offer stable operation during oxygen electrocatalysis.27,28 Among several reported materials for catalysing O2 (either ORRs or OERs or both), recently the number of papers based on phosphates has dramatically increased. This review primarily focuses on phosphate-based polyanion materials for oxygen electrocatalysis. Following, the major advantages of phosphate compounds are assessed and discussed in detail.

2.1. Environmentally benign

Currently, no systematic experimental techniques are used across the community to compare the electrocatalytic activities based on materials rather than those that evaluate the performance of the overall device.29,30 However, if we consider only the best-performing oxygen electrocatalysts, IrO2 and RuO2 for OERs and Pt/C for ORRs are the best-known conventional state-of-the-art catalysts whose activity is highly influenced by the pH value. Further, they are very expensive and may not be suitable for large-scale practical applications such as in metal–air batteries or other devices. Additionally, LaxSryMn1−yO3 perovskite type electrocatalysts are widely used in the OER-driven next-generation energy storage systems.11 However, the environmental cost of these cutting-edge technologies will be significantly impacted if the substance comprises rare earth metals such as La/Sr or other 4d elements.31,32 Since Li-rich metal-oxides undergo anion redox reactions at high operating potentials, polyanionic materials based on phosphate are both much safer and less harmful to the environment than oxides if comprised of 3d-tranistion metals.33 Because of strong covalent P–O bonds in phosphate-based polyanionic materials, they have the least possibility of O2 gaseous evolution caused due to anion redox process at higher potentials.34,35 Thus, the limiting nature of oxide cathode materials has stimulated the search of new phosphate-based polyanionic materials to cater the demand of high energy density vs. cost equilibrium while offering operational safety at a higher voltage.

2.2. Direct synthesis

Typically phosphate-based intercalation materials can be simply prepared via a low-cost, solvothermal ‘splash combustion synthesis route’, which is the most straightforward method for achieving carbon-coated nano-range particle sizes.36,37 In general, solution combustion is sufficiently adapted as an unsophisticated technique for nanoscale material synthesis to satisfy the requirement of charge conductivity and pulverization inhibition of the electrodes for energy conversion and storage devices.38,39 For instance, Bashir et al. have summarised the unique properties or combinations of properties offered by nanoscale electrodes for various energy storage devices.40 They emphasized the merits of size reduction, such as how nanoparticulate electrodes or electrocatalysts increase electrode/electrolyte contact areas, spatial confinement and surface contribution, ultimately increasing electrode reaction rates, which are crucial for advanced energy storage devices.

There have been several reports where the electrocatalysts require complex strategies such as CVD (chemical vapour deposition) and AVD (atomic vapour deposition) methods for the synthesis of catalysts.41–43 Nonetheless, it is cumbersome to implement them to increase the loading/structuring/active site of catalysts as well as enhanced intrinsic activity of each active site, which is necessary to achieve better material performances at the device level.44 Mohamed et al. showed the synthesis of ternary spinel MCo2O4 (M = Mn, Fe, Ni, and Zn) porous nanorods by a hydrothermal reaction exhibiting bifunctional (ORR and OER) activity, suitable for Li–O2 batteries.45 For the synthesis of precious metal oxides or highly crystalline multimetallic nanocrystals, as being highly efficient for OER or ORR activity, multistep synthesis has been opted with the use of highly toxic oleylamine solvent as a stabilizing agent.46–48 Thus, despite aforementioned complex strategies, it would be encouraging to obtain cost-efficient and durable phosphate-based materials prepared via a ultrafast facile combustion route, where the proposed catalysts would also be carbon coated and suitable for high-performance electrocatalysis devices.49–51

2.3. Recycling

Recycling old batteries is currently a critical necessity. Irrespective of the purity at the end of the life of a battery, the cathode materials are primarily subjected to metallurgical processes as well as direct repair processes for the reclamation of materials.52 Recycling will not only reduce the battery waste, but also mitigate the harmful effects of mining activities as well as critical metal sustainability in our high-tech society.53–55 The metallurgical way to recycle spent cathodes such as pyrometallurgical and hydrometallurgical approaches demand acid/alkaline leaching and cause huge greenhouse gas emissions due to high temperature and energy to smelt transition metals with toxic by-products. However, the direct repair approach is straightforward to use the cycled cathodes as such irrespective of the impurity. Liu et al. provided an innovative solution to avoid the use of corrosive acids showing acid-free mechanochemically induced isomorphic substitution of Na in LiFePO4, forming NaFePO4 and recreating Li metal precursor (Li2CO3).56 Phosphate-based intercalation materials have the potential to be employed in electrocatalytic devices because of their promising stability brought about by a strong P–O bond, in contrast to the instability of Pt/C electrocatalysts in other types of energy storage devices.27,57 In order to fabricate air-electrodes for catalysis-driven energy storage devices, it is simple to reclaim the used insertion cathode materials from Li/Na-ion batteries.

3. Battery insertion materials as oxygen electrocatalysts

First-row transition metal oxides, whether they are spinels or perovskites having edge-shared or corner-shared octahedra or tetrahedra, can exhibit comparable OER and ORR catalytic activities to those of RuO2 and IrO2.58 In addition to their usage in oxygen electrocatalysis-based devices, capable of reversible Li+ (de)intercalation into their frameworks, the first-row transition metal oxides (Mn, Fe, Co, and Ni) have also embarked a remarkable history in the development of LIBs.59 Moving from the layered oxides to a three-dimensional polyanionic framework system, LiFePO4 triggered the Li+ ion storage in olivine frameworks, where transition metal oxides (TM–O6) have edge sharing as well as corner sharing with phosphate (PO4) units. Shao-Horn group for the first time examined and compared the OER activities, at pH 7 and 13, of well-known LIB cathode insertion materials: oxides vs. olivine.60 They demonstrated superior activity in Co-based materials, particularly LiCoO2 and LiCoPO4 (Table 1 and Fig. 2a). Whether it is oxide or phosphate, it involves the formation of an amorphous surface layer on the catalyst. However, in case of phosphates, these amorphous surface layers are thicker at a higher pH than that of oxides due to more P being leached from the material, as shown in Fig. 2b.
Table 1 Overview of ORR and OER (vs. RHE) activities for crystalline phosphate-based-polyanionic intercalation materials in alkaline solutions
Materials Crystal structure (space group) Intercalation-based electrochemical activity Electrocatalytic activity Electrolyte solution Ref.
ORR OER
Specific capacity (mA h g−1) Working potential (V) Ref. E onset (V) E 1/2 (V) n e− (no. of electrons) Tafel slope (mV dec−1) E onset (V) η@10 mA cm−2 (mV) Tafel slope (mV dec−1)
With no alkali metal-cation
Co2P2O7/NPGA B121/c1 Inactive 0.91 0.80 3.72 1.45 340 96 0.1 M KOH 65
Co3(PO4)2/NC P121/c1 Inactive 0.96 0.83 3.99 56 296 53 0.1 M KOH 89
ZnCo2(PO4)2 P21/n Inactive 0.87 0.73 3.90 450 0.1 M KOH 66
ZnCoP2O7 P21/n Inactive 0.86 0.72 3.82 460 0.1 M KOH 66
Li-cation based
LiFePO4 Pnma 162 3.45 90 N.A. 1.52 0.1 M KH2PO4 60
LiMnPO4 Pnma ∼91 4.10 91 0.85 0.89 OER inactive 0.1 M KOH 92
LiCoPO4 Pnma ∼93 ∼4.8 91 N.A. 1.60 60 0.1 M KOH 60, 63 and 70
N.A. 1.68 120 0.1 M KH2PO4 60
Li0.7Co0.75Fe0.25PO4/rGO Pnma 3.4 & 4.8 69 N.A. 1.61 53 0.1 M KOH 69
LiCo0.60Fe0.40PO4 Pnma ∼135 3.4 & 4.8 68 N.A. 1.47 345 44 0.1 M KOH 68
LiNiPO4 Pnma ∼120 4.93 & 5.27 71 N.A. 1.65 1 M KOH 67
LiNi0.75Fe0.25PO4/rGO Pnma N.A. N.A. 1.45 295 47 1 M KOH 67
Li2CoP2O7 P21/c ∼85 4.93 & ∼4.9 93 N.A. 1.81 84 0.5 M Na2HPO4 63
Na-cation based
NaCoPO4 Pnma Inactive 94 0.77 0.69 - 109 N.A. 0.1 M NaOH 63
0.85 0.73 3.58 87 1.55 390 52 1 M NaOH 73
NaFePO4 Pnma ∼100 2.88 & 3.02 95 1.04 0.79 3.87 ∼94 OER inactive 0.1 M KOH 96
NaFe(PO3)3 Pa[3 with combining macron] ∼22 2.8 97 ORR Inactive 550 90 1 M KOH 78
NaMn(PO3)3 N.A. ORR Inactive 189
NaCo(PO3)3 ∼56 3.2 77 1.54 340 76
0.83 60 1.52 77 0.1 M NaOH 79
NaCoFe2(PO4)3 C2/c 55 ∼3.0 80 0.75 108 1.55 0.1 M KOH 80
NaCo4(PO4)3 P21/n N.A. N.A. 520 121 Phosphate Buffer 98
Na2FePO4F Pbcn 100 ∼3.05 99 0.89 0.77 119 1.69 570 130 0.1 M KOH 83
Na2MnPO4F P21/c ∼60 ∼3.7 100 0.91 0.84 137 1.72 250 0.1 M KOH 83
Na2CoPO4F Pbcn 55 4.2 84 0.90 0.83 93 1.61 500 110 0.1 M KOH 83 and 84
0.85 0.79 3.7 105 1.59 420 0.1 M NaOH
β-Na2MnP2O7 P[1 with combining macron] 80 ∼3.6 101 0.87 0.84 ∼3.5 118 1.56 470 64 0.1 M KOH 76
α-Na2MnP2O7 P1 N.A. 0.87 0.83 4.0 131 1.58 145
Na2CoP2O7 Pna21 ∼80 ∼3.0 102 0.78 115 N.A. 0.1 M NaOH 73
0.86 0.74 96 1.50 51 1 M NaOH
0.87 3.50 82 N.A. 0.1 M KOH 75
N.A. 339 79 1 M KOH
P[1 with combining macron] ∼80 ∼4.3 103 0.87 3.90 78 N.A. 0.1 M KOH
N.A. 374 64 1 M KOH
K-cation based
KFePO4 P21/c ∼25 ∼2.8 85 1.01 3.84 104 N.A. 0.1 M KOH 96
KCo(PO3)3 P[6 with combining macron]c2 N.A. 104 ∼0.81 85 1.52 156 0.1 M NaOH 79
K2CoP2O7 P42/mnm Inactive 87 0.87 0.78 3.84 114 1.48 370 106 0.1 M KOH 87
0.95 0.82 109 N.A. Phosphate Buffer
K2Co(PO3)4 Cc N.A. 0.83 0.87 59 1.44 1.57 80 0.1 M NaOH 88



image file: d3qm01088k-f2.tif
Fig. 2 OER activity of LiCoPO4, a commonly used cathode material for lithium-ion battery devices under (a) neutral conditions and (b) post-mortem analysis of cycled catalysts in 0.1 M KPi (pH 7) and 0.1 M KOH (pH 13). HRTEM images (left) and FFTs (right) of the LiCoPO4 surface showing the evolution of the amorphous region, i.e. change in surface morphology at alkaline pH, which is distinguished from crystalline regions with the white dashed lines. Reproduced from Lee et al.,60 with permission from the American Chemical Society. The OER activity of sodium ion battery's cobalt phosphate-based cathode (Na2CoP2O7 and NaCoPO4) vs. lithium-ion battery's cobalt phosphate-based cathode (Li2CoP2O7 and LiCoPO4). (c) CV curves of 1st and 100th cycles at neutral pH from 0.7 to 1.5 V (versus NHE), with the thermodynamic potential for water oxidation at 0.816 V (versus NHE). (d) Bulk electrolysis of Na2CoP2O7, NaCoPO4, Li2CoP2O7 (inset), and LiCoPO4 (inset) performed under a constant potential of 1.4 V (versus NHE). Reproduced from Kim et al.,63 with permission from Springer Nature.

Later, Manthiram group characterized and compared the OER activity in layered oxide battery materials containing various compositions of Mn, Co and Ni in LiTMO2. They observed that Ni-rich compositions exhibited the highest OER activity, while Mn-rich compounds were found to suppress the OER activity.61 The root cause is related to the singly occupied eg orbital along with an oxidation state of +3 for the transition metal as well as surface atomic arrangements, which is explained in detail in the following sections.61,62 Certainly, fresh or spent battery insertion materials can be put on anvil to design economic electrocatalysts showing ORRs, OERs, or bifunctional activities, suitable for metal–air batteries. In this spirit, the recent advances in the exploration of various phosphate-based battery insertion cathodes are elaborated in the following sections.

4. Phosphate-based polyanionic insertion materials as catalysts

Before enlisting the significant phosphate-based battery materials exhibiting electrocatalysis until now, at first, consideration should be given to how phosphates gained interest in oxygen electrocatalysis. The story begins with in situ-generated Co–Pi (phosphate in an approximately 1/2 ratio with cobalt), reported by Kanan and Nocera in 2008 exhibiting remarkable OER activity in earth-abundant catalysts at neutral pH under ambient conditions.64 Later Ren et al. developed N, P codoped reduced graphene-oxide-aerogel-supported (rGOA-supported) Co2P2O7 (CoPi) particles (CoPi/NPGA) via a simple hydrothermal route followed by pyrolysis route delivering excellent bifunctional activities in these Co–Pi materials.65 Taking advantage of cobalt phosphate, Baby et al. modified the catalytic activity of Co by a divalent cation (Zn) substitution to realize robust bifunctional activity in ZnCo2(PO4)2 and ZnCoP2O7 (pyro)phosphates.66 The bifunctional oxygen electrocatalytic performance of Zn-substituted Co phosphate and pyrophosphate is compared with non-substituted Co-phosphate Co2P2O7/NPGA and Co3(PO4)2/NC (Table 1). During the ORR activity, the onset potential (Eonset) for Co2P2O7/NPGA, Co3(PO4)2/NC, ZnCo2(PO4)2 and ZnCoP2O7 was found to be 0.91, 0.96, 0.87 and 0.86 V (vs. RHE), respectively, with the increasing order of half-wave potential (E1/2): ZnCoP2O7 (0.72 V) < ZnCo2(PO4)2 (0.73 V) < Co2P2O7/NPGA (0.80 V) < Co3(PO4)2/NC (0.83 V) in 0.1 M KOH. The LSV curve showed a higher saturation current for Zn-substituted Co catalysts, which is comparable to that of the Pt/C benchmark, which yielded 5.98 mA cm−2 exchange current density. In case of Co2P2O7/NPGA, Co3(PO4)2/NC, the untreated catalysts with no N and P doping, exhibited inferior ORR activity. For the OER activity, at a current density of 10 mA cm−2, the overpotential from 1.23 V (vs. RHE) for these catalysts was reported to be 340 mV, 296 mV, 450, and 460 mV for Co2P2O7/NPGA, Co3(PO4)2/NC, ZnCo2(PO4)2 and ZnCoP2O7 respectively. It is worth mentioning that without any further addition of N or P or reduced graphene oxide into the structure, ZnCo2(PO4)2 and ZnCoP2O7 showed comparable ORR and OER activities. Consequently, it can be claimed that Zn-substituted Co-phosphates increased the catalytic activity of untreated Co-phosphates.65,66

4.1. Phosphate (PO4)

As mentioned above, besides LiFePO4, LiCoPO4 and LiMnPO4 olivine structures containing only one transition metal, the OER activity in olivines containing mixed divalent cations has also been investigated by Ma et al. in 2016 and Wu et al. in 2020.67,68 Apart from the partial substitution of Co/Ni or Fe in the host structure, Gershinsky et al. achieved lithium sub-stoichiometric Li0.7Co0.75Fe0.25PO4 compound by the direct synthesis having an olivine structure and identified excellent electrochemical water oxidation.69 These findings coincide with those of Cui group's work demonstrating increased OER activity in electrochemically extracted olivine-type lithium transition metal phosphates.70 For LIB cathodes Li0.7Co0.75Fe0.25PO4/rGO, LiCo0.60Fe0.40PO4, LiNiPO4 and LiNi0.75Fe0.25PO4/rGO showing a specific discharge capacity ≥120 mA h g−1, the OER activity is presented in Table 1, showing a clear influence of transition metals on water oxidation properties into the host LiTMPO4 olivine-type structure.68,69,71 The onset potential for the OER activity of LiCoPO4, Li0.7Co0.75Fe0.25PO4/rGO and LiCo0.60Fe0.40PO4 was found to be 1.60, 1.61 and 1.47 V (vs. RHE), respectively, in 0.1 M KOH, while for LiNiPO4 and LiNi0.75Fe0.25PO4/rGO, it showed Eonset of 1.65 and 1.45 V (vs. RHE), respectively (Table 1).67,69 In all these cases, the activities of the end member/undoped LiTMPO4 were enhanced because of the synergistic effect between Fe and Co or Ni.67,68

4.2. Pyrophosphate (P2O7)

Kim et al. comprehensively evaluated the influence of PO4vs. P2O7 units on OER activity using PO4-chemistry in one formula unit (f.u.) in its crystal structure. For this study, LiCoPO4, Li2CoP2O7, NaCoPO4 and Na2CoP2O7 cathode materials were taken into consideration, showing a specific discharge capacity ≤90 mA h g−1 for LIBs/SIBs. The four chosen Co-based phosphate catalysts exhibited a range of catalytic activity, championed by Na2CoP2O7 exhibiting the highest exchange current during water oxidation. It clearly explained the crucial role of metal coordination, which serves as a platform to fully comprehend the catalytic activity of these materials under neutral conditions. Referring to Fig. 3a, the crystal structures of LiCoPO4, Li2CoP2O7, NaCoPO4 and Na2CoP2O7 revealed the presence of different networks between Co polyhedra and various Co coordination, resulting in different catalytic activities with the increasing trend of LiCoPO4 < Li2CoP2O7 < NaCoPO4 < Na2CoP2O7 (Fig. 2c and d). The distorted local Co–O4 geometry in Na2CoP2O7 displayed the highest OER activity compared to others, where the rotation of the flexible pyrophosphate ligands allowed Co–O5i.e. an additional Co–O bond at the surface due to water adsorption and further water oxidation catalysis at neutral pH.63,72 Furthermore, Gond et al. evaluated the increased specific OER and ORR activities for Na2CoP2O7 denting this material as a bifunctional catalyst in an alkaline pH environment.73 They also investigated the importance of interplay between the TM and P2O7 units, showing polymorphism vs. bifunctional electrocatalysis in Na2CoP2O7 and Na2MnP2O7. Pyrophosphates are likely to exhibit polymorphism, providing a variety of crystal structures acting as potential cathodes for LIBs and SIBs.74 Later, realizing these materials for ORRs and OERs, different polymorphs of Na2CoP2O7, such as orthorhombic phase o-Na2CoP2O7 (Pna21) and triclinic phase t-Na2CoP2O7 (P[1 with combining macron]) and triclinic polymorphs of Na2MnP2O7 like β-Na2MnP2O7 (P[1 with combining macron]) and α-Na2MnP2O7 (P1) have been synthesized at low and high temperatures, respectively, as shown in Fig. 3b.75,76 Acting as insertion materials for the sodium-ion batteries, o-Na2CoP2O7, t-Na2CoP2O7, and β-Na2MnP2O7 delivered a specific discharge capacity of ∼80 mA h g−1 at 3.0 V, 4.3 V, and 3.6 V, respectively. In the case of Na2CoP2O7, the trend for ORR exchange current density was o-Na2CoP2O7 < t-Na2CoP2O7, while for OER the trend was opposite; t-Na2CoP2O7 < o-Na2CoP2O7. However, in case of Na2MnP2O7, the bifunctional activity of β-Na2MnP2O7 was found to be superior to that of α-Na2MnP2O7 (the data are summarised in Table 1).
image file: d3qm01088k-f3.tif
Fig. 3 Crystal structure of selected phosphate-based polyanion insertion materials with the mobile Na/Li atom (yellow/green) in the 3D frameworks. (a) Cobalt crystal field of Co-phosphate catalysts: NaCoPO4, LiCoPO4, Na2CoP2O7, and Li2CoP2O7 showing their local environment around the Co-subunit (blue for Co–O4 or Co–O6, grey for PO4 unit) and the corresponding spin state of Co atoms. The presented image is reproduced from Kim et al.,63 with permission from Springer Nature. (b) Structural comparison of two polymorphs of Na2CoP2O7 and Na2MnP2O7 w.r.t. temperature, where Co–O, Mn–O, and P–O subunits are shown in blue, brown and grey, respectively. The bridging of pyrophosphate (P2O7) units and Co–O octahedra in P[1 with combining macron] Na2CoP2O7 are edge shared and corner shared, while in Pna21 Na2CoP2O7, the P2O7 units and Co–O tetrahedra are only corner shared. Reproduced from Gond et al.75 with permission from the American Chemical Society. In Na2MnP2O7, both polymorphs P1 and P[1 with combining macron] have only corner shared Mn-O6 and P2O7 subunits. Reproduced from Gond et al.76 with permission from the Royal Society of Chemistry. (c) Na2FePO4F (pink for Fe–O4F2 and green for PO4 unit), Na2CoPO4F (blue for Co–O4F2 and grey for PO4 unit), and Na2MnPO4F (olive green for Mn–O4F2 and brown for PO4 unit) crystallize in orthorhombic, orthorhombic, and monoclinic systems respectively, where all the transition metal octahedra are corner shared to PO4 tetrahedra. Reproduced from Sharma et al.83 with permission from the Royal Society of Chemistry.

4.3. Metaphosphate (P3O9)

Pursuing the story further, Gond et al. for the first time reported excellent OER activity in NaCo(PO3)3, which is a 3.2 V cathode material for SIBs.77,78 This framework is built from the interlinked PO4 tetrahedra connecting the CoO6 octahedra. The Co-analogue showed superior electrocatalytic activity to the Fe and Mn analogues in metaphosphate NaTM(PO3)3. Recently, Murugesan et al. have reported the bifunctional activity of NaCo(PO3)3 in 0.1 M NaOH and extended it to KCo(PO3)3 witnessing metaphosphate as a bifunctional catalyst for metal air-battery applications.79 The LSV curves showed ORR onset potentials of 0.83 V and ∼0.81 V for NaCo(PO3)3 and KCo(PO3)3, respectively. However, both materials delivered similar OER activity with an onset potential of 1.52 V.

4.4. Alluaudite [(PO4)3]

The study further extended to the phosphate-based alluaudite class of materials in particular NaCoFe2(PO4)3. This material works as a ∼3.0 V cathode without any material optimization delivering a discharge capacity of 60 mA h g−1 at C/20.80 Exploiting the Co species, it showed bifunctional activity with onset potentials of 0.75 V and 1.55 V (vs. RHE) for ORR and OER processes respectively.

4.5. Fluorophosphates (PO4F)

Another interesting class of insertion materials comprises fluorophosphates having a general formula Na2TMPO4F (TM = Fe, Mn, and Co; crystal structure is represented in Fig. 2c), which tend to exhibit high working potentials concerning their phosphate analogues.81 In principle, the participation of fluorine in a Na2TMPO4F material possesses a larger ionicity of the TM–F bond than that of the TM–O one (occurring in PO4 analogues). Due to the higher electronegativity of the F atom, fluorophosphates show a higher redox potential for the TM redox activity.82 The catalytic performance of this high-voltage battery cathode materials was unknown before our group explored the oxygen bifunctional electrocatalytic capabilities of fluorophosphate containing Fe, Mn, and Co transition metals (Table 1).83 Among them, Na2CoPO4F was found to deliver excellent bifunctional properties exhibiting onset potentials of 0.854 V and 1.591 V (vs. RHE), for ORR and OER activities respectively in 0.1 M NaOH.84

In parallel, interest in studying the electrocatalytic properties of KIB cathode materials has recently grown. Lack of adequate electrode materials that can reversibly intercalate K+ with high capacity, nominal voltage, rate kinetics, and stable cycle life makes this field challenging than the Li/Na-analogues.85 One reason lies in their relative cationic size (0.76 Å for Li+ < 1.02 Å for Na+ < 1.38 Å for K+). Considering one of the simplest phosphate-based cathode materials, KFePO4 was found to yield a reversible discharge capacity of ∼47 mA h g−1 in the amorphous form. However, its crystalline form showed a sloping voltage profile with very low specific capacity originating from electric double layer capacitance instead of any efficient intercalation.86 Murugesan et al. reported the ORR activity of monoclinic KFePO4 with an onset potential of 1.01 V (Pt/C Eonset = 0.79 V vs. RHE). Its pyrophosphate derivative, K2CoP2O7, also worked as a potential electrocatalyst showing excellent bifunctional properties with onset potentials of 0.87 V and 1.48 V for ORR and OER activities, respectively (Table 1).87 Similarly, K2Co(PO3)4 exhibited bifunctional properties with unavailable charge–discharge intercalation properties of K-ions, where onset potentials of 0.83 V and 1.44 V for ORR and OER activity, respectively, were obtained in 0.1 M NaOH.88

4.6. Influence of transition metals (TM = Co, Ni, Mn and Fe)

In Li+ intercalating cathode materials, the operating voltages normally increase with the higher amount of d electrons (up to Ni). In 2004, Ceder group estimated the working potential for LiTMPO4 (TM = Fe, Mn, and Co), which was consistent with the previous experimentally measured operating potential.105 The obtained potentials for the redox activity of TM based on both the methods merged to one simple conclusion explained with ionic molecular-orbital theory (MOT). It is based on the band structure of the material, which splits 3d-orbitals into three lower t2g and two upper image file: d3qm01088k-t11.tif bands in case of octahedral oxygen coordination. Typically, for the early TM in LiTMPO4, the valence bands are mainly well separated from metal-deprived bands with oxygen-p character, while later as the d electron increases, the t2g and image file: d3qm01088k-t12.tif drop in energy and the former mix with the oxygen-p bands. Until Co, the M–O bond length noticeably decreases, due to the increase in nuclear charge on the metal, leading to an effectively smaller metal ion. However, afterwards for Ni, the M–O bond length increases due to the filling of the antibonding image file: d3qm01088k-t13.tif bands. Towards the end (for Cu and Zn), because of the increase in the nuclear charge of the metal ion, the metal-d band character moves to a lower energy and virtually intermixed with the oxygen-p bands, resulting in higher average intercalation voltage that might be limited by unstable electrolytes at that potential.106

Using the same MOT approach, the designing and discovery of new cost-effective highly active catalysts for electrochemical energy conversion and storage can be estimated. Motivated by Nørskov, Shao-horn group proposed a d-band theory activity descriptor for metal surfaces to explain the mechanisms for oxygen electrocatalysis in the transition metal oxides in perovskite-type structures. This landmark contribution accurately defined the transition metal oxides such as ABO3 perovskites composed of rare and alkaline earth (A) and 3-d TM cations (B) exhibiting ORR/OER activities comparable to the state-of-the-art catalysts IrO2, RuO2 and Pt/C.9,62,107,108 They developed the descriptor for ORRs, revealing M-shaped relationship with the increase in d-electron (for B site TM cations), with the maximum activity for d4 and d7, resulting in a volcano shape plot as a function of eg-filling of TM cations at B site in the case of perovskites. A similar volcano shape trend was achieved for OERs with eg-filling vs. potential at a particular current density.62,108 According to them, the identification of eg filling (of antibonding states) of surface transition metal cations is a universal activity descriptor for the oxides. As per the catalysis surface science based on the reaction mechanism (alkaline pH in the introduction section) for TM in the octahedra geometry, the σ-bonding eg orbital has a stronger overlap with the oxygen-related adsorbate, in particular, the OOH* intermediates than does the π-bonding t2g orbital. The reported eg filling close to unity experimentally facilitates the highest ORR and OER activity, due to more efficient electron transfer between surface cation and adsorbed reaction intermediates.62

Inspired by this, the oxygen electrocatalytic activity in selected phosphate-based catalysts containing different transition metals, where simply the number of electrons in the 3d orbital and eg (antibonding σ*-orbital; TMO6 and bonding σ-orbital; TMO4) orbital is represented during ORRs and OERs, respectively, is shown in Fig. 4 and the related data are tabulated in Table 2. Fig. 4a and b are recreated from Suntivich et al. work for ORR and OER activities, respectively,62,107,108 providing deep insights into the obtained activities in LiCo0.60Fe0.40PO4,68 LiNi0.75Fe0.25PO4/rGO,67 NaCoPO4,63 NaFe(PO3)3,78 NaCo(PO3)3,78 NaCo4(PO4)3,98 Na2FePO4F,83 Na2CoPO4F,83 β-Na2MnP2O7,76o-Na2CoP2O7,75t-Na2CoP2O7,75 KCo(PO3)3,79 K2CoP2O7,87 and K2Co(PO3)488 materials. Respective crystal structure, valence or oxidation state of transition metal, causing the number of electrons in 3d-orbitals, its spin state, and eg filling state (antibonding σ*-orbital for octahedra and bonding σ-orbital for tetrahedra geometry) are tabulated against potential in V (vs. RHE) at −3.0 mA cm−2 and 10.0 mA cm−2 for ORR and OER activities, respectively. Among all the known phosphate-based catalysts displaying intercalation chemistry, most of the articles are based on Co-based phosphate materials for ORR and OER electrocatalysis (Fig. 4 shows the predominant Co-based catalysts in the plot).


image file: d3qm01088k-f4.tif
Fig. 4 ORR and OER activities of phosphate-based polyanionic intercalation materials in alkaline solutions. (a) ORR activity: potential at a current density of 3 mA cm−2 of the selected phosphate-based catalysts with the respective transition metal centre d-electron, following the M-shaped relationship, as reported for perovskite oxides from Shao-horn et al.108 (b) OER activity: potential at a current density of 10 mA cm−2 of phosphate-based electrocatalysts with the respective eg electrons in d-orbital of transition metal responsible for the activity, the dashed volcano lines are shown only for guidance from the literature Shao-horn et al. study for perovskite oxides.9,62 Data symbols are as follows: brown triangles for Mn, blue triangles for Co, red triangles for mixed transition metal, and green open circles for Fe-based phosphate-containing electrocatalysts. The data presented are replotted and were originally reported as: LiCo0.60Fe0.40PO4,68 LiNi0.75Fe0.25PO4/rGO,67 NaCoPO4,63 NaFe(PO3)3,78 NaCo(PO3)3,78 NaCo4(PO4)3,98 Na2FePO4F,83 Na2CoPO4F,83 β-Na2MnP2O7,76o-Na2CoP2O7,75t-Na2CoP2O7,75 KCo(PO3)3,79 K2CoP2O7.87 The electrocatalytic activity of the aforementioned catalysts as well as the spin state of the catalytic transition metal centre are detailed in Table 2.
Table 2 Summary of the literature studies on the spin states (H.S.; high spin, L.S.; low spin, I.S.; intermediate spin) of the selected phosphate-based electrocatalysts based on ORR and OER performance
Materials Crystal structure (space group) Valence electrons in 3d-orbital Spin state eg filling assignment Potential, V@−3 mA cm−2 (ORR) Potential, V@10 mA cm−2 (OER) Ref.
LiCo0.60Fe0.40PO4 Pnma Co+2/+3 d7 and d6 H.S./I.S. t52g e2g (d7) and t4.52g e1.5g (d6) N.A. 1.57 69 and 109
Fe+2 d6 L.S. t62g e0g (d6)
LiNi0.75Fe0.25PO4/rGO Pnma Ni+2 d8 L.S. t62g e2g (d8) N.A. 1.52 67
Fe+2/+3 d6 and d5 H.S. t42g e2g (d6) and t32g e2g (d5)
NaCoPO4 Pnma Co+2 d7 H.S. t52g e2g (d7) 0.69 1.61 63
NaFe(PO3)3 Pa[3 with combining macron] Fe+2 d6 H.S. t42g e2g (d6) Inactive 1.78 97
NaCo(PO3)3 Pa[3 with combining macron] Co+2 d7 H.S. t52g e2g (d7) 0.72 1.57 78
NaCo4(PO4)3 P21/n Co+2 d7 H.S. t52g e2g (d7) 1.75 98
Na2FePO4F Pbcn Fe+2 d6 H.S. t42g e2g (d6) 0.67 1.80 83
Na2CoPO4F Pbcn Co+2 d7 H.S. t52g e2g (d7) 0.64 1.73 84
β-Na2MnP2O7 P[1 with combining macron] Mn+2 d5 H.S. t32g e2g (d5) 0.68 1.70 76
t-Na2CoP2O7 P[1 with combining macron] Co+2 d7 H.S. t52g e2g (d7) 0.71 1.60 75
o-Na2CoP2O7 Pna21 Co+2 d7 H.S. e4 t23 (d7) 0.68 1.57 75
KCo(PO3)3 P[6 with combining macron]c2 Co+2 d7 H.S. t52g e2g (d7) 0.66 79
K2CoP2O7 P42/mnm Co+2 d7 H.S. e4 t23 (d7) 0.70 1.60 87
K2Co(PO3)4 Cc Co+2 d7 H.S. t52g e2g (d7) 0.72 1.57 88


For the ORR activity, the catalyst exhibiting a current density ≥ −3.0 mA cm−2 is considered and shown in Fig. 4a. Such catalysts are enlisted with the potential (written next to the material) at a current density of −3.0 mA cm−2vs. the respective 3d-electron in TM. Based on this potential, onset potential and Tafel slope from Tables 1 and 2, the decrease in ORR activity trend is shown here with potential in V (vs. RHE), at −3.0 mA cm−2 in brackets: K2Co(PO3)4 (0.72 V) ≥ NaCo(PO3)3 (0.72 V) > t-Na2CoP2O7 (0.71 V) > K2CoP2O7 (0.70 V) > NaCoPO4 (0.69 V) > o-Na2CoP2O7 (0.68 V) ≥ β-Na2MnP2O7 (0.68 V) > Na2FePO4F (0.67 V) > KCo(PO3)3 (0.66 V) > Na2CoPO4F (0.64 V). The observed trend clearly shows the validation of the M-shaped relationship from the Suntivich et al. descriptor for ORRs, implying that the increase in d-electron enhances the activity, with the maximum ORR activity for d4 (the present review study deployed of d4 transition metals for ORR) and d7. The same descriptor of eg-filling stands well to explain the trend for OER activity in Mn, Fe, Co and Ni containing phosphate-based intercalation cathodes studied so far as catalysts for oxygen evolution chemistry in alkaline pH. Fig. 4b represents such cathodes whose potential (E10 in V) at 10.0 mA cm−2 is shown vs. the number of electrons in the eg orbital except for o-Na2CoP2O7 and K2CoP2O7, which possess TM–O4, thus it represents e (bonding σ-orbital, will be overfilled >2 for Co2+). The spatial mismatch between the oxygen ligand of TM–O4 and the d-orbitals provided by transition-metal cations on the tetrahedral sites suggests weak interaction with molecular oxygen and oxygenated species.

The declining trend of OER activity is likely to be shown here, representing potential in V (vs. RHE), at 10.0 mA cm−2 in brackets: LiNi0.75Fe0.25PO4/rGO (1.52 V) > LiCo0.60Fe0.40PO4 (1.57 V) ≥ NaCo(PO3)3 (1.57 V) > o-Na2CoP2O7 (1.57 V) > K2Co(PO3)4 (1.57 V) > K2CoP2O7 (1.60 V) > t-Na2CoP2O7 (1.60 V) > NaCoPO4 (1.61 V) > β-Na2MnP2O7 (1.70 V) > Na2CoPO4F (1.73 V) > NaCo4(PO4)3 (1.75 V) > NaFe(PO3)3 (1.78 V) > Na2FePO4F (1.80 V). In summary, for ORR as well as OER processes, the Co-containing phosphates are found to be superior due to the appropriate number (d7 and eg-filling close to ≥1) of electrons in the eg orbital that causes efficient bonding of O–Co–OOH* oxygen-related intermediates. In particular, LiCo0.60Fe0.40PO4 showed eg-filling close to 1.2 (in Co), laid on the imaginary volcano plot in Fig. 4b. Moreover, the Fe-doped Co/Ni-phosphates (LiCo0.60Fe0.40PO4 and LiNi0.75Fe0.25PO4/rGO) exhibited dramatically improved OER activity compared to the pure phases, due to the synergistic coupling effects for the favorable electronic structure of the surface Co, improving the rate-limiting step which is the formation of OOH* derived from O* during the OER.67,68 The metal cations and ligands (here phosphates) also affect the kinetics of intermediate adsorption/desorption and overall physicochemical processes governing catalytic activities as discussed below.

4.7. Influence of alkali cations

Besides establishing a correlation between 3d-electrons and eg occupancy with the oxygen chemistry as an activity descriptor, metal-oxygen covalency is also a very important activity descriptor. Based on the study conducted by Wei et al. for spinels with Co or Ni as active sites (e.g. Co3O4, ZnCo2O4, and NiCo2O4), because of different cations at the tetrahedra site, the ORR and OER activities in these materials deviated from the eg-filling theory, occurring due to the more covalent metal-oxygen bond nature at the octahedral site.110 Further, sticking to one particular cation (La), in LaTMO3 oxide (where TM = Mn, Co and Ni), the potential to obtain 25 μAcm−2 during the ORR follows the trend LaNiO3 > LaCoO3 > LaMnO3 with values of 908 (±8) mV, 847 (±3) mV and 834 (±24) mV (versus RHE) respectively. It witnesses superior ORR activity in Ni not only due to eg-filling theory, but also because of the higher covalent nature of the TM–O bond due to the higher electronegativity of the metal cation.108 Similarly, changing cations (M) in MxTM(PO4)y will greatly influence the overpotential linked to the electronegativity of the TM–O bond. Interestingly, a comparison between the ORR and OER activities reveals such influence supporting the truth of the covalent TM–O bond regardless of the eg-filling theory, as shown in Fig. 5 and Table 2. Considering Na2CoP2O7 and K2CoP2O7 having Pna21 and P42/mnm space groups, respectively, they possess Co–O4 tetrahedra in both cases with the only difference of metal cations (Na and K). It establishes evidence of the Co–O bond covalency when the electronegativity of neighbouring metal cation differs.111 A careful examination of Fig. 5 showing LSV curves of different Co-containing phosphate-based catalysts will reflect two types of Na2CoP2O7 of Pna21 named o-Na2CoP2O7 and o-Na2CoP2O7*, where the one labelled with * has improved oxygen electrocatalytic activity due to the synergistic effect between carbon and o-Na2CoP2O7, obtained after the pre-treatment of the catalyst with Super P. Still, it can be noted that K2CoP2O7 performs better than Na2CoP2O7 for total oxygen electrocatalysis (ORR and OER activities in Fig. 5a and b, respectively). In particular, comparing them in 0.1 M KOH, the overall oxygen electrocatalysis is seen to be better in K2CoP2O7 than in Na2CoP2O7. In support of this statement, comparing the ORR activity in 0.1 M KOH for both catalysts exhibits a similar onset potential (0.87 V vs RHE), but K2CoP2O7 ended up displaying a large limited current density (−4.23 mA cm−2) than Na2CoP2O7 (−3.53 mA cm−2) at 0.3 V vs RHE (Fig. 5a). It demonstrates a more efficient mass transfer among K2CoP2O7 due to the presence of K-cations (electronegativity trend is K < Na) in the electrocatalyst. A similar superior influence of K-cation is observed for the OER activity, where the achieved current density at 1.7 V (vs. RHE) was found to be just 2.79 mA cm−2 for Na2CoP2O7, while a ∼6-fold higher current density (ca. 16.82 mA cm−2) was achieved for K2CoP2O7 (Fig. 5b). It clearly witnesses the positive impact of TM–O bond covalency happening due to less electronegative alkali metal cations in alliance with the transition metals.
image file: d3qm01088k-f5.tif
Fig. 5 Linear sweeping voltammograms (LSVs) for different Co-containing phosphate-based-polyanionic intercalation materials in alkaline aqueous electrolytes (pH 13–13.7) performed at a scan rate of 10 mV s−1 with a rotation of 1600 rpm. (a) Oxygen reduction activity compared with Vulcan carbon XC (purple dashed line) and Pt/C (Pt is 20 wt%, grey dashed line) in O2-saturated 0.1 M KOH. (b) Oxygen evolution activity compared with state-of-the-art RuO2 catalyst in Ar-saturated 0.1 M and 1 M KOH. Presented data are replotted together for a better comparison using the originally reported data as follows: NaCoPO4,63 NaCo(PO3)3,78 NaCoFe2(PO4)3,80 Na2CoPO4F,83o-Na2CoP2O7,63o-Na2CoP2O7*,75t-Na2CoP2O7,75 and K2CoP2O7.87

4.8. Influence of phosphate units as ligands

Continuing the above discussion, it is not only the type of cations that influences the metal covalency affecting the affinity towards the oxygen intermediates responsible for oxygen electrocatalysis, but the neighbouring ligands also determine the spatial arrangements of these intermediates binding temporarily to TM–O and ease of reaction. Structural conformations of PO4-ligands highly influence the activity. Explicitly considering the case of simple PO4, P2O7 and PO4F, to complex multiple units of interlinked PO4 as in metaphosphates or alluaudite phosphates, the bifunctional oxygen electrocatalysis was found to be different. It is described in detail for NaCoPO4,63 NaCo(PO3)3,78 NaCoFe2(PO4)3,80 Na2CoPO4F83 and o-Na2CoP2O763 in the present section. Their LSV profiles for ORRs and OERs reactions are presented in Fig. 5.

LSVs from the ORR activity conducted in 0.1 M KOH showed the highest limited current density for NaCoFe2(PO4)3 with a value of −4.32 mA cm−2 at 0.3 V (vs RHE) (Fig. 5a) close to −4.48 mA cm−2 of commercial Pt/C, followed by Na2CoPO4F, o-Na2CoP2O7 and NaCoPO4 which achieved −3.29, −2.12, and −1.77 mA cm−2, respectively.63,80,83o-Na2CoP2O7* and other polymorphs labelled as t-Na2CoP2O7 were treated with Super P carbon, where the as-synthesised phases were ball milled with conductive carbon exhibiting enhanced activity when compared to the untreated o-Na2CoP2O7 (Fig. 5a).63 In particular, ignoring the bimetallic NaCoFe2(PO4)3 electrocatalyst which showed superior limited current density, the effect might be an outcome of the synergistic effect of Fe and Co as reported for LiCo0.60Fe0.40PO4.69 Thus, based on the ORR activity trend of Na2CoPO4F > o-Na2CoP2O7 > NaCoPO4, in Na2CoPO4F, the presence of the F ligand in PO4-based materials forms F2–Co–O4 octahedra, as compared with NaCoPO4, which has CoO6 units (Fig. 3). The active centre with CoO4F2 showed superior activity to CoO6, which could be due to the presence of most electronegative F atoms in the vicinity of the Co metal, that is going to interact with the oxygen intermediates OOH* (as explained in the introduction section).83 However, metal coordination varies in the case of P2O7vs. PO4, where CoO4 is corner shared with the P2O7 unit, while CoO6 is edge shared with the PO4-unit. In this case, P2O7 offers more degrees of spatial rearrangements when the oxygen intermediate forms a temporary bond with an active metal centre. Nevertheless, the eg-filling theory is more reasonable when it comes to the ORR activity while comparing two polymorphs of Na2CoP2O7. Here, regardless of CoO4 corner shared with the P2O7 unit in o-Na2CoP2O7* (Fig. 5a) possessing more possibility of spatial rearrangements, CoO6 which is edge shared with the P2O7 unit in t-Na2CoP2O7 showed a higher exchange current density.75

Furthermore, from Fig. 5b, the OER activity results conducted at pH ∼ 13 showed current densities of 2.79, 3.20, 4.87, 7.09 and 8.03 mA cm−2 (at 1.7 V vs. RHE) for o-Na2CoP2O7, NaCoPO4, NaCoFe2(PO4)3, Na2CoPO4F, and commercial RuO2, respectively. In principle, based on Kim et al.'s report, the onset potential for the OER activity and exchange current density for Na2CoP2O7 is superior to that of NaCoPO4 at neutral pH or alkaline pH. Our results also overlapped with the reported findings proving favourable binding of water molecules in Na2CoP2O7 as compared to NaCoPO4 due to flexible local cobalt coordination in Na2CoP2O7.63 However, in 0.1 KOH, the current density at 1.7 V and onset potentials are quite comparable as mentioned above, while the onset potential for o-Na2CoP2O7 outperforms at a more basic alkaline pH. Due to the compositional mixing of Co and Fe in NaCoFe2(PO4)3 alluaudite-type structure, the bulk electronic structure of CoO6 (metal 3d and O 2p states greatly influenced) lowered surface adsorption energetics and contributed a higher current density at 1.7 V (vs. RHE) than o-Na2CoP2O7 and NaCoPO4.9,111 Similar to the superior ORR activity in CoO4F2 type of TM active centres, Na2CoPO4F exhibits a current density (7.09 mA cm−2) very similar to commercial RuO2 (8.03 mA cm−2) at pH 13. The structure of Na2CoPO4F, which is more electrophilic in nature due to the F-atom, allows for the minimum binding of water molecules and promotes greater activity than that of o-Na2CoP2O7, NaCoPO4 and NaCoFe2(PO4)3, demonstrating the impact of the ligands on electrocatalysis. Comparing the OER activity in 1 M KOH, NaCo(PO3)3 showed an exceptionally high current density of 89.72 mA cm−2 at 1.7 V (vs. RHE) that is way higher than that of RuO2 (55.43 mA cm−2). The explanation behind the superior activity in NaCo(PO3)3 again suggests the importance of local Co-coordination and preference of corner-sharing to the neighbouring PO4-units. For clarity, it is important to mention that the current densities at 1.7 V (vs. RHE) for t-Na2CoP2O7 and o-Na2CoP2O7 polymorphs are 49.81 and 69.43 mA cm−2, respectively. Thus, for the material possessing CoO6 as an active metal centre, the trend for OER activity is NaCo(PO3)3 > t-Na2CoP2O7, as in both the cases the active metal centre eg-orbital directly interacts with incoming intermediates, but the main difference is the edge-shared and corner-shared octahedra with the pyrophosphate unit [(P2O7)4−]. The isolated CoO6 active centre in NaCo(PO3)3 facilitates efficient surface binding properties and allows all possible spatial rearrangement required during electrocatalysis. Due to the edge-shared octahedra in t-Na2CoP2O7, despite the direct interaction between the eg-orbital and the incoming intermediates, the system undergoes some stress. o-Na2CoP2O7* showed superior OER activity to Co–O4 as the catalytic centre (Fig. 5b).

5. Metal–air batteries using phosphate-based materials

Inspired by the demonstration of low-cost bifunctional electrocatalysts using PO4-based battery insertion materials, they have been implemented in (hybrid) metal–air batteries.112,113 Real-time pictures, schematic illustrations and galvanostatic cycling results of two such devices are shown in Fig. 6. Promising Na+-ion intercalation into the host structures of Na0.5Co0.5Mn0.5O2 and NaCoPO4F promoted their potential towards sea water batteries and hybrid Na–air batteries, respectively. K2Co(PO3)4, another potential cathode material for K-ion batteries, could also be implemented as air-cathode for hybrid Na–air battery applications. The application of bifunctional electrocatalysts for rechargeable Zn–air batteries (ZABs) is also widely known, where Co2P2O7/CoPi, ZnCoP2O7, t-Na2CoP2O7, o-Na2CoP2O7, and K2CoP2O7 have been used as cathodes for the fabrication of such storage devices.
image file: d3qm01088k-f6.tif
Fig. 6 Rechargeable secondary devices made of selected phosphate-based materials. Hybrid Na–air battery: (a) digital image of a hybrid Na–air battery illuminating a green LED with the Na2CoPO4F air cathode, containing a pouch cell (made of sodium metal dipped in an organic non-aqueous electrolyte and a solid membrane, NASICON, which is exposed outside) dipped in a beaker-type cell filled with aqueous electrolytes. (b) Schematic illustration of a hybrid Na–air battery describing the processes involved at the air electrode, cathode (Na2CoPO4F84) during charge and discharge. (c) Corresponding hybrid Na–air battery performance at different current rates. Reproduced from Sharma et al.84 with permission from the American Chemical Society. Zn–air battery: (d) image of the Zn–air battery configuration with the catalyst, anode, and simple set-up using a 20 mL plastic box. (e) Schematic diagram of a Zn–air battery describing the processes involved at the air electrode, in particular, with Na2CoP2O775 as the cathode and Zn metal as the anode, submerged in 6.0 M KOH containing 0.2 M Zn acetate as the additive. (f) Galvanostatic (current density, Id = 4 mA cm−2) discharge–charge cycling curves for two polymorphs of Na2CoP2O7, rechargeable Zn–air batteries. Reproduced from Gond et al.75 with permission from the American Chemical Society.

Sea water batteries and hybrid Na–air batteries comprise sodium (Na) metal as the anode and bifunctional electrocatalysts as the air cathode separated by a selective ionic conducting membrane, e.g. NASICON-type Na3Zr2SiPO12. Such secondary storage systems are unique combinations of non-aqueous, i.e. organic solvent-based and aqueous electrolytes (seawater in case of seawater battery) at the anode and cathode sides, respectively, as shown in Fig. 6a and b. A typical charge–discharge plot for sea water or hybrid Na–air batteries is shown in Fig. 6c. A suite of materials such as Na0.5Co0.5Mn0.5O2, NaCoPO4F and K2Co(PO3)4 showing bifunctional activities have been employed as air-cathodes for sea water/hybrid Na–air battery applications.84,88,114 Manikandan et al. demonstrated an outstanding electrochemical performance of P2-type layered Na0.5Co0.5Mn0.5O2 oxide electrocatalyst as the cathode for rechargeable seawater battery applications, where the half-cell battery showed a voltage gap of ∼0.78 V with an 80% voltage efficiency for 50 cycles (5 h charge and 5 discharge) at 0.1 mA current. Similarly, NaCoPO4F and K2Co(PO3)4 showed a potential gap (overpotential) of ∼0.40 and 0.70 V, respectively, in the case of hybrid Na–air batteries. The voltage efficiencies in the case of NaCoPO4F and K2Co(PO3)4 as the cathode were found to be 88% and 72% at current densities of 10 and 20 μA cm−2, respectively (Table 3).

Table 3 Metal–air battery performance using selected phosphate-based battery insertion materials as air cathodes (round-trip efficiency calculation based on charge–discharge voltage plateaus)
Electrocatalysts Electrolyte Type of metal–air battery Current density (mA cm−2) Open-circuit voltage (OCV in V) Overpotential gap (V) Round trip efficiency (%) Cycles Ref.
Na0.5Co0.5Mn0.5O2 Sea water Sea water 0.05 3.10 0.78 80 50 114
NaCoPO4F 0.1 M NaOH Hybrid Na–air 0.01 3.10 0.40 ∼88 30 84
K2Co(PO3)4 0.1 M NaOH Hybrid Na–air 0.02 2.85 0.70 ∼72 14 88
Co2P2O7/CoPi 6.0 M KOH Zn–air 10 1.41 0.70 ∼65 20 65
ZnCoP2O7 6.0 M KOH + 0.2 M Zn(Ac)2 Zn–air 1.0 1.50 0.80 60 25 66
t-Na2CoP2O7 6.0 M KOH + 0.2 M Zn(Ac)2 Zn–air 4.0 1.34 1.13 ∼42 20 75
o-Na2CoP2O7 6.0 M KOH + 0.2 M Zn(Ac)2 Zn–air 4.0 1.33 0.93 ∼53 20 75
K2CoP2O7 6.0 M KOH + 0.2 M Zn(Ac)2 Zn–air 1.7 1.35 <0.81 ∼52 112 87


Similarly, Zn–air battery fabrication is another sustainable approach towards energy storage employing bifunctional electrocatalytic active materials as cathodes, as illustrated in Fig. 6d. Zn–air batteries comprise Zn–metal as the anode and bifunctional electrocatalysts as the air cathode dipped in 6.0 M KOH electrolyte with or without 0.2 M Zn acetate as an additive. It is schematically shown in Fig. 6e, followed by the typical galvanostatic discharge–charge plot in Fig. 6f. The metal–air battery performance achieved by selected phosphate-based insertion materials as air cathodes is summarised in Table 3, where round trip efficiency was calculated from charge–discharge voltage plateaus. The assembled Zn–air batteries based on phosphate-based non-intercalation materials such as Co2P2O7/CoPi and ZnCoP2O7 as air cathodes exhibit high open-circuit voltages (OCVs) of 1.41 V and 1.50 V, respectively, compared to other phosphate-based intercalation materials such as t-Na2CoP2O7o-Na2CoP2O7, and K2CoP2O7 as air cathodes showing OCVs of 1.34 V, 1.33 V and 1.35 V, respectively.65,66,75,87 The ratio of the discharge to charge voltages at the end of the cycle gives the round-trip efficiency. At current densities of 10, 1.0, 4, and 1.7 mA cm−2, for Co2P2O7/CoPi, ZnCoP2O7, Na2CoP2O7, and K2CoP2O7, respectively, towards the end of the discharge–charge cycling, an overpotential gap of 0.70/0.80, 1.13, 0.93, and <0.81 V was achieved. The roundtrip efficiencies of ∼65, 60, ∼42, ∼53 and ∼52% were obtained for Co2P2O7/CoPi (end of 20 cycles), ZnCoP2O7 (end of 25 cycles), t-Na2CoP2O7 (end of 20 cycles), o-Na2CoP2O7 (end of 20 cycles), and K2CoP2O7 (end of 100 cycles), respectively.

6. Summary and outlook

In summary, we have broadly overviewed phosphate-based polyanionic battery insertion materials, which exhibit bifunctional (ORR and OER) electrocatalytic activity. Phosphorus plays vital roles in life, including genome stability, energetics, regulation, and signalling. In the energy storage sector, phosphate-based polyanion materials (e.g. LiFePO4 and Na3V2(PO4)2F3) have realized commercial battery applications combining their low cost and chemical/thermal stability.24,115 Moreover, exploiting the structure and transition metal centre, many PO4-based (non) intercalation compounds can work as economic electrocatalysts. Using these compounds as oxygen electrocatalysts, (hybrid) metal–air batteries can be harnessed for energy storage.

Metal–air batteries, like Li–air/O2 batteries, require advanced electrocatalytic materials that are not expensive and environmentally friendly. Current conventional catalysts such as IrO2 and RuO2 are expensive, limiting their practical applications. Polyanionic phosphate-based materials offer a variety of polymorphs that are safer and less harmful to the environment than the LaxSryMn1−yO3 perovskite-type electrocatalyst, with the highest OER activity known to date. These phosphate-based electrocatalysts can be easily prepared via conventional solid-state and wet chemical routes like a solution-combustion method. Depending on the synthesis conditions (e.g. annealing temperature), various polymorphs with diverse structures can be realized exhibiting electrocatalytic properties. The nanoscale electrodes synthesized via the solution-combustion approach showed superior electrocatalytic properties due to the increased electrode/electrolyte contact area, spatial confinement, one-pot carbon coating, and surface contribution towards adsorption and desorption of intermediates.

The electrochemical properties such as existing working voltage and specific discharge capacity of the polyanionic phosphate-based materials are briefly discussed, followed by their electrocatalytic properties. There are many PO4-based cathode materials (e.g. Na2CoPO4F), which cannot be used in batteries as their electrochemical activity is limited by the unavailability of stable high-voltage electrolytes. Nonetheless, exploiting their bifunctional electrocatalytic behavior, they can be alternately employed in metal–air batteries for energy storage. In addition, intrinsic oxygen electrocatalytic properties of cathode materials can enable spent cathodes to have an additional life after death. For example, the LiNi1−x−yMnxCoyO2 (NMC)-type oxide cathodes that are commonly used in LIBs can be directly extracted from the dead LIBs and can be converted into nanosized catalysts to improve their bifunctional (ORR/OER) catalytic activities, enabling their application as air-cathode components in zinc-air batteries.116

Recent studies on electrocatalysis have revealed that the underlying electrocatalytic activity of polyanionic phosphate-based materials can be tuned by various factors such as the type of transition metal centre (Co, Ni, Mn, and Fe), alkali cation, and conformation of the phosphate unit as the ligand present in the phosphate framework structures. Their role and impact on the ORR and OER activity have been discussed in this review. Herein, we also summarized the mechanism underlying the variation in activity arising due to the change in TM, alkali cation and PO4-unit, which is a correlation between 3d-electrons, eg occupancy and metal-oxygen covalency. They can be used as an activity descriptor, affirmatively extended from the d-band theory activity descriptor, to explain the mechanisms for oxygen electrocatalysis.

Phosphate-based polyanionic materials offer a rich playground for material discovery with wide structural diversity and polymorphism. The local structure can be further tuned by iso/aliovalent doping, which can affect the final electrocatalytic performance. The nanosizing and degree of crystallinity can be altered to improve their performance. They form economic and eco-friendly candidates for oxygen electrocatalysis, which can be readily employed in (hybrid) Na–air batteries and Zn–air batteries. Most importantly, exploiting the bifunctional activity and integrating the carbon matrices into polyanionic materials during combustion followed by Ar-annealing can have an interesting aspect towards carbon-free air-electrode. It can be an effective strategy to improve the amount of charge storage and prolonged cycle life in metal–air batteries with high active material loadings. With suitable material advancement, these polyanionic phosphates can ideally be implemented in Li–air/O2 battery applications, which offer a vast area for future research.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The current work was financially supported by the Technology Mission Division (Department of Science and Technology, Govt. of India) under the aegis of Materials for Energy Storage (MES-2018) program (DST/TMD/MES/2K18/207), and the Swedish Energy Agency (P2020-90216 and 50674-1). PB thanks the Alexander von Humboldt Foundation (Bonn, Germany) for a 2022 Humboldt fellowship for experienced researchers.

References

  1. J. H. Kang, J. Lee, J. W. Jung, J. Park, T. Jang, H. S. Kim, J. S. Nam, H. Lim, K. R. Yoon, W. H. Ryu, I. D. Kim and H. R. Byon, Lithium–air batteries: Air-breathing challenges and perspective, ACS Nano, 2020, 14, 14549–14578 CrossRef CAS PubMed.
  2. K. M. Abraham and Z. Jiang, A polymer electrolyte-based rechargeable lithium/oxygen battery, J. Electrochem. Soc., 1996, 143, 1–5 CrossRef CAS.
  3. Y. Shao, S. Park, J. Xiao, J. G. Zhang, Y. Wang and J. Liu, Electrocatalysts for nonaqueous lithium–air batteries: Status, challenges, and perspective, ACS Catal., 2012, 2, 844–857 CrossRef CAS.
  4. T. M. Gür, Review of electrical energy storage technologies, materials and systems: challenges and prospects for large-scale grid storage, Energy Environ. Sci., 2018, 11, 2696–2767 RSC.
  5. L. Wang, Y. Hao, L. Deng, F. Hu, S. Zhao, L. Li and S. Peng, Rapid complete reconfiguration induced actual active species for industrial hydrogen evolution reaction, Nat. Commun., 2022, 13, 5785 CrossRef CAS PubMed.
  6. L. Li, D. Yu, P. Li, H. Huang, D. Xie, C. C. Lin, F. Hu, H. Y. Chen and S. Peng, Interfacial electronic coupling of ultrathin transition-metal hydroxide nanosheets with layered MXenes as a new prototype for platinum-like hydrogen evolution, Energy Environ. Sci., 2021, 14, 6419–6427 RSC.
  7. Y. Hao, D. Yu, S. Zhu, C. H. Kuo, Y. M. Chang, L. Wang, H. Y. Chen, M. Shao and S. Peng, Methanol upgrading coupled with hydrogen product at large current density promoted by strong interfacial interactions, Energy Environ. Sci., 2023, 16, 1100–1110 RSC.
  8. S. Zhao, F. Hu, L. Yin, L. Li and S. Peng, Manipulating electron redistribution induced by asymmetric coordination for electrocatalytic water oxidation at a high current density, Sci. Bull., 2023, 68, 1389–1398 CrossRef CAS PubMed.
  9. J. Hwang, R. R. Rao, L. Giordano, Y. Katayama, Y. Yu and Y. Shao-Horn, Perovskites in catalysis and electrocatalysis, Science, 2017, 358, 751–756 CrossRef CAS PubMed.
  10. M. Risch, A. Grimaud, K. J. May, K. A. Stoerzinger, T. J. Chen, A. N. Mansour and Y. Shao-Horn, Structural changes of cobalt-based perovskites upon water oxidation investigated by EXAFS, J. Phys. Chem. C, 2013, 117, 8628–8635 CrossRef CAS.
  11. Y. Yang, W. Yin, S. Wu, X. Yang, W. Xia, Y. Shen, Y. Huang, A. Cao and Q. Yuan, Perovskite-type LaSrMnO electrocatalyst with uniform porous structure for an efficient Li–O2 battery cathode, ACS Nano, 2016, 10, 1240–1248 CrossRef CAS PubMed.
  12. Y. Meng, G. Ni, X. Jin, J. Peng and Q. Y. Yan, Recent advances in the application of phosphates and borates as electrocatalysts for water oxidation, Mater. Today Nano, 2020, 12, 100095 CrossRef.
  13. W. T. Hong, M. Risch, K. A. Stoerzinger, A. Grimaud, J. Suntivich and Y. Shao-Horn, Toward the rational design of non-precious transition metal oxides for oxygen electrocatalysis, Energy Environ. Sci., 2015, 8, 1404–1427 RSC.
  14. F. Li, T. Zhang and H. Zhou, Challenges of non-aqueous Li–O2 batteries: electrolytes, catalysts, and anodes, Energy Environ. Sci., 2013, 6, 1125–1141 RSC.
  15. Y. C. Lu, B. M. Gallant, D. G. Kwabi, J. R. Harding, R. R. Mitchell, M. S. Whittingham and Y. Shao-Horn, Lithium–oxygen batteries: bridging mechanistic understanding and battery performance, Energy Environ. Sci., 2013, 6, 750–768 RSC.
  16. Y. C. Lu and Y. Shao-Horn, Probing the reaction kinetics of the charge reactions of nonaqueous Li–O2 batteries, J. Phys. Chem. Lett., 2013, 4, 93–99 CrossRef CAS PubMed.
  17. D. G. Kwabi, N. Ortiz-Vitoriano, S. A. Freunberger, Y. Chen, N. Imanishi, P. G. Bruce and Y. Shao-Horn, Materials challenges in rechargeable lithium-air batteries, MRS Bull., 2014, 39, 443–452 CrossRef CAS.
  18. Y. C. Lu, H. A. Gasteiger, M. C. Parent, V. Chiloyan and Y. Shao-Horn, The influence of catalysts on discharge and charge voltages of rechargeable Li-oxygen batteries, Electrochem. Solid-State Lett., 2010, 13, 13–17 Search PubMed.
  19. Z. Ma, X. Yuan, L. Li, Z. F. Ma, D. P. Wilkinson, L. Zhang and J. Zhang, A review of cathode materials and structures for rechargeable lithium–air batteries, Energy Environ. Sci., 2015, 8, 2144–2198 RSC.
  20. A. Gulick, Phosphorus as a factor in the origin of life, Am. Sci., 1955, 43, 479–489 CAS.
  21. M. W. Bowler, M. J. Cliff, J. P. Waltho and G. M. Blackburn, Why did nature select phosphate for its dominant roles in biology?, New J. Chem., 2010, 34, 784–794 RSC.
  22. K. W. Knouse, D. T. Flood, J. C. Vantourout, M. A. Schmidt, I. M. McDonald, M. D. Eastgate and P. S. Baran, Nature chose phosphates and chemists should too: How emerging P(V) methods can augment existing strategies, ACS Cent. Sci., 2021, 7, 1473–1485 CrossRef CAS PubMed.
  23. F. H. Westheimer, Why nature chose phosphates, Science, 1987, 235, 1173–1178 CrossRef CAS PubMed.
  24. A. Masias, J. Marcicki and W. A. Paxton, Opportunities and challenges of lithium ion batteries in automotive applications, ACS Energy Lett., 2021, 6, 621–630 CrossRef CAS.
  25. D. Andre, S. J. Kim, P. Lamp, S. F. Lux, F. Maglia, O. Paschos and B. Stiaszny, Future generations of cathode materials: an automotive industry perspective, J. Mater. Chem. A, 2015, 3, 6709–6732 RSC.
  26. X. Yu and A. Manthiram, Sustainable battery materials for next-generation electrical energy storage, Adv. Energy Sustainability Res., 2021, 2, 2000102 CrossRef CAS.
  27. P. J. Ferreira, G. J. la O’, Y. Shao-Horn, D. Morgan, R. Makharia, S. Kocha and H. A. Gasteiger, Instability of Pt/C electrocatalysts in proton exchange membrane fuel cells: a mechanistic investigation, J. Electrochem. Soc., 2005, 152, A2256 CrossRef.
  28. K. J. J. Mayrhofer, K. Hartl, V. Juhart and M. Arenz, Degradation of carbon-supported Pt bimetallic nanoparticles by surface segregation, J. Am. Chem. Soc., 2009, 131, 16348–16349 CrossRef CAS PubMed.
  29. C. C. L. McCrory, S. Jung, J. C. Peters and T. F. Jaramillo, Benchmarking heterogeneous electrocatalysts for the oxygen evolution reaction, J. Am. Chem. Soc., 2013, 135, 16977–16987 CrossRef CAS PubMed.
  30. C. C. L. McCrory, S. Jung, I. M. Ferrer, S. M. Chatman, J. C. Peters and T. F. Jaramillo, Benchmarking hydrogen evolving reaction and oxygen evolving reaction electrocatalysts for solar water splitting devices, J. Am. Chem. Soc., 2015, 137, 4347–4357 CrossRef CAS PubMed.
  31. G. Charalampides and K. I. Vatalis, Global production estimation of rare earth elements and their environmental impacts on soils, J. Geosci. Environ. Prot., 2015, 03, 66–73 Search PubMed.
  32. V. Balaram, Rare earth elements: A review of applications, occurrence, exploration,analysis, recycling, and environmental impact, Geosci. Front., 2019, 10, 1285–1303 CrossRef CAS.
  33. A. Grimaud, W. T. Hong, Y. Shao-Horn and J. M. Tarascon, Anionic redox processes for electrochemical devices, Nat. Mater., 2016, 15, 121–126 CrossRef CAS PubMed.
  34. H. Chen and M. S. Islam, Lithium extraction mechanism in Li-rich Li2MnO3 involving oxygen hole formation and dimerization, Chem. Mater., 2016, 28, 6656–6663 CrossRef CAS.
  35. R. Sharpe, R. A. House, M. J. Clarke, D. Förstermann, J. J. Marie, G. Cibin, K. J. Zhou, H. Y. Playford, P. G. Bruce and M. S. Islam, Redox chemistry and the role of trapped molecular O2 in Li-rich disordered rocksalt oxyfluoride cathodes, J. Am. Chem. Soc., 2020, 142, 21799–21809 CrossRef CAS PubMed.
  36. P. Barpanda, T. Ye, S. C. Chung, Y. Yamada, S. I. Nishimura and A. Yamada, Eco-efficient splash combustion synthesis of nanoscale pyrophosphate (Li2FeP2O7) positive-electrode using Fe(III) precursors, J. Mater. Chem., 2012, 22, 13455–13459 RSC.
  37. B. Senthilkumar, C. Murugesan, L. Sharma, S. Lochab and P. Barpanda, An overview of mixed polyanionic cathode materials for sodium-ion batteries, Small Methods, 2019, 3, 1800253 CrossRef CAS.
  38. F. T. Li, J. Ran, M. Jaroniec and S. Z. Qiao, Solution combustion synthesis of metal oxide nanomaterials for energy storage and conversion, Nanoscale, 2015, 7, 17590–17610 RSC.
  39. A. Varma, A. S. Mukasyan, A. S. Rogachev and K. V. Manukyan, Solution combustion synthesis of nanoscale materials, Chem. Rev., 2016, 116, 14493–14586 CrossRef CAS PubMed.
  40. S. Bashir, P. Hanumandla, H. Y. Huang and J. L. Liu, Nanostructured Materials for Next-Generation Energy Storage and Conversion, Fuel Cells, 2018, 4, 517–542 Search PubMed.
  41. X. Guo, P. Liu, J. Han, Y. Ito, A. Hirata, T. Fujita and M. Chen, 3D nanoporous nitrogen-doped graphene with encapsulated RuO2 nanoparticles for Li–O2 batteries, Adv. Mater., 2015, 27, 6137–6143 CrossRef CAS PubMed.
  42. Z. Jian, P. Liu, F. Li, P. He, X. Guo, M. Chen and H. Zhou, Core–shell-structured CNT@RuO2 composite as a high-performance cathode catalyst for rechargeable Li–O2 batteries, Angew. Chem., Int. Ed., 2014, 53, 442–446 CrossRef CAS PubMed.
  43. J. Lu, Y. Lei, K. C. Lau, X. Luo, P. Du, J. Wen, R. S. Assary, U. Das, D. J. Miller, J. W. Elam, H. M. Albishri, D. A. El-Hady, Y. K. Sun, L. A. Curtiss and K. Amine, A nanostructured cathode architecture for low charge overpotential in lithium-oxygen batteries, Nat. Commun., 2013, 4, 1–10 Search PubMed.
  44. Z. W. She, J. Kibsgaard, C. F. Dickens, I. Chorkendorff, J. K. Nørskov and T. F. Jaramillo, Combining theory and experiment in electrocatalysis: Insights into materials design, Science, 2017, 355, eaad4998 CrossRef PubMed.
  45. S. G. Mohamed, Y. Q. Tsai, C. J. Chen, Y. T. Tsai, T. F. Hung, W. S. Chang and R. S. Liu, Ternary spinel MCo2O4 (M = Mn, Fe, Ni, and Zn) porous nanorods as bifunctional cathode materials for lithium–O2 batteries, ACS Appl. Mater. Interfaces, 2015, 7, 12038–12046 CrossRef CAS PubMed.
  46. C. Chen, Y. Kang, Z. Huo, Z. Zhu, W. Huang, H. L. Xin, J. D. Snyder, D. Li, J. A. Herron, M. Mavrikakis, M. Chi, K. L. More, Y. Li, N. M. Markovic, G. A. Somorjai, P. Yang and V. R. Stamenkovic, Highly crystalline multimetallic nanoframes with three-dimensional electrocatalytic surfaces, Science, 2014, 343, 1339–1343 CrossRef CAS PubMed.
  47. Y. Lee, J. Suntivich, K. J. May, E. E. Perry and Y. Shao-Horn, Synthesis and activities of rutile IrO2 and RuO2 nanoparticles for oxygen evolution in acid and alkaline solutions, J. Phys. Chem. Lett., 2012, 3, 399–404 CrossRef CAS PubMed.
  48. L. Chen, Y. Wang, X. Zhao, Y. Wang, Q. Li, Q. Wang, Y. Tang and Y. Lei, Trimetallic oxyhydroxides as active sites for large-current-density alkaline oxygen evolution and overall water splitting, J. Mater. Sci. Technol., 2022, 110, 128–135 CrossRef CAS.
  49. Q. Liu, M. Ranocchiari and J. A. Van Bokhoven, Catalyst overcoating engineering towards high-performance electrocatalysis, Chem. Soc. Rev., 2022, 51, 188–236 RSC.
  50. M. Jin, X. Zhang, S. Niu, Q. Wang, R. Huang, R. Ling, J. Huang, R. Shi, A. Amini and C. Cheng, Strategies for designing high-performance hydrogen evolution reaction electrocatalysts at large current densities above 1000 mA cm−2, ACS Nano, 2022, 16, 11577–11597 CrossRef CAS PubMed.
  51. D. Yu, Y. Hao, S. Han, S. Zhao, Q. Zhou, C. H. Kuo, F. Hu, L. Li, H. Y. Chen, J. Ren and S. Peng, Ultrafast combustion synthesis of robust and efficient electrocatalysts for high-current-density water oxidation, ACS Nano, 2023, 17, 1701–1712 CrossRef CAS PubMed.
  52. X. Zhang, L. Li, E. Fan, Q. Xue, Y. Bian, F. Wu and R. Chen, Toward sustainable and systematic recycling of spent rechargeable batteries, Chem. Soc. Rev., 2018, 47, 7239–7302 RSC.
  53. R. M. Izatt, S. R. Izatt, R. L. Bruening, N. E. Izatt and B. A. Moyer, Challenges to achievement of metal sustainability in our high-tech society, Chem. Soc. Rev., 2014, 43, 2451–2475 RSC.
  54. T. N. Q. Le, Q. D. Tran, N. N. Tran, C. Priest, W. Skinner, M. Goodsite, C. Spandler, N. J. Cook and V. Hessel, Critical elements: opportunities for microfluidic processing and potential for ESG-powered mining investments, Green Chem., 2022, 24, 8879–8898 RSC.
  55. M. Iturrondobeitia, C. Vallejo, M. Berroci, O. Akizu-Gardoki, R. Minguez and E. Lizundia, Environmental impact assessment of LiNi1/3Mn1/3Co1/3O2 hydrometallurgical cathode recycling from spent lithium-ion batteries, ACS Sustainable Chem. Eng., 2022, 10, 9798–9810 CrossRef CAS.
  56. K. Liu, Q. Tan, L. Liu and J. Li, Acid-free and selective extraction of lithium from spent lithium iron phosphate batteries via a mechanochemically induced isomorphic substitution, Environ. Sci. Technol., 2019, 53, 9781–9788 CrossRef CAS PubMed.
  57. A. Manthiram and J. B. Goodenough, Lithium-based polyanion oxide cathodes, Nat. Energy, 2021, 6, 844–845 CrossRef CAS.
  58. H. Osgood, S. V. Devaguptapu, H. Xu, J. Cho and G. Wu, Transition metal (Fe, Co, Ni, and Mn) oxides for oxygen reduction and evolution bifunctional catalysts in alkaline media, Nano Today, 2016, 11, 601–625 CrossRef CAS.
  59. J. B. Goodenough and Y. Kim, Challenges for rechargeable Li batteries, Chem. Mater., 2010, 22, 587–603 CrossRef CAS.
  60. S. W. Lee, C. Carlton, M. Risch, Y. Surendranath, S. Chen, S. Furutsuki, A. Yamada, D. G. Nocera and Y. Shao-Horn, The nature of lithium battery materials under oxygen evolution reaction conditions, J. Am. Chem. Soc., 2012, 134, 16959–16962 CrossRef CAS PubMed.
  61. T. Maiyalagan, K. A. Jarvis, S. Therese, P. J. Ferreira and A. Manthiram, Spinel-type lithium cobalt oxide as a bifunctional electrocatalyst for the oxygen evolution and oxygen reduction reactions, Nat. Commun., 2014, 5, 1–8 Search PubMed.
  62. J. Suntivich, K. J. May, H. A. Gasteiger, J. B. Goodenough, Y. Shao-horn, F. Calle-vallejo, A. D. Oscar, M. J. Kolb, M. T. M. Koper, J. Suntivich, K. J. May, H. A. Gasteiger, J. B. Goodenough and Y. Shao-horn, A Perovskite Oxide Optimized for Molecular Orbital Principles, Science, 2011, 334(6061), 2010–2012 CrossRef PubMed.
  63. H. Kim, J. Park, I. Park, K. Jin, S. E. Jerng, S. H. Kim, K. T. Nam and K. Kang, Coordination tuning of cobalt phosphates towards efficient water oxidation catalyst, Nat. Commun., 2015, 6, 8253 CrossRef CAS PubMed.
  64. M. W. Kanan and D. G. Nocera, In situ formation of an oxygen-evolving catalyst in neutral water containing phosphate and Co2+, Science, 2008, 321, 1072–1075 CrossRef CAS PubMed.
  65. J. T. Ren, G. G. Yuan, L. Chen, C. C. Weng and Z. Y. Yuan, Rational dispersion of Co2P2O7 fine particles on N,P-codoped reduced graphene oxide aerogels leading to enhanced reversible oxygen reduction ability for Zn–air batteries, ACS Sustainable Chem. Eng., 2018, 6, 9793–9803 CrossRef CAS.
  66. A. Baby, D. Singh, C. Murugesan and P. Barpanda, The design of zinc-substituted cobalt (pyro)phosphates as efficient bifunctional electrocatalysts for zinc–air batteries, Chem. Commun., 2020, 56, 8400–8403 RSC.
  67. S. Ma, Q. Zhu, L. Chen, W. Wang and D. Chen, Large-scale synthesis of LiNi0.75Fe0.25PO4 covalently anchored on graphene nanosheets for remarkable electrochemical water oxidation, J. Mater. Chem. A, 2016, 4, 8149–8154 RSC.
  68. X. Wu, Y. Lin, Y. Ji, D. Zhou, Z. Liu and X. Sun, Insights into the enhanced catalytic activity of Fe-doped LiCoPO4 for the oxygen evolution reaction, ACS Appl. Energy Mater., 2020, 3, 2959–2965 CrossRef CAS.
  69. Y. Gershinsky and D. Zitoun, Direct chemical synthesis of lithium sub-stochiometric olivine Li0.7Co0.75Fe0.25PO4 coated with reduced graphene oxide as oxygen evolution reaction electrocatalyst, ACS Catal., 2018, 8, 8715–8725 CrossRef CAS.
  70. Y. Liu, H. Wang, D. Lin, C. Liu, P. C. Hsu, W. Liu, W. Chen and Y. Cui, Electrochemical tuning of olivine-type lithium transition-metal phosphates as efficient water oxidation catalysts, Energy Environ. Sci., 2015, 8, 1719–1724 RSC.
  71. M. H. Nasir, N. K. Janjua and J. Santoki, Electrochemical performance of carbon modified LiNiPO4 as Li-ion battery cathode: A combined experimental and theoretical study, J. Electrochem. Soc., 2020, 167, 130526 CrossRef CAS.
  72. E. Loni, M. H. Siadati, A. Shokuhfar, D. Leybo and D. Kuznetsov, Sodium–cobalt pyrophosphate electrocatalyst for water splitting, J. Solid State Chem., 2020, 290, 121510 CrossRef CAS.
  73. R. Gond, K. Sada, B. Senthilkumar and P. Barpanda, Bifunctional electrocatalytic behavior of sodium cobalt phosphates in alkaline solution, ChemElectroChem, 2018, 5, 153–158 CrossRef CAS.
  74. P. Barpanda, S. I. Nishimura and A. Yamada, High-voltage pyrophosphate cathodes, Adv. Energy Mater., 2012, 2, 841–859 CrossRef CAS.
  75. R. Gond, S. Singh, X. Zhao, D. Singh, R. Ahuja, M. Fichtner and P. Barpanda, Pyrophosphate Na2CoP2O7 polymorphs as efficient bifunctional oxygen electrocatalysts for zinc-air batteries, ACS Appl. Mater. Interfaces, 2022, 14, 40761–40770 CrossRef CAS PubMed.
  76. R. Gond, S. P. Vanam and P. Barpanda, Na2MnP2O7 polymorphs as efficient bifunctional catalysts for oxygen reduction and oxygen evolution reactions, Chem. Commun., 2019, 55, 11595–11598 RSC.
  77. R. Gond, R. P. Rao, V. Pralong, O. I. Lebedev, S. Adams and P. Barpanda, Cubic sodium cobalt metaphosphate [NaCo(PO3)3] as a cathode material for sodium ion batteries, Inorg. Chem., 2018, 57, 6324–6332 CrossRef CAS PubMed.
  78. R. Gond, D. K. Singh, M. Eswaramoorthy and P. Barpanda, Sodium cobalt metaphosphate as an efficient oxygen evolution reaction catalyst in alkaline solution, Angew. Chem., 2019, 8418–8423 CrossRef.
  79. C. Murugesan, M. Musthafa, S. Lochab and P. Barpanda, Cobalt metaphosphates as economic bifunctional electrocatalysts for hybrid sodium–air batteries, Inorg. Chem., 2021, 60, 11974–11983 CrossRef CAS PubMed.
  80. D. Dwibedi, R. Gond and P. Barpanda, Alluaudite NaCoFe2(PO4)3 as a 2.9 V cathode for sodium-ion batteries exhibiting bifunctional electrocatalytic activity, Chem. Mater., 2019, 31, 7501–7509 CrossRef CAS.
  81. K. Kubota, K. Yokoh, N. Yabuuchi and S. Komaba, Na2CoPO4F as a high-voltage electrode material for Na-ion batteries, Electrochemistry, 2014, 82, 909–911 CrossRef CAS.
  82. N. Recham, J. N. Chotard, J. C. Jumas, L. Laffont, M. Armand and J. M. Tarasco, Ionothermal synthesis of Li-based fluorophosphates electrodes, Chem. Mater., 2010, 22, 1142–1148 CrossRef CAS.
  83. L. Sharma, N. Bothra, R. K. Rai, S. Pati and P. Barpanda, Metal fluorophosphate polyanionic insertion hosts as efficient bifunctional electrocatalysts for oxygen evolution and reduction reactions, J. Mater. Chem. A, 2020, 8, 18651–18658 RSC.
  84. L. Sharma, R. Gond, B. Senthilkumar, A. Roy and P. Barpanda, Fluorophosphates as efficient bifunctional electrocatalysts for metal–air batteries, ACS Catal., 2020, 10, 43–50 CrossRef CAS.
  85. T. Hosaka, T. Shimamura, K. Kubota and S. Komaba, Polyanionic compounds for potassium-ion batteries, Chem. Rec., 2019, 19, 735–745 CrossRef CAS PubMed.
  86. I. Sultana, M. M. Rahman, S. Mateti, N. Sharma, S. Huang and Y. Chen, Approaching reactive KFePO4 phase for potassium storage by adopting an advanced design strategy, Batteries Supercaps, 2020, 3, 450–455 CrossRef CAS.
  87. K. Sada, R. Gond, N. Bothra, S. K. Pati and P. Barpanda, Potassium cobalt pyrophosphate as a nonprecious bifunctional electrocatalyst for zinc–air batteries, ACS Appl. Mater. Interfaces, 2022, 14, 8992–9001 CrossRef CAS PubMed.
  88. C. Murugesan, S. P. Panjalingam, S. Lochab, R. K. Rai, X. F. Zhao, D. Singh, R. Ahuja and P. Barpanda, Cobalt tetraphosphate as an efficient bifunctional electrocatalyst for hybrid sodium-air batteries, Nano Energy, 2021, 89, 106485 CrossRef CAS.
  89. T. Ahamad, M. Naushad, R. Hassan Mousa, N. Khalaf and S. M. Alshehri, Synthesis and characterization cobalt phosphate embedded with N doped carbon for water splitting ORR and OER, J. King Saud Univ., Sci., 2020, 32, 2826–2830 CrossRef.
  90. A. K. Padhi, K. S. Nanjundaswamy and J. B. Goodenough, Phospho-olivines as positive-electrode materials for rechargeable lithium batteries, J. Electrochem. Soc., 1997, 144, 1184–1194 Search PubMed.
  91. T. Muraliganth and A. Manthiram, Understanding the shifts in the redox potentials of olivine LiM1−yMyPO4 (M = Fe, Mn, Co, and Mg) solid solution cathodes, J. Phys. Chem. C, 2010, 114, 15530–15540 CrossRef CAS.
  92. M. J. Lee, J. S. Kang, D. Ahn, D. Y. Chung, S. Park, Y. J. Son, J. M. Yoo, H. Shin, Y. S. Kang, N. E. Sung, K. S. Lee and Y. E. Sung, Lithium manganese phosphate-carbon composite as a highly active and durable electrocatalyst for oxygen reduction reaction, Electrochim. Acta, 2017, 245, 219–226 CrossRef CAS.
  93. H. Kim, S. Lee, Y. U. Park, H. Kim, J. Kim, S. Jeon and K. Kang, Neutron and X-ray diffraction study of pyrophosphate-based Li2–xMP2O7 (M = Fe, Co) for lithium rechargeable battery electrodes, Chem. Mater., 2011, 23, 3930–3937 CrossRef CAS.
  94. A. Chiring, M. Mazumder, S. K. Pati, C. S. Johnson and P. Senguttuvan, Unraveling the formation mechanism of NaCoPO4 polymorphs, J. Solid State Chem., 2021, 293, 121766 CrossRef CAS.
  95. Y. Zhu, Y. Xu, Y. Liu, C. Luo and C. Wang, Comparison of electrochemical performances of olivine NaFePO4 in sodium-ion batteries and olivine LiFePO4 in lithium-ion batteries, Nanoscale, 2013, 5, 780–787 RSC.
  96. C. Murugesan, S. Lochab, B. Senthilkumar and P. Barpanda, Earth-abundant alkali iron phosphates (AFePO4) as efficient electrocatalysts for the oxygen reduction reaction in alkaline solution, ChemCatChem, 2018, 10, 1122–1127 CrossRef CAS.
  97. R. Gond, S. S. Meena, S. M. Yusuf, V. Shukla, N. K. Jena, R. Ahuja, S. Okada and P. Barpanda, Enabling the electrochemical activity in sodium iron metaphosphate [NaFe(PO3)3] sodium battery insertion material: Structural and electrochemical insights, Inorg. Chem., 2017, 56, 5918–5929 CrossRef CAS PubMed.
  98. H. Wan, R. Ma, X. Liu, J. Pan, H. Wang, S. Liang, G. Qiu and T. Sasaki, Rare cobalt-based phosphate nanoribbons with unique 5-coordination for electrocatalytic water oxidation, ACS Energy Lett., 2018, 3, 1254–1260 CrossRef CAS.
  99. L. Sharma, P. K. Nayak, E. De La Llave, H. Chen, S. Adams, D. Aurbach and P. Barpanda, Electrochemical and diffusional investigation of Na2FeIIPO4F fluorophosphate sodium insertion material obtained from FeIII precursor, ACS Appl. Mater. Interfaces, 2017, 9, 34961–34969 CrossRef CAS PubMed.
  100. X. Lin, X. Hou, X. Wu, S. Wang, M. Gao and Y. Yang, Exploiting Na2MnPO4F as a high-capacity and well-reversible cathode material for Na-ion batteries, RSC Adv., 2014, 4, 40985–40993 RSC.
  101. P. Barpanda, T. Ye, M. Avdeev, S. C. Chung and A. Yamada, A new polymorph of Na2MnP2O7 as a 3.6 V cathode material for sodium-ion batteries, J. Mater. Chem. A, 2013, 1, 4194–4197 RSC.
  102. P. Barpanda, J. Lu, T. Ye, M. Kajiyama, S. C. Chung, N. Yabuuchi, S. Komaba and A. Yamada, A layer-structured Na2CoP2O7 pyrophosphate cathode for sodium-ion batteries, RSC Adv., 2013, 3, 3857–3860 RSC.
  103. H. Kim, C. S. Park, J. W. Choi and Y. Jung, Defect-controlled formation of triclinic Na2CoP2O7 for 4 V sodium-ion batteries, Angew. Chem., Int. Ed., 2016, 55, 6662–6666 CrossRef CAS PubMed.
  104. W. B. Park, S. C. Han, C. Park, S. U. Hong, U. Han, S. P. Singh, Y. H. Jung, D. Ahn, K. S. Sohn and M. Pyo, KVP2O7 as a robust high-energy cathode for potassium-ion batteries: Pinpointed by a full screening of the inorganic registry under specific search conditions, Adv. Energy Mater., 2018, 8, 1–12 CAS.
  105. F. Zhou, M. Cococcioni, K. Kang and G. Ceder, The Li intercalation potential of LiMPO4 and LiMSiO4 olivines with M = Fe, Mn, Co, Ni, Electrochem. Commun., 2004, 6, 1144–1148 CrossRef CAS.
  106. M. Aydinol, A. Kohan, G. Ceder, K. Cho and J. Joannopoulos, Ab initio study of lithium intercalation in metal oxides and metal dichalcogenides, Phys. Rev. B: Condens. Matter Mater. Phys., 1997, 56, 1354–1365 CrossRef CAS.
  107. A. Grimaud, K. J. May, C. E. Carlton, Y. L. Lee, M. Risch, W. T. Hong, J. Zhou and Y. Shao-Horn, Double perovskites as a family of highly active catalysts for oxygen evolution in alkaline solution, Nat. Commun., 2013, 4, 1–7 Search PubMed.
  108. J. Suntivich, H. A. Gasteiger, N. Yabuuchi, H. Nakanishi, J. B. Goodenough and Y. Shao-Horn, Design principles for oxygen-reduction activity on perovskite oxide catalysts for fuel cells and metal–air batteries, Nat. Chem., 2011, 3, 546–550 CrossRef CAS PubMed.
  109. Y. Duan, S. Sun, S. Xi, X. Ren, Y. Zhou, G. Zhang, H. Yang, Y. Du and Z. J. Xu, Tailoring the Co 3d-O 2p covalency in LaCoO3 by Fe substitution to promote oxygen evolution reaction, Chem. Mater., 2017, 29, 10534–10541 CrossRef CAS.
  110. C. Wei, Z. Feng, G. G. Scherer, J. Barber, Y. Shao-Horn and Z. J. Xu, Cations in octahedral sites: A descriptor for oxygen electrocatalysis on transition-metal spinels, Adv. Mater., 2017, 29 CAS.
  111. J. Zaanen and G. A. Sawatzky, Systematics in band gaps and optical spectra of 3D transition metal compounds, J. Solid State Chem., 1990, 88, 8–27 CrossRef CAS.
  112. X. Lang, Z. Hu and C. Wang, Bifunctional air electrodes for flexible rechargeable zn-air batteries, Chin. Chem. Lett., 2021, 32, 999–1009 CrossRef CAS.
  113. Q. Wang, Q. Feng, Y. Lei, S. Tang, L. Xu, Y. Xiong, G. Fang, Y. Wang, P. Yang, J. Liu, W. Liu and X. Xiong, Quasi-solid-state Zn-air batteries with an atomically dispersed cobalt electrocatalyst and organohydrogel electrolyte, Nat. Commun., 2022, 13, 3689 CrossRef CAS PubMed.
  114. P. Manikandan, K. Kishor, J. Han and Y. Kim, Advanced perspective on the synchronized bifunctional activities of P2-type materials to implement an interconnected voltage profile for seawater batteries, J. Mater. Chem. A, 2018, 6, 11012–11021 RSC.
  115. J. M. Tarascon, Na-ion versus Li-ion batteries: Complementarity rather than competitiveness, Joule, 2020, 4, 1616–1620 CrossRef.
  116. M. Jiao, Q. Zhang, C. Ye, Z. Liu, X. Zhong, J. Wang, C. Li, L. Dai, G. Zhou and H. M. Cheng, Recycling spent LiNi1−x−yMnxCoyO2 cathodes to bifunctional NiMnCo catalysts for zinc-air batteries, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2202202119 CrossRef CAS PubMed.

This journal is © the Partner Organisations 2024