Next Article in Journal
Novel Bicyclic P,S-Heterocycles via Stereoselective hetero-Diels–Alder Reactions of Thiochalcones with 1-Phenyl-4H-phosphinin-4-one 1-Oxide
Previous Article in Journal
Synthesis and Application of Salicylhydrazone Probes with High Selectivity for Rapid Detection of Cu2+
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Central Metal Ion on Binuclear Metal Phthalocyanine-Based Redox Mediator for Lithium Carbonate Decomposition

1
School of Materials Science and Chemical Engineering, Ningbo University, Ningbo 315211, China
2
School of Biological and Chemical Engineering, NingboTech University, Ningbo 315100, China
3
College of Chemical and Biological Engineering, Zhejiang University, Hangzhou 310058, China
4
Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(9), 2034; https://doi.org/10.3390/molecules29092034
Submission received: 23 February 2024 / Revised: 24 April 2024 / Accepted: 24 April 2024 / Published: 28 April 2024

Abstract

:
Li2CO3 is the most tenacious parasitic solid-state product in lithium–air batteries (LABs). Developing suitable redox mediators (RMs) is an efficient way to address the Li2CO3 issue, but only a few RMs have been investigated to date, and their mechanism of action also remains elusive. Herein, we investigate the effects of the central metal ion in binuclear metal phthalocyanines on the catalysis of Li2CO3 decomposition, namely binuclear cobalt phthalocyanine (bi-CoPc) and binuclear cobalt manganese phthalocyanine (bi-CoMnPc). Density functional theory (DFT) calculations indicate that the key intermediate peroxydicarbonate (*C2O62−) is stabilized by bi-CoPc2+ and bi-CoMnPc3+, which is accountable for their excellent catalytic effects. With one central metal ion substituted by manganese for cobalt, the bi-CoMnPc’s second active redox couple shifts from the second Co(II)/Co(III) couple in the central metal ion to the Pc(-2)/Pc(-1) couple in the phthalocyanine ring. In artificial dry air (N2-O2, 78:22, v/v), the LAB cell with bi-CoMnPc in electrolyte exhibited 261 cycles under a fixed capacity of 500 mAh g−1carbon and current density of 100 mA g−1carbon, significantly better than the RM-free cell (62 cycles) and the cell with bi-CoPc (193 cycles).

Graphical Abstract

1. Introduction

Along with the fast-paced electrification, intelligentization, and mobilization of industry and daily life, the demand for energy storage devices with higher energy density becomes increasingly urgent. Lithium–air batteries (LABs) possess the highest theoretical energy density (~11,400 Wh/kg, excluding O2 mass) among all electrochemical energy storage technologies and have drawn tremendous attention in the past two decades. However, major challenges must be conquered before the practicality of LABs, originating from the highly reactive cathode/anode materials and (intermediate) products (such as metallic lithium, 1O2, LiO2, O2·, etc.), as well as the unstable solid–solid contact between discharge/parasitic reaction products and the air cathode [1,2,3,4].
In LABs, Li2CO3 is regarded as the “Achilles’ heel” [5] because it is both virtually inevitable to form and difficult to remove. On the one hand, Li2CO3 can be produced both from the absorption of adventitious CO2 and from the degradation of cell components (electrolyte solvents, lithium salts, carbon electrodes, etc.). On the other hand, Li2CO3 is the most tenacious parasitic solid-state product in LABs due to its high oxidative decomposition potential, low electronic/ionic conductivity, and low solubility in most organic solvents. Containing O2 in the battery system and converting LABs to lithium–oxygen batteries (LOBs) can mitigate Li2CO3 accumulation, but at the cost of greatly reduced energy density (~3450 Wh/kg in theory, including O2 mass). Therefore, it is necessary to develop suitable catalysts for Li2CO3 decomposition to enable the utilization of ambient air, especially the redox mediators (RMs) that can overcome the solid–solid contact difficulty between solid-state Li2CO3 and an air cathode.
Since the first discovery of Li2CO3 decomposing RM (a planar binuclear cobalt phthalocyanine, bi-CoPc) [6], several other RMs have been reported effective towards Li2CO3 decomposition, such as LiBr [7,8], conjugated cobalt polyphthalocyanine (CoPPc) [9], cobalt tetramethoxyphenyl porphyrin (CoTMPP) [10], ruthenocene (Ruc) [11], (1R,2R)-(-)-N,N-bis(3,5-di-t-butylsalicylidene)-1,2-cyclohexanediaminocobalt(II) (salen-Co(II)) [12], N,N,N′,N′-tetramethyl-p-phenylenediamine (TMPD) [13], 5,10-dimethylphenazine (DMPZ) [13], S-phenyl carbonothioate (SPC) [14], 2,5-di-tert-butyl-1,4-dimethoxybenzene (DBDMB) [15], etc. However, the Li2CO3 decomposition mechanism remains elusive. Ideally, Li2CO3 should decompose via Equation (1), which reverses the discharge reaction to reproduce O2 and CO2; however, numerous investigations have reported substantially less-than-stoichiometric O2 in the evolving gas due to the undesirable decomposition paths (Equations (2) and (3)) and following 1O2/O2 attacks on electrolyte components [16,17,18,19]. Furthermore, the RMs’ exact mechanism of action is still debated, especially the effectiveness of 1-electron RMs in catalyzing Li2CO3 decomposition. Our previous work proposed the necessity of 2-electron RMs [6], which is supported by some other 2- or multiple-electron ones, such as LiBr [7,8], CoPPc [9], CoTMPP [10], and SPC [14]. However, recently, some 1-electron RMs have also been reported to be effective, such as Ruc [11], salen-Co(II) [12], DMDMB [15], TMPD [13], DMPZ [13], etc.
2 Li 2 CO 3 4 e 4 Li + + 2 CO 2 + O 2 3
2 Li 2 CO 3 4 e 4 Li + + 2 CO 2 + O 2 1
  2 Li 2 CO 3 3 e 4 Li + + 2 CO 2 +   O 2
In this work, a planar binuclear cobalt manganese phthalocyanine (bi-CoMnPc, structure shown in Figure S1) is comparatively tested with bi-CoPc to investigate the effects of central metal ions in binuclear metal phthalocyanines on the catalysis of Li2CO3 decomposition. Density functional theory (DFT) calculations indicate that the key intermediate peroxydicarbonate (*C2O62−) is stabilized by bi-CoPc2+ and bi-CoMnPc3+, which is accountable for their excellent catalytic effects. With one central metal ion substituted by manganese for cobalt, the bi-CoMnPc’s second active redox couple for catalyzing Li2CO3 decomposition shifts from the second Co(II)/Co(III) couple in the central metal ion to the Pc(−2)/Pc(−1) couple in the phthalocyanine ring. In artificial dry air, the LAB cell with bi-CoMnPc exhibited significantly improved cyclability than the RM-free one and even the cell with bi-CoPc.

2. Results and Discussion

Cyclic voltammetry (CV) in the Ar atmosphere was conducted to identify the electrochemically active redox couples in bi-CoMnPc. As displayed in Figure 1a, bi-CoMnPc shows a pair of redox peaks (marked as 1a/1c) around 2.69 V vs. Li/Li+, with emphasized shoulders (marked as 1′a/1′c) at slightly higher potentials. These two pairs could be assigned to the Co(I)/Co(II) couple and the Pc(−3)Pc(−2)/Pc(−2)Pc(−2) couple, respectively [6,20−22]. Since their redox potentials are imposed on each other, it is difficult to exactly differentiate them. At around 3.48 V vs. Li/Li+, a pair of small peaks are observed, marked as 2a/2c, which are assigned to the Mn(II)/Mn(III) couple [20,21,22]. Adjacent to that, another pair of redox peaks (marked as 2′a/2′c) appear at about 3.72 V vs. Li/Li+, which is assigned to the Co(II)/Co(III) couple [6,22], similar to that in bi-CoPc (IIa/II′a/IIc in Figure 1a). At 3.9~4.2 V vs. Li/Li+, a pair of broad peaks (marked as 3a/3c) are observed in the reductive and oxidative branches, which could be assigned to the redox of phthalocyanine rings, i.e., the Pc(−2)Pc(−1)/Pc(−2)Pc(−2) and Pc(−1)Pc(−1)/Pc(−2)Pc(−1) couples [6,22], similar to that in bi-CoPc (IIIa/IIIc).
According to the function mechanism of RMs, the active redox couples must have higher potential than the target solid-state discharge (or parasitic) product [6,7,8,9,10,11,12,13,14,15], which is 3.71 V vs. Li/Li+ for Li2CO3 (Equation (3)); therefore, in bi-CoMnPc, the active redox couples should be the Co(II)/Co(III), Pc(−2)Pc(−1)/Pc(−2)Pc(−2) and Pc(−1)Pc(−1)/Pc(−2)Pc(−1) couples. Compared to bi-CoPc, due to the substitution of manganese for cobalt, the redox potential of the central metal ion was lowered by ~340 mV from ~3.82 V (with the Co(II)/Co(III) couple) to ~3.48 V (with the Mn(II)/Mn(III) couple). This should have two consequences: on the one hand, the lowered potential would lose the redox couple’s ability to facilitate Li2CO3 decomposition; on the other hand, it is beneficial to facilitate the decomposition of other solid-state products such as Li2O2, LiOH, and carboxylates, as reported in previous works [23,24,25].
To investigate the feasibility of bi-CoMnPc for catalyzing Li2CO3 decomposition, linear scanning voltammetry (LSV), galvanostatic, and potentiostatic charging tests were conducted for Li2CO3-MWCNT (multiwall carbon nanotube) composite electrodes with bi-CoMnPc, bi-CoPc, or without RM in the electrolyte. As shown in Figure 1b, in the LSV test, without RM in the electrolyte, a small peak (*) emerges at 4.0 V vs. Li/Li+, which could be attributed to Li2CO3 decomposition. When bi-CoPc was added into the electrolyte, the oxidative current density was roughly doubled in the range of 3.75~4.50 V vs. Li/Li+, and the oxidative peaks corresponded well to IIa/II’a and IIIa as in the Li2CO3-free CV (Figure 1a). As for the sample with bi-CoMnPc, it is interesting to observe that the oxidative current density was further doubled than that with bi-CoPc. Furthermore, aside from the oxidative peaks 2a, 2′a, and 3a (as in Figure 1a), a new peak (**) appeared at ~4.05 V vs. Li/Li+, which is presumably attributed to Li2CO3 decomposition. After galvanostatic charging, the preloaded Li2CO3 was almost completely removed for the samples with bi-CoPc or bi-CoMnPc, leaving cavities in the MWCNT matrix, as characterized by X-ray diffraction (XRD) and scanning electronic microscopy (SEM) test, and shown in Figure 1c,d. In contrast, many Li2CO3 particles remained in the RM-free sample. Potentiostatic charging tests were also conducted, as shown in Figure 1b–d. After charging with bi-CoMnPc at 3.85 V, most of the preloaded Li2CO3 still remained in the carbon paper. In sharp contrast, when the charging potential increased to 4.15 V, Li2CO3 was almost completely removed. For comparison, the same tests were conducted with bi-CoPc at 3.75 V and 3.95 V, respectively. In accordance with our previous work [6], Li2CO3 remained at 3.75 V but was completely removed at 3.95 V.
Based on these results, it was demonstrated that, similar to bi-CoPc, bi-CoMnPc can also catalyze Li2CO3 decomposition. However, its catalysis mechanism should be a bit different from that of bi-CoPc because, unlike in bi-CoPc, the second central metal ion (Mn) in bi-CoMnPc is not redox-active for facilitating Li2CO3 decomposition. In our previous work, it was proposed that the RM-catalyzed Li2CO3 decomposition requires two electrons extracted from Li2CO3 to a single RM molecule rather than tandem electron transfer with different RM molecules [6]. Compared to that in bi-CoPc, the absence of the second Co(II)/Co(III) couple obliges the Pc(−2)Pc(−1)/Pc(−2)Pc(−2) couple in bi-CoMnPc to function as the second active redox couple to successfully extract electrons from Li2CO3 to the RM molecule and subsequently to the working electrode, thus raising the available potential for Li2CO3 decomposition.
To further investigate the mechanism of RM-catalyzed Li2CO3 decomposition, we performed DFT calculations for the Li2CO3 decomposition process with bi-CoPc or bi-CoMnPc. Three possible sets of reaction products, namely CO2 and 3O2, CO2 and 1O2, CO2 and O2, were considered.
The most favorable Li2CO3 decomposition pathways on bi-CoPc2+ are illustrated in Figure 2a,b and Equation (4), and the other pathways and corresponding structures are shown in Figures S2–S4. With bi-CoPc2+, upon dissociation of Li2CO3(s) from the solid phase to CO32− and Li+, CO32− adsorbs on the bi-CoPc2+ with both the O-Co and O-C interactions. Next, another molecular CO32− is adsorbed on the *CO32− structure through O-O interaction to generate *C2O64−/bi-CoPc2+. And then *C2O64− loses two electrons to obtain C2O62−, while bi-CoPc2+ receives two electrons to reduce back to neutral bi-CoPc. Compared to the simultaneous transfer of two electrons (*CO32− → *C2O64− → C2O62− companied by bi-CoPc2+ → bi-CoPc), the process of tandem electron transfer, as shown by the black lines in Figure 2a (*CO32− → CO3 →C2O63− → C2O62−) is not energetically favorable, regardless of whether the two electrons are gained by bi-CoPc2+ → bi-CoPc+ → bi-CoPc, 2bi-CoPc2+ → 2bi-CoPc+ or 2bi-CoPc+ → 2bi-CoPc (see Figures S2–S4).
Subsequently, C2O62− adsorbs on another bi-CoPc2+ and desorbs a CO2 molecule to produce *CO42−/bi-CoPc2+. With an additional two electrons transferred from CO42− to bi-CoPc2+, CO2 and 1O2 are accordingly produced, as well as bi-CoPc. Though *CO42− serves as a bifurcation point to lead to three possible sets of reaction products, namely CO2 and 1O2, CO2 and 3O2, and CO2 and O2, the production of the latter two is kinetically disadvantageous.
For the whole process of Li2CO3 decomposition, the crucial elementary reaction with the highest reaction energy is *C2O64−/bi-CoPc2+ → C2O62−/bi-CoPc (∆Gr = 328 kJ/mol). Considering both the thermodynamics and kinetics, the main products of Li2CO3 decomposition correspond to CO2 and 1O2 [6]. The preference for 1O2 production over 3O2 and O2 may help explain the unusually low O2 evolution in Li2CO3 decomposition compared to that in the Li2O2 decomposition scenario.
2 Li 2 CO 3 + 2 bi CoPc 2 + 4 Li + + 2 CO 2 + O 2 1 + 2 bi CoPc   G 1 = 2646   kJ / mol
2 Li 2 CO 3 + 2 bi CoMnPc 3 + 4 Li + + 2 CO 2 + O 2 1 + 2 bi CoMnPc 3 +   G 1 = 3714   kJ / mol
The reaction pathways for carbonate decomposition on bi-CoMnPc3+ are shown in Figure 2c,d and Equation (5), and more detailed information is exhibited in Figures S5–S7. Upon substitution of Mn for Co, more positive charges accumulate on Mn (the NPA charges on Mn and Co in bi-CoMnPc3+ are 1.262 and 0.849, respectively). Therefore, the electrostatic interaction between CO32− and the Mn site is slightly stronger than that between CO32− and the Co site, leading to preferable adsorption of CO32− at the manganese site. Although the initial adsorption site of CO32− at bi-CoMnPc3+ changes to Mn, the energetically most preferred reaction pathways to form soluble C2O62− intermediates are similar (*CO32− → *C2O64− → C2O62− companies by bi-CoMnPc3+ → bi-CoMnPc+). Here, the tandem electron transfer to form C2O62− accompanied by bi-CoMnPc3+ → bi-CoMnPc2+ → bi-CoMnPc+, 2bi-CoMnPc3+ → 2bi-CoMnPc2+, or 2bi-CoMnPc2+ → 2bi-CoMnPc+ are less favorable (see Figures S5–S7). In addition, compared to the decomposition route to generate 3O2 (Equation (1)) or O2 (Equation (3)), the reaction energy for producing 1O2 (Equation (2)) is also much lower, affording CO2 and 1O2 as the main products with 3714 kJ/mol of energy released for the whole decomposition process, which is much higher than that in the bi-CoPc system (∆Gr = -2646 kJ/mol). In this process, the crucial elementary reaction is *C2O64−/bi-CoMnPc3+ → C2O62−/bi-CoMnPc+. Here, the reaction energy for *C2O64− → C2O62− (∆Gr = 713 kJ/mol) is much higher than that in the bi-CoPc system (∆Gr = 328 kJ/mol). According to the experimental results, the decomposition voltage of lithium carbonate in the bi-CoMnPc system is higher than that in the bi-CoPc system.
It should be noted that, in the Li2CO3 decomposition process, the soluble *C2O62− is a crucial intermediate because it can desorb and re-adsorb onto another RM molecule (or ion) or separately diffuse onto the air cathode surface to continue decomposition [8,26]. According to the calculation and discussion above, *C2O62− is stabilized with the facilitation of bi-CoPc+/bi-CoMnPc+/bi-CoMnPc2+, and even more so with bi-CoPc2+/bi-CoMnPc3+, favoring *C2O62− desorption and subsequent adsorption onto fresh RMs (bi-CoPc2+/bi-CoMnPc3+) or diffusion onto the air cathode to enhance the decomposition process. It can thus be deduced that, for metal phthalocyanine-type RMs, 2-electron ones are superior to 1-electron ones in facilitating Li2CO3 decomposition, consistent with our experimental results (Figure 1 and Ref. [6]). And the Li2CO3 decomposition with the bi-CoMnPc/bi-CoPc facilitation can be illustrated in Scheme 1.
It should also be noted that, other than the oxidative decomposition path, Li2CO3 could also be removed through the double decomposition reaction (Equation (6)), which has a much lower equilibrium potential (2.80 V vs. Li/Li+) than Equations (1)–(3) (3.71–3.82 V). There have been several 1-electron RMs reported effective for catalyzing this reaction, such as manganese phthalocyanine (MnPc) [27], ruthenium acetylacetonate (Ru(acac)3) [28], o-phenylenediamine (OPD) [29], phenoxathiine (PHX) [30], etc. Therefore, this reaction should be carefully treated when investigating Li2CO3 decomposition mechanisms to exclude its interference.
2 Li 2 CO 3 + C 4 e   4 Li + + 3 CO 2
To evaluate the effects of bi-CoMnPc on the cell performance and compare it with bi-CoPc, fixed capacity discharge–charge cycling tests were carried out for LAB coin cells with 2.5 mmol L−1 of bi-CoMnPc, with 2.5 mmol L−1 of bi-CoPc, or without RM in the electrolyte. The cells were tested in artificial dry air, which is composed of N2 and O2 (78:22, v/v). The discharge–charge curves are displayed in Figure 3a. With the addition of bi-CoMnPc, the mid-capacity voltage was reduced by ~260 mV from 4.02 V (without RM) to 3.76 V. This is also ~70 mV lower than that with bi-CoPc, which could be attributed to the lower redox potential of the Mn(II)/Mn(III) couple in bi-CoMnPc than that of the Co(II)/Co(III) couple in bi-CoPc, and thus reduced overpotential for Li2O2/LiOH decomposition. Aside from the reduction in charging potential, the RMs also exhibited an effect on elevating the discharge voltage; the discharge plateau was raised from 2.66 V without the RM to 2.71 V with bi-CoMnPc and 2.68 V with bi-CoPc. A similar phenomenon has been observed with other metal phthalocyanine complexes, such as mononuclear cobalt phthalocyanine [6] and iron phthalocyanine [31], which could be attributed to their redox mediation effect for the catalysis of oxygen reduction reactions (ORRs).
With RM addition in the electrolyte, the cyclic life was significantly enhanced from 52 cycles (without RM) to 193 cycles (with bi-CoPc) and 261 cycles (with bi-CoMnPc), as shown in Figure 3b. Ex situ Fourier transform infrared (FTIR) spectroscopy in Figure 3c clearly shows that, after 30 cycles of operation, intense signals of Li2CO3 and lithium carboxylates became observable for the RM-free cell. In contrast, for the cell with bi-CoPc and bi-CoMnPc, the Li2CO3 and lithium carboxylate peaks are greatly suppressed. It is noted that the Li2CO3 and lithium carboxylate peaks are even less prominent for the cell with bi-CoMnPc than that with bi-CoPc, indicating bi-CoMnPc’s superior suppression effect on parasitic product accumulation. To further investigate the accumulation status of parasitic products, ex situ energy dispersive spectroscopy (EDS) mapping was used to semi-quantitatively characterize the cycled air cathodes. As shown in Figures S8–S11, the oxygen content in the air cathodes with bi-CoPc or bi-CoMnPc is roughly only half of the RM-free one, indicating their good capabilities in removing parasitic products.
It has been widely perceived that high charging potential is detrimental to the cyclability of LOBs/LABs because it would induce more parasitic reactions, consuming electrolyte/carbon electrodes and accumulating solid products (such as Li2CO3, lithium carboxylates, and polyesters) on the air cathode [32]. As discussed above, on the one hand, with the substitution of manganese for cobalt, the RM’s functioning potential for Li2CO3 decomposition is raised from ~3.95 V (with bi-CoPc) to ~4.15 V (with bi-CoMnPc), but on the other hand, its functioning potential for Li2O2/LiOH decomposition is reduced by ~260 mV from 3.74 V to 3.48 V. As a result of these competing effects, bi-CoMnPc apparently exhibited better performance in improving the cell cyclability, probably because Li2O2 is still the majority of solid-state discharge products in the CO2-free artificial dry air.

3. Materials and Methods

3.1. Materials

Binuclear cobalt phthalocyanine (bi-CoPc) and binuclear cobalt manganese phthalocyanine (bi-CoMnPc) were purchased from Shanghai Dibai Chemical Technologies Co., Ltd. (Shanghai, China). Tetraethylene glycol dimethyl ether (TEGDME) from Aladdin Scientific (Shanghai, China) was dried for at least 72 h over activated 3 Å molecular sieves. Li2CO3 was ball milled for 6 h at 400 rpm prior to use. All reagents were used as received unless specified.

3.2. Electrochemical Measurements

The cyclic voltammetry (CV), linear sweep voltammetry (LSV), and charging and discharging tests were conducted in CR2032 coin cells. The air cathode was prepared by pasting a slurry of ketjenblack (KB) EC600JD carbon black onto a piece of carbon paper (TGP-H-060, Toray, Tokyo, Japan) with a loading of ∼0.15 mgcarbon cm−2. Polytetrafluoroethylene (PTFE) was used as the binder, with KB/PTFE = 85:15 (w/w). The Li2CO3-preloaded working electrode was prepared by casting a slurry of Li2CO3-MWCNT-PVDF (4:5:1, w/w/w) onto a piece of carbon paper with a loading of ∼0.5 mgcarbon cm−2. A piece of lithium foil (200 μm thick) was used as the anode. The base electrolyte was 1 M lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) in 140 μL of TEGDME. Electrolytes containing bi-CoPc or bi-CoMnPc were prepared by dissolving them into the base electrolyte at a concentration of 2.5 mM. A piece of glass fiber filter (Whatman GF/D, Whatman, Maidstone, UK) and a piece of PP (polypropylene) membrane (Celgard 2400, Celgard, Charlotte, NC, USA) were used as the separators. The coin cells were assembled in an Ar-filled glovebox and relaxed for 12 h, and then the tests were performed either in Ar or in a simulated dry air environment with N2/O2 (78:22, v/v). Potentiostatic and galvanostatic charging tests, as well as cycling tests, were conducted using a Neware battery testing system. The current density was set to 100 mA g−1carbon, with a cutoff voltage of 2.0 V for discharge and 4.50/4.55 V for charge. CV and LSV tests were conducted with a Solartron 1470E electrochemical workstation (Solartron Metrology, Bognor Regis, UK).

3.3. Physical Characterizations

Powder X-ray diffractograms (XRD) were obtained on a D8 ADVANCE DaVinci diffractometer (Bruker, Billerica, MA, USA) with Cu Kα radiation (λ = 1.5418 Å). Fourier transform infrared (FTIR) spectra were collected on a Nicolet 6700 spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) in transmission mode or on a Cary 660+620 spectrometer (Agilent, Santa Clara, CA, USA) in attenuated total reflectance (ATR) mode. Scanning electron microscopy (SEM) was performed on a Hitachi S4800 (Hitachi, Tokyo, Japan) unit or on a Quanta 250 FEG unit (FEI, Hillsboro, OR, USA).

3.4. Computation Methods

The DFT calculation was performed by using the Gaussian 16 [33], xTB (version 6.3.2) [34,35], and ORCA 5.0 [36] programs. Structural optimization and frequency analysis were performed at the GFN2-xTB level as interfaced into the Gaussian 16 program using the gau_xtb code [37] to obtain the thermal correction to Gibbs energy in the gas phase. Stationary points were optimized without symmetry constraints, and their nature was confirmed by vibrational frequency analysis. All structures given are at minimal points and have no imaginary frequencies. Then, more accurate single-point energies were obtained at the ωB97X [38]-D3BJ [39,40]/def2-TZVPP [41] level with ORCA. The Gibbs free energy G was calculated as G = E (single-point energies obtained by wb97X-D3BJ/def2-TZVPP) + E(thermal correction to Gibbs by GFN2-xTB). And ∆G was calculated by ΔG = ΔE + ΔZPE − TΔS, where ΔE, ΔZPE, T, and ΔS indicated the single-point energy change, zero-point energy change, temperature, and entropy change, respectively. In the decomposition pathways, the intermediate label * represents adsorption on the phthalocyanine structure; otherwise, it means in the gas phase state. The total charge of the structures is considered with the change in the adsorption state, which is the sum of the charges of the binuclear metal phthalocyanine and the adsorption structure. The natural bond orbital (NBO) [42,43,44,45,46,47] calculations were performed to obtain further information.

4. Conclusions

The effects of central metal ions in binuclear metal phthalocyanines have been investigated in terms of their catalytic activity towards Li2CO3 decomposition. DFT calculations indicate that the key intermediate peroxydicarbonate (*C2O62−) is stabilized by bi-CoPc2+ and bi-CoMnPc2+, which is accountable for their improved catalytic activity compared to bi-CoPc+/bi-CoMnPc+ and pristine bi-CoPc/bi-CoMnPc. With one central metal ion substituted by manganese for cobalt, the bi-CoMnPc’s second active redox couple for catalyzing Li2CO3 decomposition shifts from the second Co(II)/Co(III) couple in the central metal ion to the Pc(-2)/Pc(-1) couple in the phthalocyanine ring. In artificial dry air, the LAB cell with bi-CoMnPc in electrolyte exhibited 261 cycles under a fixed capacity of 500 mAh g−1carbon and current density of 100 mA g−1carbon, significantly better than the RM-free cell (62 cycles) and even the cell with bi-CoPc (193 cycles). We hope this work sheds new insights into the mechanisms of RM-facilitated Li2CO3 decomposition and designs more efficient Li2CO3 decomposing RMs.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules29092034/s1, Figure S1: The structures of bi-CoPc and bi-CoMnPc. Figure S2: The Li2CO3 decomposition pathways on bi-CoPc2+ and corresponding structures with bi-CoPc2+→bi-CoPc. Figure S3: The Li2CO3 decomposition pathways on pristine bi-CoPc2+ and the corresponding structures with 2bi-CoPc2+→2bi-CoPc+. Figure S4: The Li2CO3 decomposition pathways on bi-CoPc+ and the corresponding structures with 2bi-CoPc+→2bi-CoPc. Figure S5: The Li2CO3 decomposition pathways on pristine bi-CoMnPc3+ and the corresponding structures with bi-CoMnPc3+→bi-CoMnPc+. Figure S6: The Li2CO3 decomposition pathways on pristine bi-CoMnPc3+ and the corresponding structures with 2bi-CoMnPc3+→2bi-CoMnPc2+. Figure S7: The Li2CO3 decomposition pathways on pristine bi-CoMnPc2+ and the corresponding structures with 2bi-CoMnPc2+→2bi-CoMnPc+. Figure S8: The SEM images and EDS mapping results of carbon and oxygen for the air cathode (a) without RM, (b) with bi-CoPc and (c) with bi-CoMnPc after 1 discharge-charge cycle. Figure S9: The SEM images and EDS mapping results of carbon and oxygen for the air cathode (a) without RM, (b) with bi-CoPc and (c) with bi-CoMnPc after 10 discharge-charge cycles. Figure S10: The SEM images and EDS mapping results of carbon and oxygen for the air cathode (a) without RM, (b) with bi-CoPc and (c) with bi-CoMnPc after 30 discharge-charge cycles. Figure S11: The oxygen content in the cycled air cathodes according to EDS results.

Author Contributions

Conceptualization, Z.L.; methodology, Z.L. and S.Z.; investigation, Q.Y. and L.Y.; data curation, Q.Y. and L.Y.; validation, Q.Y., L.Y. and H.H. (Haoshen Huang); formal analysis, Z.L. and S.Z.; writing—original draft preparation, Q.Y., L.Y., Z.L. and S.Z.; writing—review and editing, Z.C. and H.H. (Haiyong He); funding acquisition, Z.L., S.Z. and Z.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Key Research and Development Program of Zhejiang Province (grant number 2023C01102), the Ningbo S&T Innovation 2025 Major Special Program (grant number 2022Z022), and the Ningbo Science and Technology Project (grant number 2022-DST-004).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article and Supplementary Materials.

Acknowledgments

The authors would like to thank Jiahao Song and Junwen Zheng for their assistance in manuscript review and editing.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kwak, W.-J.; Rosy; Sharon, D.; Xia, C.; Kim, H.; Johnson, L.R.; Bruce, P.G.; Nazar, L.F.; Sun, Y.-K.; Frimer, A.A.; et al. Lithium-Oxygen Batteries and Related Systems: Potential, Status, and Future. Chem. Rev. 2020, 120, 6626–6683. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, T.; Vivek, J.P.; Zhao, E.W.; Lei, J.; Garcia-Araez, N.; Grey, C.P. Current Challenges and Routes Forward for Nonaqueous Lithium–Air Batteries. Chem. Rev. 2020, 120, 6558–6625. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, K.; Yang, D.-Y.; Huang, G.; Zhang, X.-B. Lithium–Air Batteries: Air-Electrochemistry and Anode Stabilization. Acc. Chem. Res. 2021, 54, 632–641. [Google Scholar] [CrossRef] [PubMed]
  4. Kang, J.-H.; Lee, J.; Jung, J.-W.; Park, J.; Jang, T.; Kim, H.-S.; Nam, J.-S.; Lim, H.; Yoon, K.R.; Kim, I.-D.; et al. Lithium–Air Batteries: Air-Breathing Challenges and Perspective. ACS Nano 2020, 14, 14549–14578. [Google Scholar] [CrossRef] [PubMed]
  5. Zhao, Z.; Huang, J.; Peng, Z. Achilles’ Heel of Lithium–Air Batteries: Lithium Carbonate. Angew. Chem. Int. Ed. 2017, 57, 3874–3886. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, Z.; Zhang, Y.; Jia, C.; Wan, H.; Peng, Z.; Bi, Y.; Liu, Y.; Peng, Z.; Wang, Q.; Li, H.; et al. Decomposing lithium carbonate with a mobile catalyst. Nano Energy 2017, 36, 390–397. [Google Scholar] [CrossRef]
  7. Wang, X.-G.; Wang, C.; Xie, Z.; Zhang, X.; Chen, Y.; Wu, D.; Zhou, Z. Improving Electrochemical Performances of Rechargeable Li−CO2 Batteries with an Electrolyte Redox Mediator. ChemElectroChem 2017, 4, 2145–2149. [Google Scholar] [CrossRef]
  8. Mota, F.M.; Kang, J.-H.; Jung, Y.; Park, J.; Na, M.; Kim, D.H. Mechanistic Study Revealing the Role of the Br3-/Br2 Redox Couple in CO2-Assisted Li–O2 Batteries. Adv. Energy Mater. 2020, 10, 1903486. [Google Scholar] [CrossRef]
  9. Chen, J.; Zou, K.; Ding, P.; Deng, J.; Zha, C.; Hu, Y.; Zhao, X.; Wu, J.; Fan, J.; Li, Y. Conjugated Cobalt Polyphthalocyanine as the Elastic and Reprocessable Catalyst for Flexible Li–CO2 Batteries. Adv. Mater. 2019, 31, 1805484. [Google Scholar] [CrossRef]
  10. Huang, S.; Li, Z.; Liu, Z.; Yan, Q.; Ma, B.; Wang, D.; Wei, F.; Chen, Z.; He, H. Surface Enrichment of Redox Mediator for Long-Cyclable Lithium−Air Batteries. Energy Fuels 2023, 37, 11465–11471. [Google Scholar] [CrossRef]
  11. Zhu, C.; Wang, Y.; Shuai, L.; Tang, Y.; Qiu, M.; Xie, J.; Liu, J.; Wen, W.; Chen, H.; Nan, S.; et al. Remarkable improvement of cyclic stability in Li–O2 batteries using ruthenocene as a redox mediator. Chin. Chem. Lett. 2020, 31, 1997–2002. [Google Scholar] [CrossRef]
  12. Wan, H.; Sun, Y.; Yu, J.; Shi, Q.; Zhu, Y.; Qian, Y. A chiral salen-Co(II) complex as soluble redox mediator for promoting the electrochemical performance of Li-O2 batteries. Nano Res. 2022, 15, 8101–8108. [Google Scholar] [CrossRef]
  13. Ko, S.; Yoo, Y.; Choi, J.; Lim, H.-D.; Park, C.B.; Lee, M. Discovery of organic catalysts boosting lithium carbonate decomposition toward ambient air operational lithium–air batteries. J. Mater. Chem. A 2022, 10, 20464. [Google Scholar] [CrossRef]
  14. Pipes, R.; Bhargav, A.; Manthiram, A. Phenyl Disulfide Additive for Solution-Mediated Carbon Dioxide Utilization in Li–CO2 Batteries. Adv. Energy Mater. 2019, 9, 1900453. [Google Scholar] [CrossRef]
  15. Xiong, Q.; Huang, G.; Zhang, X.-B. High-Capacity and Stable Li-O2 Batteries Enabled by a Trifunctional Soluble Redox Mediator. Angew. Chem. Int. Ed. 2020, 59, 19311–19319. [Google Scholar] [CrossRef]
  16. Jiang, F.; Ma, L.; Sun, J.; Guo, L.; Peng, Z.; Cui, Z.; Li, Y.; Guo, X.; Zhang, T. Deciphering the Enigma of Li2CO3 Oxidation Using a Solid-State Li–Air Battery Configuration. ACS Appl. Mater. Interfaces 2021, 13, 14321–14326. [Google Scholar] [CrossRef]
  17. Yang, S.; He, P.; Zhou, H. Exploring the electrochemical reaction mechanism of carbonate oxidation in Li–air/CO2 battery through tracing missing oxygen. Energy Environ. Sci. 2016, 9, 1650–1654. [Google Scholar] [CrossRef]
  18. Mahne, N.; Renfrew, S.E.; McCloskey, B.D.; Freunberger, S.A. Electrochemical Oxidation of Lithium Carbonate Generates Singlet Oxygen. Angew. Chem. Int. Ed. 2018, 57, 5529–5533. [Google Scholar] [CrossRef]
  19. Cao, D.; Tan, C.; Chen, Y. Oxidative decomposition mechanisms of lithium carbonate on carbon substrates in lithium battery chemistries. Nat. Commun. 2022, 13, 4908. [Google Scholar] [CrossRef]
  20. Lever, A.B.P.; Minor, P.C.; Wilshire, J.P. Electrochemistry of Manganese Phthalocyanine in Nonaqueous Media. Inorg. Chem. 1981, 20, 2550–2553. [Google Scholar] [CrossRef]
  21. Lin, C.-L.; Lee, C.-C.; Ho, K.-C. Spectroelectrochemical studies of manganese phthalocyanine thin films for applications in electrochromic devices. J. Electroanal. Chem. 2002, 524−525, 81–89. [Google Scholar] [CrossRef]
  22. Kobayashi, N.; Lam, H.; Nevin, W.A.; Janda, P.; Leznoff, C.C.; Koyama, T.; Monden, A.; Shirai, H. Synthesis, spectroscopy, electrochemistry, spectroelectrochemistry, Langmuir-Blodgett film formation, and molecular orbital calculations of planar binuclear phthalocyanines. J. Am. Chem. Soc. 1994, 116, 879–890. [Google Scholar] [CrossRef]
  23. Lim, H.-D.; Lee, B.; Zheng, Y.; Hong, J.; Kim, J.; Gwon, H.; Ko, Y.; Lee, M.; Cho, K.; Kang, K. Rational design of redox mediators for advanced Li–O2 batteries. Nat. Energy 2016, 1, 16066. [Google Scholar] [CrossRef]
  24. Chen, Y.; Gao, X.; Johnson, L.R.; Bruce, P.G. Kinetics of lithium peroxide oxidation by redox mediators and consequences for the lithium–oxygen cell. Nat. Commun. 2018, 9, 767. [Google Scholar] [CrossRef] [PubMed]
  25. Ahn, S.; Zor, C.; Yang, S.; Lagnoni, M.; Dewar, D.; Nimmo, T.; Chau, C.; Jenkins, M.; Kibler, A.J.; Pateman, A.; et al. Why charging Li–air batteries with current low-voltage mediators is slow and singlet oxygen does not explain degradation. Nat. Chem. 2023, 15, 1022–1029. [Google Scholar] [CrossRef] [PubMed]
  26. Qiao, Y.; Yi, J.; Guo, S.; Sun, Y.; Wu, S.; Liu, X.; Yang, S.; He, P.; Zhou, H. Li2CO3-free Li–O2/CO2 battery with peroxide discharge product. Energy Environ. Sci. 2018, 11, 1211–1217. [Google Scholar] [CrossRef]
  27. Kim, B.; Shin, K.; Henkelman, G.; Ryu, W.-H. CO2-mediated porphyrin catalysis in reversible Li-CO2 cells. Chem. Eng. J. 2023, 477, 147141. [Google Scholar] [CrossRef]
  28. Lian, Z.; Lu, Y.; Ma, S.; Wang, L.; Li, Z.; Liu, Q. An integrated strategy for upgrading Li-CO2 batteries: Redox mediator and separator modification. Chem. Eng. J. 2022, 450, 138400. [Google Scholar] [CrossRef]
  29. Wang, L.; Lu, Y.; Ma, S.; Lian, Z.; Gu, X.; Li, J.; Li, Z.; Liu, Q. Optimizing CO2 reduction and evolution reaction mediated by o-phenylenediamine toward high performance Li-CO2 battery. Electrochim. Acta 2022, 419, 140424. [Google Scholar] [CrossRef]
  30. Cao, D.; Liu, X.; Yuan, X.; Yu, F.; Chen, Y. Redox Mediator-Enhanced Performance and Generation of Singlet Oxygen in Li−CO2 Batteries. ACS Appl. Mater. Interfaces 2021, 13, 39341–39346. [Google Scholar] [CrossRef]
  31. Sun, D.; Shen, Y.; Zhang, W.; Yu, L.; Yi, Z.; Yin, W.; Wang, D.; Huang, Y.; Wang, J.; Wang, D.; et al. A Solution-Phase Bifunctional Catalyst for Lithium–Oxygen Batteries. J. Am. Chem. Soc. 2014, 136, 8941–8946. [Google Scholar] [CrossRef] [PubMed]
  32. Ottakam Thotiyl, M.M.; Freunberger, S.A.; Peng, Z.; Bruce, P.G. The Carbon Electrode in Nonaqueous Li–O2 Cells. J. Am. Chem. Soc. 2013, 135, 494–500. [Google Scholar] [CrossRef]
  33. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  34. Grimme, S.; Bannwarth, C.; Shushkov, P. A Robust and Accurate Tight-Binding Quantum Chemical Method for Structures, Vibrational Frequencies, and Noncovalent Interactions of Large Molecular Systems Parametrized for All spd-Block Elements (Z = 1–86). J. Chem. Theory Comput. 2017, 13, 1989–2009. [Google Scholar] [CrossRef] [PubMed]
  35. Bannwarth, C.; Ehlert, S.; Grimme, S. GFN2-xTB—An Accurate and Broadly Parametrized Self-Consistent Tight-Binding Quantum Chemical Method with Multipole Electrostatics and Density-Dependent Dispersion Contributions. J. Chem. Theory Comput. 2019, 15, 1652–1671. [Google Scholar] [CrossRef] [PubMed]
  36. Neese, F. The ORCA program system. WIREs Comput. Mol. Sci. 2011, 2, 73–78. [Google Scholar] [CrossRef]
  37. Lu, T. Gau_Xtb: A Gaussian Interface for Xtb Code. Available online: https://sobereva.com/soft/gau_xtb (accessed on 15 August 2023).
  38. Chai, J.-D.; Head-Gordon, M. Systematic optimization of long-range corrected hybrid density functionals. J. Chem. Phys. 2008, 128, 084106. [Google Scholar] [CrossRef] [PubMed]
  39. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef] [PubMed]
  40. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  41. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  42. Reed, A.E.; Curtiss, L.A.; Weinhold, F. Intermolecular Interactions from a Natural Bond Orbital, Donor-Acceptor Viewpoint. Chem. Rev. 1988, 88, 899–926. [Google Scholar] [CrossRef]
  43. Carpenter, J.E.; Weinhold, F.J. Analysis of the geometry of the hydroxymethyl radical by the “different hybrids for different spins” natural bond orbital procedure. J. Mol. Struct. Theochem. 1988, 46, 41–62. [Google Scholar] [CrossRef]
  44. Reed, A.E.; Weinstock, R.B.; Weinhold, F.J. Natural Atomic Orbitals and Natural Population Analysis. J. Chem. Phys. 1985, 83, 735–746. [Google Scholar] [CrossRef]
  45. Reed, A.E.; Weinhold, F.J. Natural Localized Molecular Orbitals. J. Chem. Phys. 1985, 83, 1736–1740. [Google Scholar] [CrossRef]
  46. Reed, A.E.; Weinhold, F.J. Natural Bond Orbital Analysis of Near-Hartree-Fock Water Dimer. J. Chem. Phys. 1983, 78, 4066–4073. [Google Scholar] [CrossRef]
  47. Foster, J.P.; Weinhold, F.J. Natural Hybrid Orbitals. J. Am. Chem. Soc. 1980, 102, 7211–7218. [Google Scholar] [CrossRef]
Figure 1. (a) Cyclic voltammograms of bi-CoPc and bi-CoMnPc in Ar atmosphere; (b) LSV curves of the Li2CO3-MWCNT composite electrodes with bi-CoMnPc, bi-CoPc, or without RM in the electrolyte, the asterisk (* and **) signs mark the oxidative peaks presumably attributed to Li2CO3 decomposition; (c) XRD patterns of the Li2CO3-MWCNT composite electrodes after galvanostatic charging (marked as CC) to 4.55 V, or after potentiostatic charging at 3.75/3.95 V vs. Li/Li+ with bi-CoPc, at 3.85/4.15 V with bi-CoMnPc, or at 4.15 V without RM, the asterisk (*) sign marks the unknown peak; (d) SEM images of the galvanostatically or potentiostatically charged Li2CO3-MWCNT composite electrodes.
Figure 1. (a) Cyclic voltammograms of bi-CoPc and bi-CoMnPc in Ar atmosphere; (b) LSV curves of the Li2CO3-MWCNT composite electrodes with bi-CoMnPc, bi-CoPc, or without RM in the electrolyte, the asterisk (* and **) signs mark the oxidative peaks presumably attributed to Li2CO3 decomposition; (c) XRD patterns of the Li2CO3-MWCNT composite electrodes after galvanostatic charging (marked as CC) to 4.55 V, or after potentiostatic charging at 3.75/3.95 V vs. Li/Li+ with bi-CoPc, at 3.85/4.15 V with bi-CoMnPc, or at 4.15 V without RM, the asterisk (*) sign marks the unknown peak; (d) SEM images of the galvanostatically or potentiostatically charged Li2CO3-MWCNT composite electrodes.
Molecules 29 02034 g001
Figure 2. (a) The most favorable Li2CO3 decomposition pathways on bi-CoPc2+; (b) the corresponding structure evolution of Li2CO3 decomposition on bi-CoPc2+; (c) the most favorable Li2CO3 decomposition pathways on bi-CoMnPc3+; (d) the corresponding structure evolution of Li2CO3 decomposition on bi-CoMnPc3+.
Figure 2. (a) The most favorable Li2CO3 decomposition pathways on bi-CoPc2+; (b) the corresponding structure evolution of Li2CO3 decomposition on bi-CoPc2+; (c) the most favorable Li2CO3 decomposition pathways on bi-CoMnPc3+; (d) the corresponding structure evolution of Li2CO3 decomposition on bi-CoMnPc3+.
Molecules 29 02034 g002
Scheme 1. The Li2CO3 decomposition process involving bi-CoMnPc or bi-CoPc facilitation.
Scheme 1. The Li2CO3 decomposition process involving bi-CoMnPc or bi-CoPc facilitation.
Molecules 29 02034 sch001
Figure 3. (a) The discharge–charge curves of LABs with bi-CoPc, bi-CoMnPc, or without RM in the first cycle, under a current density of 100 mA g−1carbon and a fixed capacity of 500 mAh g−1carbon, voltage range 2.0–4.55 V; (b) the evolution of capacity retention and terminal discharge voltage for the LABs with bi-CoPc, bi-CoMnPc, or without RM. The arrows are used to assign the terminal discharge voltage and specific capacity to different vertical axes; (c) the FTIR spectra of the air cathodes after 30 discharge–charge cycles with bi-CoPc, bi-CoMnPc, or without RM.
Figure 3. (a) The discharge–charge curves of LABs with bi-CoPc, bi-CoMnPc, or without RM in the first cycle, under a current density of 100 mA g−1carbon and a fixed capacity of 500 mAh g−1carbon, voltage range 2.0–4.55 V; (b) the evolution of capacity retention and terminal discharge voltage for the LABs with bi-CoPc, bi-CoMnPc, or without RM. The arrows are used to assign the terminal discharge voltage and specific capacity to different vertical axes; (c) the FTIR spectra of the air cathodes after 30 discharge–charge cycles with bi-CoPc, bi-CoMnPc, or without RM.
Molecules 29 02034 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yan, Q.; Yan, L.; Huang, H.; Chen, Z.; Liu, Z.; Zhou, S.; He, H. Effects of Central Metal Ion on Binuclear Metal Phthalocyanine-Based Redox Mediator for Lithium Carbonate Decomposition. Molecules 2024, 29, 2034. https://doi.org/10.3390/molecules29092034

AMA Style

Yan Q, Yan L, Huang H, Chen Z, Liu Z, Zhou S, He H. Effects of Central Metal Ion on Binuclear Metal Phthalocyanine-Based Redox Mediator for Lithium Carbonate Decomposition. Molecules. 2024; 29(9):2034. https://doi.org/10.3390/molecules29092034

Chicago/Turabian Style

Yan, Qinghui, Linghui Yan, Haoshen Huang, Zhengfei Chen, Zixuan Liu, Shaodong Zhou, and Haiyong He. 2024. "Effects of Central Metal Ion on Binuclear Metal Phthalocyanine-Based Redox Mediator for Lithium Carbonate Decomposition" Molecules 29, no. 9: 2034. https://doi.org/10.3390/molecules29092034

Article Metrics

Back to TopTop