2022 年 90 巻 10 号 p. 102002
This comprehensive paper, Electrode Potentials Part 2, is a continuation of Electrode Potentials Part 1: Fundamentals and Aqueous Systems. Determining the electrode potential is crucial for understanding the nature of the electrochemical properties of materials or systems; however, an accurate evaluation of the potential of a target electrode has always been a challenge. The electrode potential can be used to predict the reaction mechanisms in electrochemistry and can be directly applied to the study of electrochemical applications. This paper introduces the methodologies and strategies for measuring electrode potentials in nonaqueous and solid-state electrolytes, including organic solvent electrolytes, ionic liquid electrolytes, and oxide and sulfide solid electrolytes. Experimental details are described for basic to state-of-the-art strategies, focusing on practical methods and know-how.
The demand for electrochemical energy storage systems, such as secondary batteries and supercapacitors, increases with each passing year.1,2 “Organic electrolytes” (organic solvents containing ionic species), have become standard for secondary batteries. Additionally, “ionic liquids” (liquid state salts consisting solely of cations and anions)3–5 and “solid-state electrolytes” (oxides, sulfides, and polyanions)6–9 are attracting more attention as promising electrolytes for next-generation batteries with high safety and performance.10–12 However, measuring the precise potential of an electrode in these electrolytes has always been a challenge.
This comprehensive paper is a continuation of Electrode Potentials Part 1: Fundamentals and Aqueous Systems. Part 1 covers the fundamentals of electrode potentials based on their thermodynamic background with related materials and discusses the issues with electrode potentials in aqueous systems (potential–pH diagram, potential windows, practical reference electrodes, and mixed potentials). Part 2 focuses on the methodologies and strategies for measuring electrode potentials in nonaqueous and solid-state electrolytes. The first section of this paper discusses the general properties of common reference electrodes in nonaqueous electrolytes and their experimental and preparation methods. The second section introduces the concept of advanced reference electrodes using metal alloys and two-phase insertion-type compounds. The final section presents the experimental strategies and methods for measuring the electrode potential in a solid-state electrolyte. Detailed slides for each section are provided in the Supplementary Material.
Reference electrodes utilizing the Ag+/Ag redox couple (hereafter “Ag reference electrode”) are widely applied for electrochemical measurements in nonaqueous systems due to their high stability and easy handling.
| \begin{equation} \text{Ag$^{+}$} + \text{e$^{-}$} \rightleftharpoons \text{Ag} \end{equation} | (1) |

Schematic illustrations of selected electrodes for electrochemical measurements.
Electrochemical measurements using Ag reference electrodes are both useful and convenient. However, the universality of the obtained potentials should be considered when comparing the results from various systems, because the Ag+/Ag redox potential is known to significantly depend on the solvent.13,14 According to the IUPAC recommendation,15 the ferrocenium/ferrocene (Fc+/Fc) redox couple is the universal standard because the Fc+/Fc potential is considered to be independent of solvents owing to their bulkiness. The potential difference in various systems can be discussed on a common scale with the introduction of the Fc+/Fc redox couple.
| \begin{equation} \text{Fc$^{+}$} + \text{e$^{-}$} \rightleftharpoons \text{Fc} \end{equation} | (2) |
In general, assuming a cyclic voltammetric measurement for a reversible reaction, the potential difference (ΔEp) between anodic (Epa) and cathodic (Epc) peaks is expressed as follows:16
| \begin{equation} \Delta E_{\text{p}} = E_{\text{pa}} - E_{\text{pc}} = \frac{\Delta \widetilde{E}_{\text{p}}}{n} \end{equation} | (3) |
| \begin{equation} E_{1/2}^{\text{r}} = E^{\circ\prime} + \frac{RT}{nF}\ln \sqrt{\frac{D_{\text{R}}}{D_{\text{O}}}} \end{equation} | (4) |
| \begin{equation} E^{\circ\prime} = E_{1/2}^{\text{r}} \approx \frac{E_{\text{pa}} + E_{\text{pc}}}{2} \end{equation} | (5) |
| \begin{equation} E_{\text{vs. Fc${^{+}}$/Fc}} = E_{\text{vs. Ag${^{+}}$/Ag}} - E^{\circ\prime}(\text{Fc$^{+}$/Fc}) \end{equation} | (6) |
| \begin{equation} i_{\text{pa}} = 0.4463\sqrt{\frac{n^{3}F^{3}}{RT}} C_{\text{R}}\sqrt{D_{\text{R}}v} \end{equation} | (7) |
It should be noted that the reversible systems involved with multielectron reactions (n ≥ 2) usually show the deviation from the theory mentioned above because the Eqs. 3, 4, and 7 can be applied only for one-step n-electron reaction. Since most of multielectron reactions are practically composed of n-step one-electron reactions, the statistical (entropic) factors should be considered for calculation of the standard potentials.16,18 Assuming the systems of n equivalent, independent, and reversible active centers for one-electron transfer having the same standard potentials (E°′), the entropic factor leads to the following equation for the redox potential of the jth one-electron transfer step ($E^{\circ \prime}_{j}$).18
| \begin{equation} E_{j}^{\circ\prime} = E^{\circ\prime} - \frac{RT}{F}\ln \left(\frac{j}{n - j + 1} \right) \end{equation} | (8) |
| \begin{equation} \Delta E^{\circ\prime} = E_{n}^{\circ\prime} - E_{1}^{\circ\prime} = -\frac{2RT}{F}\ln n \end{equation} | (9) |
Various metals are also applied as reference electrodes for systems utilizing metal ions as charge carriers because these systems often allow the reversible electrochemical deposition/dissolution of metals as follows:
| \begin{equation} \text{M$^{n+}$} + \text{$n$e$^{-}$} \rightleftharpoons \text{M} \end{equation} | (10) |
Nevertheless, the handling of metal reference/counter electrodes largely affects the electrochemical behavior, leading to the misinterpretation of obtained results. In the case of the two-electrode configurations, since the degree of the overpotential of the metal counter electrodes changes with the temperature, the current density, and the scan rate of the potential (voltage), it is advisable to check the overpotential in the corresponding experimental conditions with symmetric M/M cells, as described in Part 1. In three-electrode cells with metal reference electrodes, most electrolytes including alkali metal-ion systems provide relatively stable electrode potentials that are sufficient for electrochemical measurements. However, some systems show poor reversibility and large overpotentials of the metal deposition/dissolution reactions due to the formation of passivating films, leading to the instability of the counter/reference electrode potentials. Even in the reversible alkali metal-ion systems, the surface film, referred to as the “solid electrolyte interphase (SEI)”, is formed on the alkali metal counter/reference electrodes, and their potentials shift from the ideal values because most electrolytes are intrinsically (thermodynamically) unstable toward the highly reactive alkali metal electrodes. In addition, the interfacial resistance of alkali metals usually increases with an increase in the period of contact with electrolytes.20 Since the degree of the potential deviation depends on the stability of the electrolyte components, it is preferable to measure the potential difference between fresh and old metals and evaluate the surface resistance by impedance spectroscopy.
Research and development activities of post-lithium-ion batteries such as sodium-ion and potassium-ion ones are attracting increasingly more attention.4,21–23 Most researchers discussed the latent capabilities of operating voltages based on the difference in famous alkali metal electrode potentials in aqueous solutions,14,24 and they concluded that the expected operating voltage of sodium-ion batteries is 0.3 V smaller than that of lithium-ion batteries. However, the potential difference of 0.3 V can only be applied to aqueous solutions and is fundamentally different in nonaqueous systems. In general, the standard potentials of Mn+/M redox couples in nonaqueous systems can be converted from an aqueous system as follows:14
| \begin{equation} E^{\circ} (\text{non-aq}) = E^{\circ}(\text{aq}) + \frac{\Delta G_{\text{t}}^{\circ}}{nF} \end{equation} | (11) |
| Solvent* | $\text{H}^{ + }/\frac{1}{2}\text{H}_{2}$ | Li+/Li | Na+/Na | K+/K | Ag+/Ag |
|---|---|---|---|---|---|
| Water | 0.000 | −3.040 | −2.714 | −2.936 | 0.799 |
| PC | 0.52 | −2.79 | −2.56 | −2.88 | 0.99 |
| MeCN | 0.48 | −2.73 | −2.56 | −2.88 | 0.56 |
| EtOH | 0.12 | −2.93 | −2.57 | −2.77 | 0.85 |
| DMF | −0.19 | −3.14 | −2.81 | −3.04 | 0.58 |
| NMP | −0.26 | −3.40 | −2.87 | −3.05 | 0.53 |
| DMSO | −0.20 | −3.20 | −2.85 | −3.07 | 0.44 |
*PC = propylene carbonate, MeCN = acetonitrile, EtOH = ethanol, DMF = N,N-dimethylformamide, NMP = N-methylpyrrolidone, DMSO = dimethyl sulfoxide.
The electrode potential is a key metric for evaluating electrode materials for electrochemical measurements. As described in Sections 2.1 and 2.2, for nonaqueous systems, Ag and ferrocene reference electrodes are often used in a three-electrode configuration, and alkali metal reference electrodes are used in two-electrode cells (mostly coin cells) (Fig. 2). These two different cell configurations have clear advantages and disadvantages (see Section 2.4 in Part 1 for details about the three-electrode cell); for example, the potential of the reference electrode in the three-electrode cell is practically constant, whereas that in the two-electrode cell is not. Nevertheless, for the sake of convenience, the two-electrode cell is often selected for the study of batteries. However, alkali metals used as reference electrodes can cause problems and are far from the ideal reference electrode under certain conditions. These references frequently exhibit high reactivity toward electrolytes, which may cause a shift from the ideal potential by forming a passivation layer on the surface of the alkali metals, and the overpotential is amplified as the current density increases. Moreover, the experimental temperature ranges are restricted by the melting point of metals (Li: 179 °C, Na: 97.8 °C, K: 63.7 °C).27

Schematic illustrations of (a) two-electrode cell and (b) three-electrode cell configurations for electrochemical measurements.
In this regard, alternative reference electrodes using metal alloys and two-phase insertion-type compounds are attracting attention. The potentials of alternative reference electrodes for lithium-ion batteries are summarized in Fig. 3.28 Metal alloy materials of Li–Sn, Li–Al, Li–Bi, and Li–Au were selected in some studies.29–33 Their potentials are relatively more stable than those of alkali metals during electrochemical measurements for reasonable periods. These metal alloy references are often used for electrochemical impedance spectroscopy and potentiometry in lithium-ion and solid-state batteries (see Section 3.2). However, additional lithiation processes are necessary for these metal alloys to maintain a stable chemical composition. Furthermore, the temperature, current densities, and durations for alloy formation also strongly influence both the chemical composition and surface morphology of the metal alloy.28,34 Thus, attempts have been made to use the equilibrium state of two-phase insertion-type compounds such as olivine LiFePO4, spinel Li4Ti5O12, and NASICON Na3V2(PO4)3.34–38 These materials with flat plateaus during charge–discharge are considered suitable for use as reference electrodes.

Comparison of potentials and polarization for Li, alloy materials, and advanced reference electrodes for lithium-ion batteries. Reproduced with permission under the Creative Commons Attribution License (CC BY).28
The flat potential in a two-phase electrochemical transition can be derived from the Gibbs energy as follows. According to the Nernst equation described in Part 1, the voltage (V) of an electrochemical cell (half cell with Li metal counter electrode) of lithium-ion battery is related to the chemical potential, as shown in Eq. 12:39,40
| \begin{equation} V = -\frac{\mu_{\text{Li}} - \mu_{\text{Li}}^{\text{metal}}}{e} \end{equation} | (12) |
| \begin{equation} \mu_{\text{Li}} = \frac{\partial g}{\partial x} \end{equation} | (13) |

(a) Gibbs energy and (b) voltage profiles of two-phase transition materials. (c) Concept of pseudo reference/counter electrode utilizing the flat plateau of two-phase transition materials of Na3V2(PO4)3–NaV2(PO4)3. Reproduced with permission.35 Copyright 2021, Royal Society of Chemistry.
Thus, the flat plateau region associated with a two-phase transition during the Li- or Na-ion insertion/desertion processes could behave as a nonpolarizable electrode with a small overpotential (Fig. 4c for the Na3V2(PO4)3–NaV2(PO4)3 electrode as an example).35 Although there are several attractive points for such reference electrodes, a practical drawback is the preparation of a partially charged state of these materials. These materials must be charged first to obtain the two-phase state to reach the flat plateau region. Thus far, the electrochemical method (charge in a half-cell configuration) has been applied to reach the partially charged state. However, the electrochemical process is time-consuming and unsuitable for mass production. Recently, an oxidation process was conducted using Cl2 gas to extract Na+ from Na3V2(PO4)3. This method does not harm the morphology of pristine materials and is suitable for mass production. Electrochemical measurements were carried out using a Na3V2(PO4)3–NaV2(PO4)3 two-phase (3.4 V Na+/Na) counter electrode combined with working electrode materials of Na2FeP2O7, Na3V2(PO4)3, NaCrO2, and hard carbon in organic and ionic liquid electrolytes. The Na3V2(PO4)3–NaV2(PO4)3 electrode presented an accurate charge–discharge potential of the working electrodes even at high temperatures (>100 °C) at which Na metal melts. Moreover, the Na3V2(PO4)3–NaV2(PO4)3 electrode exhibited high electrochemical stability during long cycles and low polarization compared with the Na metal electrode.
Computational methods can also be used to calculate an electrode potential based on the basic principle that the equilibrium voltage difference between positive and negative electrodes depends on the difference in their Li chemical potentials, as shown in Eqs. 12 and 13. Density functional theory (DFT) calculations have demonstrated that the electrode potentials for alkali metal batteries can be calculated from the change in the total energy.41,42 For example, the electrode potential of LiCoO2 vs. Li+/Li can be calculated from Eq. 14.
| \begin{equation} V = -\frac{[E_{\text{LiCoO}_{2}} - E_{\text{Li${_{(1 - n)}}$CoO${_{2}}$}} - nE_{\text{Li}}]}{n} \end{equation} | (14) |
All-solid-state batteries consist of only solid components. Solid electrolytes are expected to have various advantages, including high power, improved safety, and long lifetimes. As a representative solid-state system, this section outlines the basics of all-solid-state batteries, reference electrodes, and potential-window measurements of solid electrolytes.
Figure 5 shows a schematic of the structure of an all-solid-state lithium-ion battery. The battery has a layered structure consisting of a positive electrode current collector, positive electrode composite layer, separator layer (solid electrolyte layer), negative electrode composite layer, and negative electrode current collector. The positive electrode composite consists of a positive electrode active material, a solid electrolyte, a binder, and a conductive additive. Similarly, the negative electrode composite consists of a negative electrode active material, a solid electrolyte, a binder, and a conductive additive. The active material stores electricity through oxidation and reduction via Li extraction and insertion, respectively.

Schematic illustration of an all-solid-state battery.
Positive electrode active materials used commonly are LiCoO2 with a layered structure and Li(Ni, Mn, Co, Al)O2 in which Co in LiCoO2 is partially replaced with Ni, Mn, and Al.43 LiCoO2, Li(Ni, Mn, Co, Al)O2, and LiMn2O444 are 4 V class electrode active materials, and LiFePO4 is 3.5 V class.45 Research and development of 5 V class high-potential materials and sulfur-based positive electrode active materials, which are 2 V class materials but enable a significantly large capacity, are being actively conducted. Graphite is the most commonly used negative electrode active material. Other materials that can be used as negative electrode active materials include hard carbon, spinel Li4Ti5O12,46 silicon, and tin, which form alloys with Li.
Solid electrolytes are key materials for all-solid-state batteries. The requirements that should be satisfied as solid electrolytes for this purpose are listed below.
Ideally, solid electrolytes that satisfy all the above requirements should be developed. In contrast, unlike liquid systems, different solid electrolytes can be used in the positive and negative electrode composite layers and in the solid electrolyte separator layer of solid systems. Typical sulfide electrolytes include Li10GeP2S12 (LGPS),47 argyrodite Li6PS5Cl,48 and Li3PS4-based glass and glass ceramics,49–51 whereas typical oxide electrolytes include Li0.34La0.51TiO2.94,6 Li1.3Al0.3Ti1.7(PO4)3,52 and Li7La3Zr2O12 (LLZ).8 Li3.3PO3.8N0.22,53 Li3BO3-based glass, and glass ceramics54,55 are also used for thin-film or small batteries. Halide electrolytes and complex hydrides have also attracted attention.56–58 Sulfide, halide, and complex hydride electrolytes are generally characterized by high conductivity and excellent formability. Oxide-based electrolytes, on the other hand, basically have high chemical stability.
Most inorganic solid electrolytes show high Li transference numbers σLi/σtotal close to 1. In solution systems, multiple ions, such as cations and anions, exist in the solvent, but solid electrolytes are often single-ion conductors in which only the carrier ions move. The electrochemical behavior can be interpreted in a simple manner because the carrier ion concentration changes very little in the electrolyte.
3.2 Electrode potential and reference electrode in all-solid-state batteriesIt is necessary to fabricate a three-electrode electrochemical cell to accurately measure the electrode potentials. Detailed analysis requires investigation of the potential of each positive and negative electrode. However, the fabrication of a three-electrode cell in a solid-state system is more difficult than that in a liquid system. Thus, most electrochemical tests of all-solid-state batteries are performed in two-electrode configurations. As described above (Section 2.3), the electrode potential of solid solution reagion gradually shifts with a change in the composition. On the other hand, that of two-phase coexistence region shows constant voltage. Therefore, electrode active materials with two-phase coexistence reactions are useful as reference electrodes because they exhibit a constant voltage.
Li metal is typically used as the reference electrode for lithium-ion batteries. Although the Li metal electrode is the simplest counter/reference electrode in all-solid-state batteries, short-circuiting and electrolyte reductive decomposition can occur; thus, a more stable counter/reference electrode for the two-electrode configuration is needed.
Li–In alloys are mainly used as counter/reference electrodes in all-solid-state lithium batteries using sulfide-based solid electrolytes,12,59 because they have a high Li diffusion coefficient and a constant potential (0.62 V vs. Li+/Li) over a wide composition range owing to the two-phase coexistence region between In and LiIn. Figure 6 shows the initial discharge curve of the all-solid-state Li/In cell using a sulfide solid electrolyte. A constant voltage plateau is observed at 0.62 V vs. Li electrode for the wide composition range of x = 0–1 in LixIn.

Discharge curve of the all-solid-state Li/In cell, exhibiting the two-phase coexistence region.
Furthermore, the resistance of the interface between the electrode and electrolyte for the Li–In electrode is relatively small in most cases. The relatively high electrode potential effectively suppresses the short-circuiting caused by Li dendrites, and thus acts as a stable counter electrode. More practically, it is recommended that LixIn be used within the composition range of x = 0.2 and 0.8. Li–In alloy foils can be obtained simply by overlapping and pressing a Li metal foil and an In metal foil because Li–In alloys have high Li diffusion coefficients. The key to successful alloying is the removal of impurities from the Li and In metal surfaces and holding them under pressure for a period of time. Similarly, Li4Ti5O12 in the two-phase range is also useful as a counter/reference electrode in the same way.
In a three-electrode system for a lithium-ion battery using a liquid electrolyte, a Li wire or ribbon electrode located between the working and counter electrodes is typically used as the third electrode, as shown in Fig. 2. Although it is not easy to construct three-electrode cells similar to liquid systems for all-solid-state batteries, researchers have reported three-electrode cells for the characterization of all-solid-state batteries.36,60,61
In lithium-ion battery research, there is a great need to measure the potential and polarization (resistance) of the positive and negative electrodes using three-electrode cells to identify the dominant factors limiting the battery performance and the degradation sites. For example, Ikezawa et al. prepared an all-solid-state three-electrode cell using a chemically reduced Li4Ti5O12 reference electrode for impedance analysis. For the shapes and locations of the reference electrodes, they chose mesh-type reference electrodes and positioned them between the positive and negative electrodes to reduce artifacts in the electrochemical impedance spectroscopy data.36
3.3 Electrochemical window of solid electrolytesDFT calculations have clearly shown that most solid electrolytes are not thermodynamically stable against Li metal and are reduced at low potentials.62 For example, the thermodynamically stable potential range for many sulfide-based solid electrolytes is very narrow, lying between 1.5–2.5 V vs. Li+/Li. However, many of them can be used with both 4 V class positive electrode materials and 0–1 V class negative electrode materials. Figure 7 shows the cyclic voltammogram of an all-solid-state Li/Li7P3S10O/stainless steel (SS) cell. The cell shows a relatively large current of Li plating/stripping at 0 V vs. Li+/Li, and there is no large current even at 10 V vs. Li+/Li. Although this result does not strictly mean that the solid electrolyte has an electrochemical window over a wide range of 0–10 V, it clearly indicates that this solid electrolyte could be used over a wide range of potentials. In contrast to liquid electrolytes, large oxidation and reduction currents often do not flow in the solid electrolytes. This is because oxidative decomposition does not occur continuously when the products of oxidative or reductive decomposition have low electronic conductivities. Non-flowability also contributes to the suppression of continuous oxidation and reduction currents due to side reactions. As shown in Fig. 7b, the anodic and cathodic currents due to the oxidation and reduction of the solid electrolyte observed at approximately 2 V vs. Li+/Li are small, and the value is approximately 1/100 of the current density observed in Li plating/stripping. When the oxidation or reduction reaction products have very low electronic conductivities, the solid electrolyte does not undergo further decomposition. In addition, when the oxidation or reduction reaction products also have sufficient ionic conductivity, the all-solid-state battery works well, even though the solid electrolyte itself is thermodynamically unstable under high or low voltages. Instead of the self-forming oxidative and reductive decomposition layers, it is effective to introduce a buffer layer with very low electronic conductivity and high ionic conductivity in advance. It has been found that such a coating on the surface of the electrode active material particles can significantly reduce the resistance of the interface between the electrode and electrolyte.12,63 The difference of chemical potential of Li between electrodes and solid electrolytes should be an important factor to understanding the side reaction and interfacial resistance.

Cyclic voltammogram of the all-solid-state Li/Li7P3S10O/SS cell. (b) shows magnified figure of (a).
Electrochemistry has progressed tremendously over the past several decades. The evaluation of the electrode potential is, in essence, the foundation of electrochemistry, as well as the key to developing electrochemical applications. In this Part 2, a summary of several reference electrodes and the methodologies and strategies for measuring the electrode potential in nonaqueous and solid electrolytes, thereby providing further insight into the nature of the electrode potential. The electrode potential is not only required for the development of new materials for electrochemical applications at the lab scale but also must be considered in the production stage. Thus, we believe that this literature will serve as a useful guideline and strategy that can be directly utilized to develop devices.
The data that support the findings of this study are openly available under the terms of the designated Creative Commons License in J-STAGE Data at https://doi.org/10.50892/data.electrochemistry.21357846. The authors' profiles of this paper can be found on the preface.64
Jinkwang Hwang: Writing – original draft (Lead)
Takayuki Yamamoto: Writing – original draft (Lead)
Atsushi Sakuda: Writing – original draft (Lead)
Kazuhiko Matsumoto: Writing – review & editing (Supporting)
Kohei Miyazaki: Writing – review & editing (Supporting)
The authors declare no conflict of interest in the manuscript.
This paper constitutes a collection of papers edited as the proceedings of the 51st Electrochemistry Workshop organized by the Kansai Branch of the Electrochemical Society of Japan.
J. Hwang, T. Yamamoto, and A. Sakuda: These authors contributed equally to this work.
J. Hwang, T. Yamamoto, A. Sakuda, K. Matsumoto, and K. Miyazaki: ECSJ Active Members