Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 25 July 2022
Sec. Electrochemistry
This article is part of the Research Topic Rising stars in Electrochemistry 2021 View all 5 articles

Essays on Conceptual Electrochemistry: I. Bridging Open-Circuit Voltage of Electrochemical Cells and Charge Distribution at Electrode–Electrolyte Interfaces

Updated
  • Institute of Energy and Climate Research, IEK-13, Theory and Computation of Energy Materials, Jülich, Germany

We ponder over how an electrochemical cell conforms itself to the open-circuit voltage (OCV) given by the Nernst equation, where properties of the electrodes play no role. We first show, via a pedagogical derivation of the Nernst equation, how electrode properties are canceled and then take a closer look into the electrode–electrolyte interface at one electrode by linking charge and potential distributions. We obtain an equilibrium Poisson–Nernst equation that shows how the charge distribution across an electrode–electrolyte interface can be dictated by the chemical potentials of redox species. Taking a H2/O2 fuel cell as an example, we demystify the formal analysis by showing how the two electrodes delicately regulate their “electron tails” to abide by the Nernst equation. In this example, we run into a seemingly counterintuitive phenomenon that two electrodes made of the same transition metal display two distinct potentials of zero charge. This example indicates that the double layer at transition metals with chemisorption can display distinct behaviors compared to ideally polarizable double layers at sp metals.

Many concepts in electrochemistry appear simple, but they are more complicated than they seem. One example is the electrode potential, which is among the first concepts we will encounter when opening any introductory electrochemistry textbook. However, the definition of absolute electrode potential had been extensively discussed in the 1970s–1980s (Trasatti, 1980; 1982; Trasatti, 1986a;b; Trasatti, 1986c; 1990). To illustrate, Trasatti (1990) stated that “the present author has considerably contributed to the discussion with several papers during the last 15 years in an attempt to bring the various views back to a unifying approach.” Trasatti’s formal theory states that “the absolute electrode potential is the difference in electronic energy between a point inside the metal (Fermi level) and a point outside the solution.” This definition has been adopted in the recommendation of The International Union of Pure and Applied Chemistry (IUPAC) (Trasatti, 1986a).

There is another “familiar stranger,” namely, the open-circuit voltage (OCV) of an electrochemical cell, which is simply the potential difference between two terminals of the cell under the open-circuit condition. Though the definition is unequivocal and the measurement easily accessible, the obtained value can be elusive and its interpretation disputed. A famous example is the OCV of a hydrogen/oxygen fuel cell at rest, also called the rest potential: why is the measured value significantly lower, by several hundreds of millivolts, than the thermodynamic value of 1.23 V versus the standard hydrogen electrode (SHE)?

James Hoare systematically investigated this problem in the 1960s–1970s (Hoare, 1962; Hoare, 1963; Hoare, 1964a; Hoare, 1964b; Hoare, 1964c; Thacker and Hoare, 1971; Hoare, 1974; Hoare, 1978). He concluded that the measured rest potential is a mixed potential of the oxygen reduction reaction and another parasitic reaction, namely, water dissociation reaction forming adsorbed hydroxyl. There are continued interests in understanding the rest potential of practical fuel cells beyond the glass cells used in Hoare’s experiments (Zhang et al., 2006; Vilekar and Datta, 2010; Reimer et al., 2019). In these practical situations, other factors such as hydrogen crossover come into play.

As an established fact, the OCV of the full cell calculated by the Nernst equation is independent of electrode properties, such as the chemical potential of electrons, the work function, and the valency electron density. For example, the OCV of a hydrogen/oxygen fuel cell under standard conditions is 1.23 V, regardless of the electrode materials. A seemingly naïve question might be asked:

How does the OCV of an electrochemical cell abide by the Nernst equation by canceling off all electrode-specific properties in an amazingly precise manner?

Though the answer to this question is self-evident, as thermodynamic quantities should be path-independent, a revisit of it from an alternative view is not meaningless. It may reveal something interesting. This question is discussed herein by first presenting a pedagogical derivation of the OCV of a general electrochemical cell. By decomposing the OCV into several parts, we show explicitly how electrode-specific properties cancel each other. Further delving into the electrode–electrolyte interface at one electrode, we link the potential difference between electrode and electrolyte phases with the net charge distribution at the electrode–electrolyte interface, which is then correlated with the chemical potentials of redox species involved in the reaction occurring on this electrode. The formal analysis is then demystified by taking an H2/O2 fuel cell as an example. In this example, we run into a weird point that two identical platinum electrodes constituting a hydrogen-oxygen fuel cell have different potentials of zero charge. We close this essay by commenting on why Trasatti’s relationship between the potential of the zero charge of an electrochemical interface and the work function of metals does not apply to the present case.

For an overall electrochemical reaction aA+bB=cC+dD, the Nernst equation for the cell relates the OCV under any condition (E) to that under standard conditions (E0) and the species activities (ai, i=A, B, C, D), written as

E=E0RTnFlnaCcaDdaAaaBb(1)

with n being the number of electrons transferred, and E0 is calculated by

E0=ΔGr0nF(2)

ΔGr0 is the Gibbs energy change under standard conditions, calculated by

ΔGr0=cμC0+dμD0aμA0bμB0(3)

where μio (i=A, B, C, D) is the chemical potential of species i under standard conditions. Indeed, there is no electrode-specific property in the above equations for OCV.

Now, we will formulate the OCV in another way and decouple it into a serial connection of several potential differences. OCV is measured by connecting the cathode and anode with a voltmeter, as shown in Figure 1. The voltmeter has extremely high resistance, so a circuit connected with it allows for negligible current and thereby can be taken as an open circuit. The wires that connect the voltmeter with the two electrodes are of the same metal. The value on the voltmeter, herein the OCV, is expressed by the difference in the Fermi levels of these two wires:

E=ϵF,wcϵF,wae (4)

with ϵF,wi=μe,wieϕwi and ϕwi (i = a, c) denote the respective inner potential of the metal wires at the cathode side and anode side. As, for the same material, μwc=μwa, we arrive at the OCV:

E=ϕwcϕwa(5)

FIGURE 1
www.frontiersin.org

FIGURE 1. Schematic of an electrochemical cell. A voltmeter connects the cathode and anode to measure the open-circuit potential. The two wires of the voltmeter are made of the same metal.

The difference in the potential in two metal wires can be decoupled into a serial connection of four potential differences:

ϕwcϕwa=Δwcϕc+Δcϕs+Δsϕa+Δaϕwa(6)

with Δxϕy=ϕxϕy. These four Δxϕy in Eq. 6 correspond to the four circles in Figure 1. The first and the last terms are the metal contact potential difference, whereas the other two are the potential difference at the metal–solution interface. We will express them one by one.

The metal contact potential difference can be formulated with the condition that the electrons in connected metals are in electrochemical equilibrium (or Fermi levels align):

μe,wcFϕwc=μe,cFϕc(7)
μe,waFϕwa=μe,aFϕa(8)

Combining Eqs 7, 8, the two chemical potentials in metal wires are canceled:

Δwcϕc+Δaϕwa=μe,aFμe,cF(9)

We update Eq. 5 with the help of Eqs 6, 9:

E=(Δcϕsμe,cF)(Δaϕsμe,aF)(10)

where (Δiϕsμe,iF) (i = a, c) is one definition of absolute electrode potential (Trasatti, 1990). We are left with two potential differences at the metal–solution interface: Δcϕs and Δaϕsa.

Considering a redox couple in equilibrium on the cathode side, O+ne=R. The charges bore in O and R are not explicitly expressed, but their sum must cancel the n electrons. Therefore, equilibrium gives

μO,c+nFϕs+nμe,cnFϕc=μR,c(11)

Some rearrangements give

Δcϕs=μO,cμR,cnF+μe,cF(12)

In the same vein, Δaϕs reads

Δaϕs=μO,aμR,anF+μe,aF(13)

Inserting Eqs 12, 13 back to Eq. 10, μe,c and μe,a are canceled:

E=μO,cμR,cnFμO,aμR,anF(14)

Shown above is how all electrode-specific properties, namely, μe,mc and μe,ma, are canceled in the expression of OCV. If the chemical potential is expressed by the standard chemical potential and correction for the species activity is applied, Eq. 14 reduces back to Eq. 1.

Of note, a usual misunderstanding of the OCV reads OCV=ΔcϕsΔaϕs. Eq. 6 reminds us that the two contact potentials should never be forgotten. The correct equation for the OCV is Eq. 10.

Next, we look further into Δiϕs (i=a,c). The distribution of electric potential from the electrode phase to the solution phase is governed by the Poisson equation:

.(ϵ(r)ϕ(r))=ρ(r)(15)

where ϵ(r) is the dielectric permittivity, which is spatially inhomogeneous in the EDL and ρ(r) the net charge distribution, which can be decomposed into an electrode part ρM(r) and a solution part ρS(r):

ρ(r)=ρM(r)+ρS(r)(16)

We take some more lines to explain Eq. 16. This electrode–solution dichotomy of the net charge distribution is proper for the case without ion-specific adsorption. ρM(r) is the sum of the negative charge carried by valence electrons of the electrode material and the positive charge of cationic cores of the electrode material. A small portion of valence electrons of the electrode will enter into the solution counterpart, termed the electron spillover phenomenon and called vividly “electron tail.” It can be modeled by the Thomas–Fermi theories (Badiali, 1987; Kornyshev, 1989; Schmickler, 1996; Huang, 2021; Huang et al., 2021). ρS(r) is the sum of negative charge of anions and positive charge of cations in solution.

For the cases with ion-specific adsorption, chemical bonds are formed between the electrode and specifically adsorbed ions. In general, the specifically adsorbed ions are not electroneutral but still possess a fraction of charge (Schmickler and Guidelli, 2014). It then becomes principally difficult to separate ρM(r) and ρs(r) because ions, originally belonging to ρs (r), are now adsorbed onto the electrode with residual charge. A possible scheme is to treat the electrode together with the specifically adsorbed ions as a whole, whose net charge is denoted by ρM(r). This way, ρs(r) is limited to the net charge of the nonspecifically absorbed ions in solution; see a recent Minireview on the surface charging behavior of EDLs with chemisorption (Huang, 2022).

For a one-dimensional case without ion-specific adsorption, where r is replaced with x, Eq. 15 reads

d(ϵ(x)dϕ(x)dx)dx=ρ(x)(17)

In bulk solution, ρ(xs)=0 and thus dϕ(xs)dxs=0. Integrating from xs to x gives

ϵ(x)dϕ(x)dx=xsxρ(x)dx(18)

Integrating again from xs to xi gives the potential difference across each electrode–electrolyte interface (i=a,c):

Δiϕs=xsxidx1ϵ(x)xsxdxρi(x)(19)

with xs being the location of solution bulk and xi the location of the electrode bulk. Provided that dielectric properties of the electrode–solution interface are known a priori, Δiϕs is exclusively determined by the net charge distribution ρi(x).

As no assumption on the condition of the interfacial reaction is made in the derivation, Eq. 19 is valid under both equilibrium and nonequilibrium conditions.

It has been established that the potential difference across the electrode–solution interface Δiϕs is related to the chemical potentials of the redox species in solution via Eqs 12, 13. Now, with the help of Eq. 19, we arrive at an equilibrium Poisson–Nernst formula that bridges the net charge distribution at the electrode–electrolyte interface ρi(x) with the chemical potentials:

xsxidx1ϵ(x)xsxdxρi(x)=μO,iμR,inF+μe,iF(20)

Eq. 20 implies that ρi(x) can be tuned by the chemical potentials of redox species and also the electrons. Eq. 20 is valid under the equilibrium condition only due to the use of Eqs 12, 13.

Based on Eq. 20, two corollaries can be made: 1) for two electrodes of the same material where different redox species are in equilibrium, the net charge distribution will be different and 2) for two electrodes of different materials, even if the redox species—with whom they have established equilibrium—are the same, the net charge distribution will be different.

In order to understand the first corollary from a microscopic perspective, it is instructive to consider an H2/O2 fuel cell, schematically shown in Figure 2. Both electrodes made of platinum are immersed in an acidic solution with pH = 0. The cell is kept under the open-circuit condition.

FIGURE 2
www.frontiersin.org

FIGURE 2. (A) Layout of an H2/O2 fuel cell under the standard condition (1 atm O2, 1 atm H2, 1M H+, and 25 °C). The solid and dashed gray lines denote, respectively, the potential distribution before and after H2 and O2 are injected into the cell. PEM is short for the proton exchange membrane. The schematically shown E=ΔcϕsΔaϕs is only valid for the case of two electrodes made of identical material and thus μe,c=μe,a. (B) and (C) show the pullout and pushback effects by oxygen and hydrogen on the metal electron density distribution, respectively. The dashed and solid red lines denote the electron density distribution before and after the injection of H2 and O2.

At the outset, no H2/O2 gas has been injected into the cell and dissolved in the solution. Therefore, the anode and cathode sides are totally symmetric in terms of electrode material and electrolyte component. Therefore, regardless of any reaction that may occur at both interfaces, we can safely state that Δcϕs=Δaϕs and the cell voltage is exactly zero, as schematically shown in Figure 2A.

Meanwhile, dashed lines in Figures 2B,C show schematically the electron density distribution at both electrodes for this symmetric case. The detailed charge distributions can be calculated using first-principles simulations based on the density-functional theory (DFT) (Le et al., 2017; Sakong and Groß, 2018, 2020; Le et al., 2021).

From now on, let oxygen and hydrogen gas come in the cathode and anode compartments while the external electric circuit is still kept open. The redox couple of O2, H+, and H2O occupies the cathode side, whereas that of H+ and H2 occupies the anode side. When the equilibrium for each redox couple is established respectively at the cathode and the anode, Δcϕs and Δaϕs are determined by Eqs 11, 12. Under standard conditions, the Nernst equation, Eq. 1, gives E=1.23 V. Any parasitic reaction that results in a mixed potential is not considered here.

Using Eq. 9 and recalling the two electrodes as both made of Pt, that is, μe,c=μe,a, we have

ΔcϕsΔaϕs=1.23 V(21)

The differences between Δcϕs and Δaϕs must be ascribed to the difference in the net charge distribution ρ(x), according to Eq. 19, which, in turn, is tuned by the chemical potentials of redox species, according to Eq. 20. As the electrochemical cell is kept under the ideal open-circuit condition, electrons cannot be exchanged between the two electrodes. On the cathode side, after oxygen gas is introduced, oxygen reduction reaction, O2 + 4H+ + 4e → 2H2O, must occur until it equilibrates with the reverse reaction. In other words, some electrons have been consumed in the cathode platinum to establish the equilibrium. Since the circuit is open and there is no way to compensate the electron consumption, the cathode platinum must be positively charged. Following the same line of reasoning, the anode platinum must be negatively charged due to excess electrons generated from hydrogen oxidation, H2 → 2H+ + 2e. However, the amount of excess on both electrodes is unknown without a model for the electrochemical double layers. Moreover, the presence of H2 and O2 also contribute to the change of net charge distribution, as shown in the dashed and solid curves in Figures 2B,C.

After introducing hydrogen and oxygen into the anode and cathode, hydrogen repels metal electrons back into the metal skeleton, while oxygen pulls more metal electrons out of the metal skeleton. The pullout effect by oxygen increases the surface dipole moment by increasing the distance between positively charged ionic cores and negatively charged “electron tails.” The pushback effect by the hydrogen, on the contrary, does the opposite. The uplift of Δcϕs is a combined effect of positively charged electrode and the raised dipole moment due to the pullout effect. In the same logic, the suppression of Δaϕs is a combined effect of negatively charged electrode and reduced dipole moment due to the pushback effect. The ultimate thrust for adjusting the free charge distribution is the thermodynamic requirement expressed in Eq. 20.

Theory and experiment revealed that the pushback effect of adsorbed hydrogen decreases the work function of Pt(111) (Li et al., 2021), whereas the pullout effect of adsorbed oxygen increases it (Malek and Eikerling, 2018). As regards the EDL at Pt(111) contacted with an acidic aqueous solution, a mean-field model has shown that chemisorption of partially charged hydroxyl and oxygen contributes an additional surface dipole moment, leading to a second pzc and an overall nonmonotonic surface charging behavior (Huang et al., 2016; Huang et al., 2018), which are confirmed in atomistic simulations (Fernandez-Alvarez and Eikerling, 2019; Tesch et al., 2021; Braunwarth et al., 2022).

Let us relook at this problem from the perspective of pzc. The pzc is defined as the potential at which no net charge is accumulated on the electrode. The anode herein is actually the standard hydrogen electrode (SHE). It has a potential of 0 V with reference to the SHE because it takes itself as the reference. Since the anode is negatively charged, it has a pzc higher than 0 V. The cathode has a potential of 1.23 V with reference to the SHE. Since it is positively charged, it has a pzc lower than 1.23 V. Due to the pullout and pushback effects, the two pzc are arguably not equal to each other.

Ostensibly, it is strange that the two electrodes, made of the same metal platinum, could have so different pzcs. The point is that the pzc is not a property of the electrode itself but a property of the electrode–electrolyte interface; the latter could be significantly changed by the electrolyte composition. That has been demonstrated in several experimental works (Smalley, 2017; Shatla et al., 2021) and discussed in a recent modeling work (Huang et al., 2020). As an immediate implication, it is ambiguous to say the pzc of an electrode material without specifying the adjacent electrolyte solution.

As a final remark, the present case where platinum can have two different pzcs contradicts Trasatti’s relationship between the pzc of an EDL and the work function of the metal constituting the EDL (Trasatti, 1971). It should be noted that Trasatti’s relationship was established for “clean” metal surfaces without adsorption or chemisorption. However, chemisorption occurs on the surfaces of two Pt electrodes for the present case where the surface structure of the metal changes from its original state. Therefore, our analysis indicates that Trasatti’s relationship does not apply to electrocatalytic EDLs, which display distinct behaviors compared to ideally polarizable EDLs at “clean” metal surfaces. For instance, no Gouy–Chapman minimum was observed in the differential double-layer capacitance curves of Pt(111), even in the so-called pure double-layer region (Pajkossy and Kolb, 2007; Ojha et al., 2020; Ojha et al., 2022). It is also important to notice that the usual pzc is measured under a closed-circuit condition, whereas the two platinum electrodes in our case are under open-circuit condition. Therefore, the pzc under the open-circuit condition could be different from that under the closed-circuit condition because the surface state of the electrodes changes.

In conclusion, we have touched upon the seemingly trivial question that how the electrochemical cell conforms itself to the OCV stipulated by the Nernst equation. We have obtained an equilibrium Poisson–Nernst equation in Eq. 20 relating the net charge distribution across the electrode–electrolyte interface to chemical potentials of redox species. Furthermore, an H2/O2 fuel cell has been used to illustrate how the OCV is generated microscopically via tuning the “electron tail,” namely, the spillover electron, in the EDL. It is important to note that the analysis has been limited to the equilibrium state under the open-circuit condition. The closed-circuit condition with charge transfer reactions occurring at both electrodes will be addressed in a separate essay.

Data Availability Statement

The original contributions presented in the study are included in the article/supplementary material. Further inquiries can be directed to the corresponding author.

Author Contributions

JH initiated this work, and YZ joined later and contributed substantially. JH wrote the first draft, which was revised by YZ. We confirmed the submission.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

The handling editor VB declared a shared affiliation with the authors at the time of review.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

Dedicated to Professor Juan Feliu on the occasion of his official retirement from Alicante University who continues to inspire our understanding of electrochemical interfaces and reactions.

References

Badiali, J. P. (1987). The Jellium Model in Electrochemistry. Berichte Bunsenges. für Phys. Chem. 91, 270–276. doi:10.1002/bbpc.19870910406

CrossRef Full Text | Google Scholar

Braunwarth, L., Jung, C., and Jacob, T. (2022). Potential-dependent Pt (111)—water Interface: Tackling the Challenge of a Consistent Treatment of Electrochemical Interfaces. doi:10.26434/chemrxiv-2022-xndvx

CrossRef Full Text | Google Scholar

Fernandez-Alvarez, V. M., and Eikerling, M. H. (2019). Interface Properties of the Partially Oxidized Pt(111) Surface Using Hybrid DFT-Solvation Models. ACS Appl. Mat. Interfaces 11, 43774–43780. doi:10.1021/acsami.9b16326

PubMed Abstract | CrossRef Full Text | Google Scholar

Hoare, J. P. (1963). A Study of the Rest Potentials in the Gold-Oxygen-Acid System. J. Electrochem. Soc. 110, 245. doi:10.1149/1.2425724

CrossRef Full Text | Google Scholar

Hoare, J. P. (1964b). On the Mixed Potentials Observed in the Iridium-Oxygen-Acid System. J. Electrochem. Soc. 111, 988. doi:10.1149/1.2426305

CrossRef Full Text | Google Scholar

Hoare, J. P. (1978). On the Normal Oxygen Potential at a Platinum-Oxygen Alloy Diaphragm Electrode. J. Electrochem. Soc. 125, 1768–1771. doi:10.1149/1.2131291

CrossRef Full Text | Google Scholar

Hoare, J. P. (1962). Rest Potentials in the Platinum-Oxygen-Acid System. J. Electrochem. Soc. 109, 858. doi:10.1149/1.2425569

CrossRef Full Text | Google Scholar

Hoare, J. P. (1964c). Some Double-Layer Capacity Measurements on Platinum Electrodes. Nature 204, 71–73. doi:10.1038/204071b0

PubMed Abstract | CrossRef Full Text | Google Scholar

Hoare, J. P. (1964a). The Effect of Metal Dissolution on the Rest Potential in the Palladium-Oxygen-Acid System. J. Electrochem. Soc. 111, 610. doi:10.1149/1.2426193

CrossRef Full Text | Google Scholar

Hoare, J. P. (1974). The Effect of Oxygen Dissolved in Pt on the Potential of a Pt∕O2 Electrode at Rest. J. Electrochem. Soc. 121, 872. doi:10.1149/1.2401940

CrossRef Full Text | Google Scholar

Huang, J., Chen, S., and Eikerling, M. (2021). Grand-Canonical Model of Electrochemical Double Layers from a Hybrid Density–Potential Functional. J. Chem. Theory Comput.

Google Scholar

Huang, J. (2021). Hybrid Density-Potential Functional Theory of Electric Double Layers. Electrochimica Acta 389, 138720. doi:10.1016/j.electacta.2021.138720

CrossRef Full Text | Google Scholar

Huang, J., Li, P., and Chen, S. (2020). Potential of Zero Charge and Surface Charging Relation of Metal-Solution Interphases from a Constant-Potential Jellium-Poisson-Boltzmann Model. Phys. Rev. B 101, 125422. doi:10.1103/physrevb.101.125422

CrossRef Full Text | Google Scholar

Huang, J., Malek, A., Zhang, J., and Eikerling, M. H. (2016). Non-monotonic Surface Charging Behavior of Platinum: A Paradigm Change. J. Phys. Chem. C 120, 13587–13595. doi:10.1021/acs.jpcc.6b03930

CrossRef Full Text | Google Scholar

Huang, J. (2022). Surface Charging Behaviors of Electrocatalytic Interfaces with Partially Charged Chemisorbates. Curr. Opin. Electrochem. 33, 100938. doi:10.1016/j.coelec.2022.100938

CrossRef Full Text | Google Scholar

Huang, J., Zhou, T., Zhang, J., and Eikerling, M. (2018). Double Layer of Platinum Electrodes: Non-monotonic Surface Charging Phenomena and Negative Double Layer Capacitance. J. Chem. Phys. 148, 044704. doi:10.1063/1.5010999

PubMed Abstract | CrossRef Full Text | Google Scholar

Kornyshev, A. A. (1989). Metal Electrons in the Double Layer Theory. Electrochimica Acta 34, 1829–1847. doi:10.1016/0013-4686(89)85070-4

CrossRef Full Text | Google Scholar

Le, J., Iannuzzi, M., Cuesta, A., and Cheng, J. (2017). Determining Potentials of Zero Charge of Metal Electrodes versus the Standard Hydrogen Electrode from Density-Functional-Theory-Based Molecular Dynamics. Phys. Rev. Lett. 119, 016801. doi:10.1103/PhysRevLett.119.016801

PubMed Abstract | CrossRef Full Text | Google Scholar

Le, J. B., Fan, Q. Y., Li, J. Q., and Cheng, J. (2020). Molecular Origin of Negative Component of Helmholtz Capacitance at Electrified Pt (111)/Water Interface. Science Advances 6 (41), eabb1219.

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, P., Huang, J., Hu, Y., and Chen, S. (2021). Establishment of the Potential of Zero Charge of Metals in Aqueous Solutions: Different Faces of Water Revealed by Ab Initio Molecular Dynamics Simulations. J. Phys. Chem. C 125, 3972–3979. doi:10.1021/acs.jpcc.0c11089

CrossRef Full Text | Google Scholar

Malek, A., and Eikerling, M. H. (2018). Chemisorbed Oxygen at Pt(111): a DFT Study of Structural and Electronic Surface Properties. Electrocatalysis 9, 370–379. doi:10.1007/s12678-017-0436-0

CrossRef Full Text | Google Scholar

Ojha, K., Arulmozhi, N., Aranzales, D., and Koper, M. T. M. (2020). Double Layer at the Pt(111)-Aqueous Electrolyte Interface: Potential of Zero Charge and Anomalous Gouy-Chapman Screening. Angew. Chem. Int. Ed. 59, 711–715. doi:10.1002/anie.201911929

PubMed Abstract | CrossRef Full Text | Google Scholar

Ojha, K., Doblhoff-Dier, K., and Koper Marc, T. M. (2022). Double-layer Structure of the Pt(111)–Aqueous Electrolyte Interface. Proc. Natl. Acad. Sci. 119, e2116016119. doi:10.1073/pnas.2116016119

PubMed Abstract | CrossRef Full Text | Google Scholar

Pajkossy, T., and Kolb, D. M. (2007). Double Layer Capacitance of the Platinum Group Metals in the Double Layer Region. Electrochem. Commun. 9, 1171–1174. doi:10.1016/j.elecom.2007.01.002

CrossRef Full Text | Google Scholar

Reimer, U., Cai, Y., Li, R., Froning, D., and Lehnert, W. (2019). Time Dependence of the Open Circuit Potential of Platinum Disk Electrodes in Half Cell Experiments. J. Electrochem. Soc. 166, F3098–F3104. doi:10.1149/2.0121907jes

CrossRef Full Text | Google Scholar

Sakong, S., and Groß, A. (2018). The Electric Double Layer at Metal-Water Interfaces Revisited Based on a Charge Polarization Scheme. J. Chem. Phys. 149, 084705. doi:10.1063/1.5040056

PubMed Abstract | CrossRef Full Text | Google Scholar

Sakong, S., and Groß, A. (2020). Water Structures on a Pt(111) Electrode from Ab Initio Molecular Dynamic Simulations for a Variety of Electrochemical Conditions. Phys. Chem. Chem. Phys. 22, 10431–10437. doi:10.1039/c9cp06584a

PubMed Abstract | CrossRef Full Text | Google Scholar

Schmickler, W. (1996). Electronic Effects in the Electric Double Layer. Chem. Rev. 96, 3177–3200. doi:10.1021/cr940408c

PubMed Abstract | CrossRef Full Text | Google Scholar

Schmickler, W., and Guidelli, R. (2014). The Partial Charge Transfer. Electrochimica Acta 127, 489–505. doi:10.1016/j.electacta.2014.02.057

CrossRef Full Text | Google Scholar

Shatla, A. S., Landstorfer, M., and Baltruschat, H. (2021). On the Differential Capacitance and Potential of Zero Charge of Au(111) in Some Aprotic Solvents. ChemElectroChem 8, 1817–1835. doi:10.1002/celc.202100316

CrossRef Full Text | Google Scholar

Smalley, J. F. (2017). Potential of Zero Charge and its Temperature Derivative for Au(111) Electrode|Alkanethiol SAM|1.0 M Aqueous Electrolyte Solution Interfaces: Impact of Electrolyte Solution Ionic Strength and its Effect on the Structure of the Modified Electrode|Electrolyte Solution Interface. J. Phys. Chem. C 121, 9260–9272. doi:10.1021/acs.jpcc.6b10954

CrossRef Full Text | Google Scholar

Tesch, R., Kowalski, P. M., and Eikerling, M. H. (2021). Properties of the Pt(111)/electrolyte Electrochemical Interface Studied with a Hybrid DFT-Solvation Approach. J. Phys. Condens. Matter 33, 444004. doi:10.1088/1361-648x/ac1aa2

PubMed Abstract | CrossRef Full Text | Google Scholar

Thacker, R., and Hoare, J. P. (1971). Sorption of Oxygen from Solution by Noble Metals. J. Electroanal. Chem. Interfacial Electrochem. 30, 1–14. doi:10.1016/0368-1874(71)85027-x

CrossRef Full Text | Google Scholar

Trasatti, S. (1986b). Components of the Absolute Electrode Potential. Conceptions and Misinterpretations. Mater. Chem. Phys. 15, 427–438. doi:10.1016/0254-0584(86)90026-x

CrossRef Full Text | Google Scholar

Trasatti, S. (1986c). “Components of the Electrode Potential. Concepts and Problems,” in Trends in Interfacial Electrochemistry. Editor A. F. Silva (Dordrecht: Springer Netherlands), 1–24. doi:10.1007/978-94-009-4694-1_1

CrossRef Full Text | Google Scholar

Trasatti, S. (1990). The "absolute" Electrode Potential-The End of the Story. Electrochimica Acta 35, 269–271. doi:10.1016/0013-4686(90)85069-y

CrossRef Full Text | Google Scholar

Trasatti, S. (1986a). The Absolute Electrode Potential: an Explanatory Note (Recommendations 1986). Pure Appl. Chem. 58, 955–966. doi:10.1351/pac198658070955

CrossRef Full Text | Google Scholar

Trasatti, S. (1982). The Concept and Physical Meaning of Absolute Electrode Potential. J. Electroanal. Chem. Interfacial Electrochem. 139, 1–13. doi:10.1016/0022-0728(82)85100-0

CrossRef Full Text | Google Scholar

Trasatti, S. (1980). “The Electrode Potential,” in Comprehensive Treatise of Electrochemistry: The Double Layer. Editors J. O. M. Bockris, B. E. Conway, and E. Yeager (Boston, MA: Springer US), 45–81. doi:10.1007/978-1-4615-6684-7_2

CrossRef Full Text | Google Scholar

Trasatti, S. (1971). Work Function, Electronegativity, and Electrochemical Behaviour of Metals. J. Electroanal. Chem. Interfacial Electrochem. 33, 351–378. doi:10.1016/s0022-0728(71)80123-7

CrossRef Full Text | Google Scholar

Vilekar, S. A., and Datta, R. (2010). The Effect of Hydrogen Crossover on Open-Circuit Voltage in Polymer Electrolyte Membrane Fuel Cells. J. Power Sources 195, 2241–2247. doi:10.1016/j.jpowsour.2009.10.023

CrossRef Full Text | Google Scholar

Zhang, J., Tang, Y., Song, C., Zhang, J., and Wang, H. (2006). PEM Fuel Cell Open Circuit Voltage (OCV) in the Temperature Range of 23°C to 120°C. J. Power Sources 163, 532–537. doi:10.1016/j.jpowsour.2006.09.026

CrossRef Full Text | Google Scholar

Keywords: open-circuit potential, electrochemical concept, electric double layer, surface charge, potential of zero charge

Citation: Huang J and Zhang Y (2022) Essays on Conceptual Electrochemistry: I. Bridging Open-Circuit Voltage of Electrochemical Cells and Charge Distribution at Electrode–Electrolyte Interfaces. Front. Chem. 10:938064. doi: 10.3389/fchem.2022.938064

Received: 06 May 2022; Accepted: 21 June 2022;
Published: 25 July 2022.

Edited by:

Valentín Briega-Martos, Helmholtz Institute Erlangen-Nürnberg for Renewable Energy (IEK-11), Germany

Reviewed by:

Ricardo Alonso Martínez Hincapié, Max-Planck-Institut für Chemische Energiekonversion, Germany
Luis Enrique Botello, University of Alicante, Spain

Copyright © 2022 Huang and Zhang. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Jun Huang, ju.huang@fz-juelich.de

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.