Brought to you by:

A First-Principles Surface Reaction Kinetic Model for Hydrogen Evolution under Cathodic and Anodic Conditions on Magnesium

Published 27 July 2016 © 2016 The Electrochemical Society
, , Citation Christopher D. Taylor 2016 J. Electrochem. Soc. 163 C602 DOI 10.1149/2.1171609jes

1945-7111/163/9/C602

Abstract

A surface kinetic model is developed using a proposed mechanism for the anodic and cathodic branches of the hydrogen evolution observed over magnesium (i.e. the negative difference effect). The key element to the model is that hydrogen evolution requires the removal of adsorbed OH from the surface. This step is achieved in the cathodic branch by reductive desorption of OH to form OH from the surface, and in the anodic branch by the dissolution of both adsorbed OH along with the Mg atom to which it is attached (thus combined desorption of OH alongside dissolution of Mg2+). Steady state theory is applied to derive expressions for the hydrogen evolution rate as a function of potential. First-principles parameters obtained from the literature are used to compute the rate constants for the individual mechanistic steps, where available, and reasonable values based on cohesive energies are used to estimate the remaining parameters where first-principles information is not available. These rate constants are then applied to the surface kinetic model to show that the model provides reasonable agreement with the observed phenomenology for hydrogen evolution over magnesium, both in terms of the shape and order of magnitude of hydrogen evolved, and the transition potential (i.e. the open circuit potential) between the anodic and cathodic hydrogen evolution branches.

Export citation and abstract BibTeX RIS

A characteristic of corroding magnesium surfaces is the production of hydrogen under both anodic and cathodic conditions. The evolution of hydrogen under anodic conditions is atypical, since hydrogen evolution is typically a cathodic phenomenon, and, accordingly it has been given the name "the negative difference effect". This effect has been reviewed in the study of magnesium corrosion mechanisms by Song and Atrens1 and references therein. It is considered to be unusual because it is generally perceived that the hydrogen evolution reaction (HER) should be suppressed under anodic conditions, since the production of hydrogen is a reduction reaction (the alkaline form of the reaction is shown):

Equation ([1])

and therefore the reaction rate should decrease according to the expression:

Equation ([2])

where U0 is the reversible potential, k is the Boltzmann constant, T is the temperature and β is the Tafel slope. kHER,0 is the reaction rate at the equilibrium potential. That is, an increasing (i.e. more anodic) potential U should (according to standard electrochemical theory) lead to an exponential decay in the rate of hydrogen evolution.

Comparing this prediction, based on standard electrochemical theory, with the data presented by Frankel, Samaniego and Birbilis (2013),2 among many others, indicates that hydrogen evolution appears to follow this behavior up to a point (i.e. typical cathodic Tafel kinetics), but then, after the open circuit potential is traversed, becomes evolved by some yet unidentified anodic mechanism. This behavior is exemplified by the careful plots of the current associated with hydrogen evolution versus potential as shown by Frankel, Samaniego and Birbilis reproduced here with permission in Figure 1.2

Figure 1.

Figure 1. The current density associated with hydrogen evolution plotted against the applied potential for a 99.99% pure magnesium sample exposed to 0.1 M NaCl. Reproduced with permission from Corrosion Science (Elsevier, 2013).2

The same effect is shown when the hydrogen evolution rate is plotted against the applied current density, as reproduced from Samaniego, Frankel and Birbilis in Figure 2.3

Figure 2.

Figure 2. The volumetric hydrogen evolution rate plotted against the applied current density for several magnesium alloy systems, including pure Mg. Reproduced with permission from Ref. 3.

While Figure 2 demonstrates that additional alloying elements can amplify this effect, it is evident that even very pure Mg demonstrates the same phenomenology.

This non-standard electrochemical behavior of Mg corrosion was documented as early as 1953 by Petty, Davidson and Kleinberg4 and characterized primarily through a valence associated with Mg dissolution calculated by normalizing the net charge passed with the weight loss of between 1.38 and 1.66. This reduced net valence was rationalized by the dissolution of Mg as Mg+ ion, which was then hypothesized to react with water to form MgOH+ and H2, thus also accounting for the anodic evolution of hydrogen. This mechanism, however, has been in dispute (see for example work by Swiatowska et al. that reports a measured electrochemical stoichiometry of n = 25), since the Mg+ ion has never been directly characterized, and it seems unlikely to decouple the reaction sequence of Mg dissolution (which involves water complexation at the interface) to form Mg+ with the second step that generates H2. Further investigations by Samaniego, Hurley and Frankel repeating Petty's experiment, with the addition of in situ Raman spectroscopy, demonstrated that the Mg+ hypothesis was unsupported by Petty's original work.6

In 1961, Casey, Bergeon and Nagy considered the dissolution of Mg in MgCl2 solutions, and provided "a more explicit mechanism, which is easy to evisage" in comparison to the Petty univalent Mg model.7 The model has an advantage over the Petty model in that it gives more consideration to the roles of surface processes such as adsorption and the molecular chemistry of water, hydroxyl/hydroxide and hydrogen at the electrochemical interface. The scheme has four steps:

Equation ([3])

Equation ([4])

Equation ([5])

Equation ([6])

This mechanism has been used to support the behavior of the Mg reaction as a function of the H2 pressure in the experiments. This mechanism preserves some nature of this univalent dissolution of magnesium – through the desorption of a net single valent species MgOH+, that is Mg2+-OH - and has some bearing on the mechanism proposed in the present paper. Based on the recent ability to model surfaces with quantum chemical simulation, however, the formation of adsorbed species such as H2+ ads seems unlikely- molecular hydrogen readily dissociates on metal surfaces and is unlikely to exist as an adsorbed form.8 Furthermore, hydrogen adsorbed on metal surfaces almost always has a partially negative charge.8

Liu and Schlesinger9 also considered the Petty univalent Mg model, and attempted to provide a more quantitative analysis of the surface kinetics, using a microgalvanic approach and proposing some "factors" N1 and N2 that described the relative extent to which the univalent vs divalent dissolution of Mg occurred. It was left as an open question, however, as to how these factors should be determined. Presumably their geometric means, weighted by the valence of the Mg ion, would give the net valences for dissolving Mg observed by Petty. Liu and Schlesinger also proposed that models for Mg alloy corrosion should include microgalvanic effects due to the presence of electrochemically distinct species such as Zn and Al.

In 2012, Bender et al.10 proposed yet another explanation for the NDE in which the dissolving Mg (as Mg2+) reacts with water to dissociate and form OH, which precipitates as a surface film, and the protons from the dissociation of water produce H2. The reactions is postulated to take place entirely in the electrochemical double layer, thus not disturbing the bulk pH and equilibrium between water and its ions. The model was not quantitatively evaluated, however, and did not involve a mathematical treatment of the surface reactions and adsorption effects associated with the mechanisms. Furthermore, it has been mentioned by Rossrucker et al. that Mg2+ hydrolyses only weakly, throwing this mechanism into some doubt.11

Recent density functional theory investigations of the surface chemistry of water and hydrogen over magnesium surfaces have shed new light upon the physical and chemical details of the mechanisms for water dissociation and hydrogen dissociation and recombination at metal surfaces– mechanisms that should be highly relevant for understanding the microkinetic processes that undoubtedly must be contributing to the negative difference effect.12,13 Herein, a synthesis is made of the density functional theory studies in the literature, alongside a surface reaction kinetic model that elucidates the rates for hydrogen evolution and the state of actively corroding magnesium surfaces. The resulting mechanistic model draws from the ideas developed by the above cited authors, but tries to enrich the development of the models by performing a steady analysis of the reaction kinetics, using this newly available first-principles data as much as possible to infer the individual rate constants associated with each step. As the quantum chemical models continue to increase in their sophistication, and multiscale models permit inclusion of effects such as surface roughness and surface film formations, it is envisaged that models of this nature will be able to explain the rich corrosion phenomenology of Mg in even greater detail.

Proposed Mechanisms

The surface reaction kinetic model derived in this paper is built upon the following set of assumptions:

  • (1)  
    Water molecules dissociate upon adsorption on magnesium metal surfaces to produce adsorbed atomic hydrogen and adsorbed hydroxyl radical. (See Figure 3.) The exact charge states of these adsorbed species are not well defined. Quantum mechanical calculations for adsorption states indicate some localized charge transfer from the metal atoms on which they are adsorbed to the adsorbed states, but overall the reaction is charge neutral (no oxidation or reduction).13 This is Reaction 7.
    Equation ([7])
    The reaction, therefore, consumes two available metal sites Mg* and produces one surface site that has OH bound to it (Mg*OH) and one surface site with H bound to it (Mg*H).
  • (2)  
    The Mg*H sites may combine to produce hydrogen, H2, and making available the two metal sites Mg* (Figure 4). Note at this point that our mechanism has assumed that no charge transfer has yet taken place: it involves the recombination of neutral hydrogen atoms to make neutral H2.
    Equation ([8])
  • (3)  
    If Reactions 7 and 8 were to continue, the surface would eventually be covered by Mg*OH, allowing for no more hydrogen evolution. (Figure 5). However, there will be some rate at which the hydroxyl is removed from the surface by reduction to hydroxide (Figure 6a), making available once more the metal site Mg*:
    Equation ([9])
    It is during this step that the electron is consumed, making the total hydrogen evolution reaction through Equations 79 an electrochemical reduction reaction.At this point, we have been following the Tafel mechanism for hydrogen evolution reaction as elucidated in the theoretical work by Williams et al.13 This seems the most reasonable mechanism given the proclivity for Mg surfaces to reactively dissociate water and produce Mg*H and Mg*OH.12 Collectively the reaction should be responsible for the cathodic evolution of hydrogen, the reaction overall being:
    Equation ([10])
    with the states in [ ] representing the surface intermediates. The intermediate step, R3, is the one in which the electron transfer occurs, and will become slower and slower as potentials move in the anodic direction, thus leading to overall Mg*OH poisoning of the HER at anodic potentials. However, this can only account for the cathodic branch of the hydrogen evolution observed in Figures 1 and 2, so the proposed mechanism in Equations 79 cannot be the whole story.To explain the anodic branch, we now propose that one more process must be occurring:
  • (4)  
    Mg*OH sites on the surface can alternatively be removed through anodic dissolution of Mg (Figure 6b). We assume here that sites where OH is adsorbed are the most vulnerable to dissolution, given that there is already some partial charge transfer from the Mg to the OH moiety to which it is attached, and the formation of a Mg-OH bond will tend to weaken adjacent metal-metal bonds.13 There are a number of possible representations for this anodic dissolution step; we choose here the simplest one, in which Mg2+ is liberated, along with the OH ion and the release of an electron (note that Mg2+ dissolution is a 2-electron process- here we assume the second electron is transmitted through the desorption of OH):
    Equation ([11])
    The final term, Mg*, indicates that removing the Mg*OH will leave a freshly exposed Mg* site (initially Mg(m) where (m) stands for metal) which is available for further reaction according to Equation 7. Since dissolving surfaces will not be atomically smooth, but quite rough, we leave aside the details as to how the vacancy is dealt with, but rather, stick with the simplistic description for the sake of developing an analytical model (see further below). Also, the dissolution of Mg from sites that do not have OH adsorbed can occur in parallel with Reaction 11 and, as will be shown later on, does not have a mathematical effect on the details of the current model. This concept is essentially very similar to that proposed by D. Williams and R. Newman in a 2015 Faraday Discussion - "a fast-dissolving metal has more (albeit momentarily) bare metal sites which are the sites of hydrogen evolution" - although herein the detail is added that this route is necessary to overcome the otherwise difficulty of eliminating the adsorbed OH species that would prevent further reaction.14
Figure 3.

Figure 3. Water adsorption on Mg surface (Left). For example, water adsorption over an Mg atom (i.e. the atop site) on an Mg(0001) surface is shown using molecular rendering. Water adsorption on the close-packed surface has been found to be dissociative, producing OH and H adsorbed on to three-fold hollow sites on the surface (Right).

Figure 4.

Figure 4. Through diffusion H atoms can encounter one another in adjacent sites, and, if they possess significant thermal energy to cross the activation barrier, may recombine to produce H2.

Figure 5.

Figure 5. If reactions 7 – water dissociation – and 8 – hydrogen recombination – proceed, without OH adsorption, the surface will become saturated (or poisoned) with *OH, preventing further dissociative adsorption of water and production of H2.

Figure 6.

Figure 6. In this paper we propose that the necessary OH desorption to allow hydrogen evolution could be CATHODIC (a, Reaction 9, Left)- by adsorbed OH accepting electrons and desorbing as OH-, or ANODIC (b, Reaction 11, Right)– by facilitating the dissolution of an Mg atom to which it is attached. This may be one of the terrace atoms, as shown here, or – more likely, given that corroding surfaces will not be atomistically smooth – an Mg atom at a step, kink or an adatom.

As will be shown, a reaction kinetic analysis – using reasonable activation energies and Arrhenius kinetics obtained from the theoretical literature and standard reference databases – that consists of Equations 79,11 and the law of conservation (the total number of surface sites must be conserved) can quantitatively account for the negative difference effect that is visible in both of Figures 1 and 2. Of omission from the mechanistic account is the formation of magnesium hydroxide films (distinct from metallic magnesium surface covered by a full or partial monolayer of adsorbed OH, Mg*OH), suggested to also be relevant to hydrogen evolution by Salleh et al.15 We assume that this pathway is slower than the dissolution and HER kinetics, although certainly a treatment that considers the role of hydroxide in more detail could be made in a future work- especially given that theoretical data for the interactions of water and hydrogen with Mg(OH)2 surfaces are also being newly made available.12,16 Some preliminary discussion of this point based on theoretical data points for water dissociation and the formation of OH vacancies through OH dissolution from the Mg(OH)2 surface will be made toward the end of this paper.

Reaction Kinetics and the Steady State Assumption

Proposing a mechanism as in Reactions 79, 11 provides a qualitative rationale that might explain the negative difference effect, but some reasonable, quantitative model needs to be developed in order to see if the proposed mechanism is truly viable. To begin constructing the quantitative model we note that, under the proposed mechanism, the surface sites will consist of available Mg sites Mg*, hydroxylated sites Mg*OH, and hydrogenated sites Mg*H. The model in its current form only considers the actively dissolving part of the magnesium surface, and so does not consider those surfaces that have been passivated by thin films of Mg(OH)2. This neglect is not meant to assume that these surface films are not important, but that they have not yet been included in the current model, and, we are assuming implicitly that the passivation to form the Mg(OH)2 film occurs at a time-scale much slower than the Reactions 79,11. Another type of surface site not considered herein are those that may be occupied by anions, like chloride. The role of chloride in affecting surface properties will be treated in a separate publication which uses density functional theory calculations for adsorption energies of different species on Mg surfaces to construct a surface phase diagram for magnesium in aqueous media.3,17,18

With these points in mind, we will continue for now to only focus on Mg*, Mg*OH and Mg*H, and the conservation law:

Equation ([12])

where θM, θH, and θOH stand for the surface fractional coverage of metal sites, hydrogenated sites, and hydroxylated sites, respectively.

In addition, we have the kinetic rate equations (K1–K4) for Reactions 79,11:

Equation ([13])

Equation ([14])

Equation ([15])

Equation ([16])

where the upper case "Ki"s are the reaction rates, and the lower case "ki"s are rate constants. The rate constants for the electrochemical reactions k3 and k4 will be potential dependent, following Tafel type relations:

Equation ([17])

Equation ([18])

The sign of the exponential differs from k3 to k4 due to the switch from a cathodic elimination of OH to an anodic elimination of OH conjoint with Mg dissolution. Under standard conditions (i.e. concentrations of 1 M for OH and Mg2+ and at standard temperature and pressure), U0 for k3 will be −0.83 V NHE, the equilibrium potential for hydrogen evolution, and U0 for k4 will be −2.38 V NHE, the equilibrium potential for magnesium dissolution.19 To simulate the effect of different pH and background Mg2+ concentrations we can shift the potentials using the Nernst equation. Within the treatment that follows we use [Mg2+] = 2×10−5 M and pH 11 based on the solubility product of Mg(OH)2 and typical pH resulting from magnesium dissolution.20 We will assume a typical β value of 0.5, although, at some point, it should be feasible to determine β entirely from first-principles.

The kinetic rate equations (K1–K4, i.e. Equations 1316) are expressions for the rate of change in the consumption and production of the reaction sites, quantitatively expressed by the fractional surface coverages θM, θH, and θOH. Thus, they contain the differential equations that control the evolution of θM, θH, and θOH over time. For example, according to Equation 7, θM will decrease at twice the rate at which θH and θOH increase. However, the differential equations for these variables will be tightly coupled. The differential equations resulting from the simultaneous occurrence of Reactions 79,11 are:

Equation ([19])

Equation ([20])

Equation ([21])

Since the adsorbed surface species are intermediates to the hydrogen evolution reaction, as illustrated in Equation 10, the steady state approximation can be applied- that is, that the rates of change captured in these three differential equations should, on average, go to zero. Here, again, it is appropriate to remember that corroding Mg surfaces are eventually protected by the formation of a Mg(OH)2 film, however, this film is observed to form behind the evolving active corrosion front (see work by G. Williams),21 hence we assume this is at a time scale several orders of magnitude away from the events captured in Reactions 79,11. In fact, it seems likely that this film might form following the reaction sequence 79,11, as a consequence of the re-precipitation of Mg(OH)2 once a critical pH and magnesium concentration are reached in the near-surface environment.

When we apply the steady state assumption, it becomes possible to solve for the steady state coverages θM, θH, and θOH and hence the reaction rates K1–K3, Equations 1315. The reaction rate K2 is the rate corresponding to hydrogen evolution reaction (i.e. the observable volumetric evolution rate shown in Figure 2). The algebraic expression for K2 resolves to:

Equation ([22])

Equation ([23])

Equation ([24])

The variables α and γ are introduced for simplicity, and the somewhat complex form for the rate equation arises due to the necessity of solving a quadratic equation when the steady state approximation is applied to the differential equations. The variable α can be related to the relative rate of elimination of the Mg*OH sites relative to hydrogen recombination, and the variable γ can be related to the relative rates of hydrogen recombination compared to water dissociation.

Having now presented the basis for, and the results of, the derivation of the hydrogen evolution rate, K2, as a function of the fundamental rate constants k1-4, we will now use a first-principles approach to quantify the rate constants, and hence see if the proposed combination of mechanisms can account for the negative difference effect of hydrogen evolution on pure magnesium.

Estimation of Rate Constants from First Principles

While it is not straightforward to estimate the rate constants, k1–k4, we can in fact make some reasonable guesses by applying the Arrhenius equation which resolves to the requirement of estimating a prefactor A and an activation energy Ea:

Equation ([25])

Reaction 7 is the dissociation of water over the magnesium surface. According to the density functional investigations by Dr. Williams, the barrier to water dissociation over a pair of vacant magnesium sites is 1.06 eV calculated with explicit solvent molecules.c,12 The value for the prefactor A is taken to be 1.0 × 1013, i.e. the standard "attempt frequency" based on the typical vibrational frequency for atoms and molecules in the condensed phase.22

Similarly, the barrier to hydrogen recombination (Reaction 8) was computed to be 1.04 eV.12 Other theoretical estimates for this reaction have a similar barrier.23,24 According to Pozzo, the attempt frequency for hydrogen recombination is around 1.0 × 1014.23

Reaction 9 is the cathodic desorption of hydroxide. This reaction will be electrochemical in nature, and so the Arrhenius term should be coupled with the Tafel equation. The activation barrier for desorbing hydroxide can be estimated using some reasonable physical arguments. The initial state- adsorbed hydroxyl- is likely to have a semblance of the solvation shell, so we assume it has approximately 1/2 of the solvation energy in the initial state. During the transition to a completely solvated hydroxide ion, the hydroxyl must first desorb, which has an estimated energy of 5.0 eV.d The hydroxyl species must then receive an electron to become OH-; the energy associated with this is 3.7 eV, from experimentally determined electron affinity of OH.25 At the equilibrium potential for hydrogen evolution this value should be modified to the value 3.7−4.44 = −0.74 eV, since the absolute potential of the hydrogen evolution reaction is 4.44 V.26 The hydroxide ion then receives the rest of its solvation shell; half of the solvation enthalpy of OH- is 2.4 eV, using the enthalpies tabulated by Marcus.27 Put altogether, this gives an estimated barrier to the cathodic desorption of OH of 1.9 eV. The standard prefactor of 1.0 × 1013 is used here, and the Tafel equation with a β value of 0.5 applied.

The final reaction step considered in the scheme is the dissolution of the magnesium ion. We estimate the barrier using the cohesive energy of Mg, 1.51 eV.28 Density functional theory studies of the dissolution pathway for a copper adatom on copper showed that the early stage of dissolution was associated with the breaking of the metal bonds,29 and hence, the cohesive energy should serve as a reasonable estimate for the activation barrier. Alternative estimates and/or values derived from detailed studies made using density functional theory could be made in future work. The standard prefactor of 1.0 × 1013 is used. The Tafel relation is combined with this equation, using a β value of 0.5, as discussed earlier in the derivation.

Implementation of the Model

The rate of hydrogen evolution, K2 (Equation 14), can now be determined by using the rate constants estimated in the preceding section to determine the constants α and γ. α is a function of potential, and so we calculate α for potentials ranging between −2.8 and −1.2 V NHE. Consequently, we arrive at a HER that has two branches: a cathodic branch, and an anodic branch. see Figure 7.

Figure 7.

Figure 7. (a). Cathodic (short dash) and Anodic (long dash) branches and the sum of cathodic and anodic effects of the HER (solid line) determined by Equation 22. (b). Plotted side-by-side with the data from Ref. 2. The model curve "this work" has been scaled so that the asymptotic value corresponds to the maximum rate in the experimental curve, due to the uncertainty surrounding the rate constant pre-factors in the first-principles model.

The model data plotted in Figure 7a has a strong resemblance in structure and a local minimum at similar potential to the experimental data shown in Figure 12 highlighted here in the side by side plot shown in Figure 7b. Since the pre-factor is estimated to within an order of magnitude in the rate constants, we only show the relative rate scaled to the asymptotic HER values. As seen, there is a difference in the absolute potential at which the current drops to a minimum, as well as some differences in the steepness of the curve. These differences could be due to a multitude of factors, some of which may include (a) errors in the assumed rate constants, (b) neglecting the effect of other species (chloride, for instance), (c) ignorance of any microstructure effects, such as microgalvanic coupling within the model developed herein, (d) neglect of solid-state (rather than adsorption only) films of MgO and Mg(OH)2 as well as mixed oxy-hydroxides, and last but not least (e) faulty assumptions in the expressions used to derive the model of equations 79,11.

One key feature to note in Figure 7a is the plateau in the hydrogen evolution rate at the cathodic and anodic extremes plotted. This plateau is a consequence of the reaction limitation regarding the kinetics of water dissociation and hydrogen recombination. That is to say, when the potential is sufficiently negative, the OH no longer poisons the surface due to cathodic desorption of OH, and so the reaction can proceed as fast as the surface can dissociate water, and/or as fast as the hydrogen atoms can recombine. At the positive potentials, the OH poisoning is remedied by the anodic desorption of MgOH+ (or some other synergistic combination of Mg2+ + OH, the quantum chemical calculations have not yet made that clear) and hence the same scenario also applies. The graph is also symmetric because the two reactions (cathodic desorption of OH and anodic desorption/dissolution of Mg-OH) are both one-electron steps and we have assumed the same Tafel constants in our model. That symmetry may break as further first-principles or well-resolved experimental data can be obtained. In reality, the precipitation or solid-state formation of Mg(OH)2 films on the surface will significantly perturb the shape of these curves as anodic potentials are encountered; that is beyond the scope of this interpretive, analytical model. It should also be noted that the curves are scaled to these plateau values due to the difficulty in calculating pre-factors for rate constants using first-principles.

To highlight the importance of surface adsorption effects on controlling the reaction behavior, in Figure 8 are shown the relative surface coverages of free metal (θM), hydroxide (θOH) and hydrogen (θH) that can be extracted from the steady state analysis performed using the rate equations derived in this paper. The coverage is shown as a function of the electrochemical potential. It is seen that the hydroxide coverage reaches a maximum at the open circuit potential, and then decreases as anodic or cathodic polarization occurs. The free metal sites (θM) maintain around 15% coverage, except for in the vicinity of the OCP, and it can be presumed that the conventional 2-electron dissolution of Mg can proceed at these sites following the usual Butler-Volmer kinetics. It could be that a combination of this 2-electron dissolution from free metal sites alongside the 1-electron dissolution at the OH-adsorbed sites is responsible for the non-stoichiometric net valence for Mg reported by Petty et al.,4 and further advanced by Atrens.34

Figure 8.

Figure 8. First-principles prediction of fractional, or relative, surface coverage (ranging from 0–1) for free metallic sites (θM, no adsorbate), versus hydroxide adsorption (θOH) and hydrogen adsorption (θH), as a function of potential.

The hydrogen evolution rate can also be plotted against the applied current density, as was done for the experimental data shown in Figure 2. The theoretical version, obtained using the first-principles rate constants in this work, is reproduced in Figure 9.

Figure 9.

Figure 9. Model predicted hydrogen evolution rate, K2, as a function of applied current density.

Once again, the phenomenology between Figures 2 (from experimental data) and Figure 9 (the current model) is in considerably good agreement, indicative that the proposed model, grounded in the latest density functional theory studies of fundamental magnesium surface chemical properties, provides a potentially useful model for accounting for the negative difference effect.

Further exploration of the phenomenology surrounding the role of surface orientation, defects (i.e. vacancies, steps, kinks, ...), the dependence of adsorption isotherms with changes in the coverage of OH, O or H, and alloying elements should shed further light on the details of this mechanism, and could be made using a first-principles approach such as density functional theory. As more details are provided through either theory or experimental characterization, the simplistic analytical model provided in this paper would need to give way to 3-dimensional simulations using techniques such as kinetic Monte Carlo that can take into account the evolving shape of the surface, improving the realism of the model, but at the same time taking away from the intuitive benefits the analytical model derived herein provides.30,31

Further Questions to Address

At present the estimated rate constants do not allow for precise prediction of the absolute rates of hydrogen evolution or the anodic/cathodic current densities associated with the change in potential. Pre-factors can be more accurately determined from first-principles when certain kinds of accelerated molecular dynamic simulations are employed, for example.32

An additional area of concern is that the formation of hydroxide films is not taken into account in this model- rather, we assume that the hydroxide films take a longer time scale to form compared to the events that lead to the fast moving corrosion front in magnesium alloys. It might be the case that, even when the system is cathodically polarized, a nanoscale hydroxide film is present on the material, and, as shown recently, this hydroxide film might have considerable activity for HER.15

There are some density functional theory studies for hydrogen recombination and water dissociation on magnesium oxide surfaces with and without defects that might provide some useful starting points for estimating rate constants.16,33 Those estimated rate constants could then potentially be used in a surface reaction kinetic model that uses a model similar to that derived in this work to provide analogous predictions for the activity of a surface that has the Mg(OH)2 structure, with defect sites providing the place of Mg* as potential hosts for H* and OH* obtained from water dissociation and serving as the catalytic points for hydrogen recombination. The mechanisms may be slightly different due to the alternative ways in which water, hydrogen and hydroxyl/hydroxide can interact with ideal, stepped and defective surfaces of the oxides and hydroxides. Example steps to add to the proposed mechanism include the formation of defects due to Mg2+ or Mg(OH)+ dissolution, and the healing of defects via interactions with water, possibly involving the concomitant H2 evolution, see, for example Ref. 34.

We can make already some rough estimates for how processes may be different for the hydroxide film by looking at theoretical calculations that have already been performed on Mg(OH)2 and MgO surfaces. Water dissociation on Mg(OH)2 has been shown to have a barrier of 27.6 kcal/mol (1.2 eV, similar to the value calculated by Dr. K. S. Williams for Mg(0001), as used in this work) and OH dissolution of 43.2 kcal/mol (1.9 eV, essentially equal to the value used here for Mg surface).16 The hydrogen recombination barrier on a defective MgO surface is 0.61 eV,33 suggesting that the overall rate at which these processes occur might be faster on an oxidized surface. Using these rough numbers for oxidized surfaces gives outputs that are entirely analogous to those shown in Figures 7 and 9.

A Note on Univalent Mg

The theory concerning the formation of Mg+ as a precursor to Mg2+ evolution34 and its tentative placement on the traditional Pourbaix diagram35 may have some fundamental origin in the proposed reaction step 11, which we recount here:

Equation ([26])

This reaction has the "appearance" of being a one-electron dissolution process, but, in reality, there is a second electron that we consider to have transferred to OH during the dissolution event. One could also write a possible pathway for Mg dissolution as:

Equation ([27])

which is effectively the same, but we consider the OH to be a ligand that supports the dissolution of Mg (similarly to Bender et al.10 and Casey et al.7). This microkinetic reaction step could be occurring, and would give the "appearance" of a univalent Mg as an intermediate to Mg dissolution. Since it is conjectured that 2-electron electrochemical steps are less likely (and therefore kinetically much slower) than 1-electron steps, it is possible that the dissolution of Mg occurs in this way rather than as the direct 2-electron generation of Mg2+.

The model can also be adapted to include a second route for dissolution of Mg2+ from the sites that do not have OH adsorbed. These are the free sites represented in the surface kinetic model as θM. Including the rate of dissolution for Mg2+ from these sites could be coupled into the kinetic model, in which case Mg2+ dissolution will occur in parallel with the processes occurring here, but mediated according to the extent of relevant surface sites, see Figure 8. Many of the macroscopic observations occurring in experimental tests will be a combination of the mechanisms occurring in this work, alongside features such as the precipitation of Mg(OH)2, and the formation of distinct anodic and cathodic sites on the surface, that cannot be well-represented in the simplified mathematical model of this work. However, incorporation of all these effects in a wider multiscale model could well be achieved in a well-coordinated and collaborative effort as future work. As suggested above, one route could be to use kinetic Monte Carlo. Another route could be to use cellular automata.36

Summary

A reaction scheme has been presented whereby H2 can be generated on Mg via Tafel recombination of hydrogens obtained by the chemical (i.e. not electrochemical) dissociation of water on Mg, and the "cathodic consumption" of electrons is attributed to the desorption of OH. Under anodic potentials, this cathodic consumption is disallowed and so HER would otherwise stop due to OH poisoning of the surface. We propose that, in the case of Mg, the OH can be removed anodically through the dissolution of the Mg atom to which the OH is adsorbed – assuming that OH might be facilitating the dissolution, leading to a one-electron dissolution pathway. This effectively creates fresh surface whereby more water can react with Mg producing H and OH. Using first-principles data from density functional theory and reference tables to estimate rate constants, we show that the proposed pathway can account for both cathodic and anodic HER, and the theoretically predicted phenomenology is in remarkable agreement with the measured relations between HER and potential and HER and applied current density. While there are many theoretical questions that remain to be answered, it appears that the proposed model might reasonably hold the key to understand the negative difference effect on Mg.

Acknowledgments

The author gratefully acknowledges several exceptionally beneficial and stimulating discussions that led to the development of this model and the improvement of this work with Dr. Kristen S. Williams (Boeing), Michael F. Francis (EPFL) and Gerald S. Frankel (Ohio State University). The recommendations and input from peer-reviewers was also gratefully received and benefited the paper. The Ohio State University is gratefully acknowledged for supporting this work.

Footnotes

  • Depending on the way the calculation is carried out, the barrier can range from 0.9 to 1.8 V.13

  • This value can be derived from the data reported by Dr. K. S. Williams where water dissociation to adsorbed OH* and H* over Mg(0001) has a value of −2.27 eV12 and subtracting out the energy associated with the hydrogen atomic adsorption and the water dissociation contribution in vacuum.

Please wait… references are loading.