Home Physical Sciences Electro-chemo-mechanical properties of anodic oxide (passive) films formed on Cu, Ni and Fe
Article Publicly Available

Electro-chemo-mechanical properties of anodic oxide (passive) films formed on Cu, Ni and Fe

  • Masahiro Seo

    Masahiro Seo is an emeritus professor at Hokkaido University, Sapporo. He received his PhD degree in applied chemistry in 1971 from Hokkaido University. He joined as a research associate in 1972 at the Fontana Corrosion Center of the Ohio State University. Dr. Seo was a professor (from 1990 to 2005) at Hokkaido University. His research interests are focused on aqueous corrosion and passivation processes of metal thin films, measurement of stresses generated during anodic oxidation of transition metals, and inhibition effects of underpotential deposition on aqueous corrosion of metals. He is an ECS fellow and an honorary member of Japan Society of Corrosion Engineering.

    EMAIL logo
Published/Copyright: December 4, 2023

Abstract

Electro-chemo-mechanical properties of anodic oxide (passive) films formed on metals have been reviewed focusing on the results of stress variations caused by anodic oxidation of Cu, Ni, and Fe thin film electrodes in deaerated pH 8.4 borate buffer solution at 25 °C. The surface stress varies toward compressive direction due to adsorption of OH on Cu from aqueous solution as well as adsorption of oxygen on metals from gas phase. The stresses are generated with the growth of three-dimensional anodic oxide films on metals. The magnitude and sign (tensile or compressive) of the intrinsic film stress were determined by taking the residual stress of the substrate and the dielectrostriction into consideration. The tensile or compressive intrinsic film stress depends on p-type or n-type semiconductive properties of the anodic oxide films, which is explained in terms of the void formation or oxide formation in the metal side at the metal/film interface. Furthermore, the stress variation toward compressive direction during cathodic reduction of the anodic oxide films is explained in terms of the volume expansion due to the formation of intermediate species.

1 Introduction

Passivity of metal is due to the formation of protective oxide film (i.e., passive film) on the metal. Recent application of in-situ scanning tunneling microscopy (STM) to corrosion studies (Kunze et al. 2001; Marcus and Maurice 2017; Maurice et al. 2000; Seyeux et al. 2005; Strehblow et al. 2001; Zuili et al. 2000) has revealed that adsorption of oxygen or OH on a single crystal metal such as Cu and Ni in aqueous solution induces a structural change or reconstruction of a topmost surface to form a superlattice-structure as a two-dimensional precursor for a three-dimensional passive film. Electro-sorption of electrolyte ions on a solid metal in aqueous solution alters a surface stress of the metal (Haiss et al. 1998), accompanying a structural change to keep the surface more stable, implying that the surface stress is closely linked to the stability of the surface structure.

A surface stress is different from a surface tension for a solid surface (Ibach 1997; Seo 2020). A surface stress is a reversible work per area to deform a surface elastically (i.e. by changing the surface atomic density), while a surface tension is a reversible work per area to deform a surface plastically (i.e. keeping the surface atomic density constant). The unit (J/m2 or N/m) of a surface stress is equal to that of a surface tension. Although the unit of N/m is conventionally used for surface stress and surface tension, the dimension of J/m2 is employed in this article. A surface stress is a tensor quantity, while a surface tension is a scalar. For a solid surface with a three-fold or higher rotation axis of symmetry, a surface stress is isotropic and can be taken as a scaler. When a liquid surface such as Hg electrode is stretched, the substrate atoms move easily to the surface to keep the surface atomic density constant. As a result, for a liquid surface, a surface stress is equal to a surface tension.

On a clean metal surface in vacuum, surface atoms miss the bonds because of the absence of atoms in the vacuum side of the surface plane. The electronic charge accumulated around surface atoms contracts the equilibrium length of lateral bonds between surface atoms due to a strong electrostatic attractive force, which induces a tensile stress (Ibach 1997). The surface stress on a clean metal surface, therefore, is always tensile irrespective of the surface structure. The chemisorption of foreign atoms on the metal surface re-builds the bonds between metal atoms and adsorbate atoms on the surface. If the adsorbate is electronegative with respect to the substrate metal, the electronic charge is removed from the bonds between the surface metal atoms to reside on the adsorbate atoms. The removal of the electronic charge between surface metal atoms reduces the tensile stress and even alters the surface stress to be compressive (Ibach 1997). Conversely in the case where the adsorbate atoms are electron donors, the electronic charge between surface metal atoms is enhanced to increase the tensile stress (Ibach 1997). The sign of the surface stress is defined as plus for tensile and minus for compressive, respectively in accordance with the sign convention of bulk stresses.

Although a surface stress for a liquid is equal to a surface tension, a solid has two different physical quantities of surface stress and surface tension (Ibach 1997; Seo 2020). The relationship between surface stress g and surface tension γ for an isotropic solid is represented by the Shuttleworth equation (Shuttleworth 1950):

(1)g=γ+γε,

where ε is the surface elastic strain. Furthermore, the derivative of surface stress g with respect to electrode potential E under constant elastic strain for a solid electrode in electrolyte solution is given by Gokhshtein (1976):

(2)(gE)ε=q(qε)E,

where q is the surface charge density of the electrode. Equation (2) means that the electrocapillary maximum of g does not coincide with potential of zero charge (pzc) corresponding to the electrocapillary maximum of γ (i.e. q = 0) since the second term qε of the right-hand side in Equation (2) is not zero.

When an anodic oxide film is formed and grown on a metal electrode during anodic polarization, a stress is generated in the film (Bradhurst and Leach 1966; Nelson and Oriani 1993; Ueno et al. 2000; Vanhumbeeck and Proost 2008a). The stability of the anodic oxide film or passive film depends on the sign and magnitude of the stress generated in the film. A film stress is defined by the stress divided with the film thickness. The unit of the film stress, therefore, is J/m3 or Pa. The product of the film stress and film thickness corresponds to the dimension of the surface stress (J/m2). The properties of solid electrode surfaces associated with correlation between electrochemical and mechanical phenomena are named “electro-chemo-mechanical (ECM) properties” (Seo 2020). The study on electro-chemo-mechanical properties of passive metal surfaces is important for better-understanding of the growth and stability of the passive film. However, there have been few corrosion studies from the electro-chemo-mechanical viewpoint. A cantilever bending method (Ibach 1997; Láng and Barbero 2012; Seo 2020) has been used for measuring the surface stress and film stress. The cantilever bending method can measure the stresses generated on the one side of a rectangular thin plate from the changes in curvature of its plate. The principle and instrumentation of the cantilever bending method have been explained elsewhere (Ibach 1997; Láng and Barbero 2012; Seo 2020). The details of the cantilever bending method are abbreviated in this review.

The surface stress variation due to oxygen adsorption on a single crystal metal such as Cu (Harrison et al. 2006), Ni (Ibach 1997), and Pt (Ibach 1997) has been measured and compared with that (Feibelman 1997; Harrison et al. 2006; Hong et al. 2004) obtained theoretically. Unfortunately, no experimental results of the surface stress variation due to OH adsorption on metals in aqueous solution have been reported so far. The surface stress variation due to OH adsorption on Cu in aqueous solution will be firstly reported and the factors controlling the surface stress variation will be discussed from the ECM viewpoint.

The stress evolution during anodization of a valve metal such as Al (Bradhurst and Leach 1966; Nelson and Oriani 1993), Ti (Ueno et al. 2000; Vanhumbeeck and Proost 2008a), and Zr (Van Overmeere and Proost 2010) has been frequently measured to evaluate the stress (i.e. film stress) generated in the anodic oxide film. The magnitude and sign of the film stress depend on many parameters such as growth mechanism, electric field, and defective structure. The results of the stress evolution for the growth of anodic oxide (passive) films on Cu, Ni and Fe in pH 8.4 borate buffer solution at 25 °C will be reviewed and the major parameters which influence the film stress will be discussed from the ECM viewpoint. Furthermore, it will be shown that the stress variation during cathodic reduction of the anodic oxide films is directly associated with the cathodic reduction mechanism. The borate buffer solution has been employed for many studies (e.g. Sato and Kudo 1971; Sato et al. 1971; Sato et al. 1976; Seo et al. 1988) of passive films because stable anodic oxide films are formed in the solution. Physicochemical properties such as composition, thickness and defective structure obtained for anodic oxide films in the borate buffer solution are indispensable for quantitative analysis of the film stress. It is remarked that all electrode potentials described in this review are referred to the standard hydrogen electrode (vs. SHE).

2 Surface stress variation due to adsorption of oxygen or OH on single crystal metals

2.1 Oxygen adsorption from gas phase

Figure 1 shows the typical surface stress variation due to oxygen adsorption on single crystal Ni (111) from gas phase: the value of surface stress g is plotted as a function of oxygen coverage θ at 77 °C (Grossmann et al. 1995; Ibach 1997). It is noted that g in the ordinate of Figure 1 is referred to zero at θ = 0 since an absolute value of g cannot be measured by the cantilever bending method (Ibach 1997; Láng and Barbero 2012; Seo 2020). The surface stress g decreases almost linearly toward compressive direction with increasing θ up to about 0.25. Single oxygen atom occupies the face-centered cubic (fcc) three-fold hollow site on Ni (111) to form a p (2 × 2)structure of the oxygen overlayer at θ = 0.25. The g versus θ curve of Figure 1 deviates downward from the linearity at about θ = 0.3, which is responsible for the phase transition of the oxygen overlayer from a p(2 × 2) to a (3×3)R30° structure. The oxygen coverage corresponding to the (3×3)R30° structure is θ = 0.33.

Figure 1: 
						Typical surface stress variation due to oxygen adsorption on single crystal Ni (111) from gas phase: the value of g is plotted as a function of oxygen coverage θ at 77 °C for oxygen adsorbed on single crystal Ni (111) (Grossmann et al. 1995; Ibach 1997). The value of g in the ordinate of Figure 1 is referred to zero at θ = 0. Copyright ©1997. Reproduced with permission from Elsevier.
Figure 1:

Typical surface stress variation due to oxygen adsorption on single crystal Ni (111) from gas phase: the value of g is plotted as a function of oxygen coverage θ at 77 °C for oxygen adsorbed on single crystal Ni (111) (Grossmann et al. 1995; Ibach 1997). The value of g in the ordinate of Figure 1 is referred to zero at θ = 0. Copyright ©1997. Reproduced with permission from Elsevier.

The fcc three-fold hollow site on Ni (111) is the most stable for single oxygen atom adsorption. The oxygen binding energy (−2.18 eV) of the hexagonal closest packing (hcp) three-fold hollow site on Ni (111) is less than the oxygen binding energy (−2.26 eV) of the fcc three-fold hollow site (Olatunji-Ojo and Taylor 2013). The structural analysis of the oxygen overlayer on Ni (111) (Schwennicke and Pfnür 1996), however, has indicated that the oxygen atoms partially reside in the hcp three-fold hollow sites at oxygen coverages between θ = 0.25 and 0.33. The downward deviation of the g versus θ curve from the linearity at about θ = 0.3 in Figure 1 is attributed to the oxygen atoms residing in the hcp three-fold hollow sites (Ibach 1997). It is deduced that the oxygen atoms in the hcp three-fold hollow sites contribute largely to the surface stress variation toward compressive direction as compared to the oxygen atoms in fcc three-fold hollow sites (Ibach 1997).

The surface stress varies toward compressive direction due to oxygen adsorption on Ni (100) as well as on Ni (111) (Grossmann et al. 1995; Ibach 1997). Single oxygen atom occupies the four-fold hollow site on Ni (100) to form a c (2 × 2) structure of the oxygen overlayer at θ = 0.5. The surface stress variation measured at 77 °C for the c (2 × 2) oxygen overlayer (θ = 0.5) on Ni (100) is Δg = −5.35 J/m2 which is significantly larger than that (∆g = −1.15 J/m2) for the (3×3)R30° oxygen overlayer (θ = 0.33) on Ni (111) (Grossmann et al. 1995; Ibach 1997). The surface stress variation measured at 300 K for the c (2 × 2) oxygen overlayer (θ = 0.5) on Cu (100) is ∆g = −1.0 J/m2 (Harrison et al. 2006), which is much less than that for the c (2 × 2) oxygen overlayer on Ni (100).

Although the cantilever bending method is unable to measure an absolute value of the surface stress of oxygen-overlayered surface as well as clean metal surface, its absolute value can be estimated from first-principles calculations based on density-functional theory (DFT) (Feibelman 1997; Harrison et al. 2006; Hong et al. 2004). Table 1 shows the comparison between experimental and theoretical values of the surface stress variations Δg due to the formation of oxygen overlayers on single crystals Cu, Ni and Pt from gas phase. The values of Δg in Table 1 are taken as the difference between the oxygen-overlayered and clean metal surfaces. For Cu (100), the theoretical value of ∆g is larger than the experimental value, while for Ni (100), the situation is reverse. At present, the causes listed for the above reverse situation have not been made clear (Harrison et al. 2006; Hong et al. 2004). Nevertheless, it is clear from Table 1 that the formation of oxygen overlayers on single crystal metals induces the surface stress variation toward compressive direction.

Table 1:

Surface stress variations ∆g due to the formation of oxygen overlayers on single crystals Cu, Ni, and Pt from gas phase.

Substrate metal Oxygen overlayer superstructure Δg (J/m2)a
Experimental Theoretical
Cu (100) c(2 × 2) −1.0 (Harrison et al. 2006) −3.07 (Harrison et al. 2006)
Cu (100) ( 2 × 2 2 ) R 45 ° −0.6 (Harrison et al. 2006) −2.03 (Harrison et al. 2006)
Ni (100) c(2 × 2) −5.35 (Ibach 1997) −1.6 (Hong et al. 2004)
Ni (111) p(2 × 2) −0.75 (Ibach 1997)
Ni (111) ( 3 × 3 ) R 30 ° −1.15 (Ibach 1997)
Pt (111) p(2 × 2) −4.0 (Ibach 1997) −3.2 (Feibelmann 1997)
  1. aThe values of ∆g are taken as the difference in g between the oxygen-overlayered and clean metal surfaces.

Since oxygen is electronegative with respect to the substrate metal, the oxygen adsorption relaxes the tensile stress and leads to the surface stress variation toward compressive direction. Carbon is a strong adsorbate for Ni (100) as compared with oxygen. It has been observed that the carbon adsorption on Ni (100) induces the surface reconstruction from a c (2 × 2) to a (2 × 2) p4g structure, in which the nearest neighbor distance of Ni atoms becomes larger in the reconstructed phase (Hong et al. 2004; Ibach 1997). The surface stress variation toward compressive direction due to the carbon adsorption on Ni (100) is initially much steeper than that due to the oxygen adsorption on Ni (100) (Ibach 1997). The surface stress variation, however, levels off to nearly constant at a coverage of θ ≈ 0.3, corresponding to the onset of the reconstruction (Ibach 1997). The level-off of the surface stress variation means the relaxation of the compressive stress. The increase in the nearest neighbor distance of Ni atoms due to the surface reconstruction induces the surface stress variation toward tensile direction and contributes to the relaxation of the compressive stress. Any surface reconstructions, however, have not been observed for the oxygen adsorption on Ni (100) and Ni (111) from gas phase at room temperature (Harrison et al. 2006; Ibach 1997).

By contrast, it has been reported that for the oxygen adsorption on Cu (100) from gas phase, the surface reconstruction of the oxygen overlayer from a c (2 × 2) to a (2×22)R45° structure takes place at room or elevated temperature (Harrison et al. 2006; Jensen et al. 1990; Kittel et al. 2001; Robinson et al. 1990). In the (2×22)R45° structure, every fourth [010] atomic row of the surface layer Cu atoms is missing and the nearest neighbor distance of the surface Cu atoms perpendicular to the missing rows (i.e. along the 22 direction) becomes large as compared with that in the unreconstructed phase (Harrison et al. 2006). As shown in Table 1, for the oxygen overlayers on Cu (100), both the experimental and theoretical values of Δg for the (2×22)R45° structure are less than those for the c (2 × 2) structure (Harrison et al. 2006), indicating that the missing row-surface reconstruction induces the surface stress variation toward tensile direction to relax the compressive stress, which is brought by the increase in the nearest neighbor distance of the surface layer metal atoms as well as the (2 × 2) p4g surface reconstruction for the carbon adsorption on Ni (100).

2.2 OH adsorption from aqueous solution

Figure 2 shows the voltammogram (A) and the volt-stressogram (g vs. E curve) (B) measured during a potentiodynamic polarization at a potential sweep rate of 10 mV/s in deaerated pH 8.4 borate buffer solution at 25 °C for the Cu thin film electrode with a thickness of 300 nm (prepared on a glass plate by magnetron sputtering) (Kim and Seo unpublished data). The value of g in the ordinate of Figure 2B is referred to zero at a cathodic limit potential of −0.8 V. The characteristic current peak positions in the voltammogram of Figure 2A are very close to those in the voltammogram of the single crystal Cu (111) electrode (Kunze et al. 2001, 2003a; Marcus and Maurice 2017; Maurice et al. 2000; Strehblow et al. 2001) measured in 0.1 M NaOH solution if the potential shift by 0.27 V due to the pH difference between two solutions is corrected. The in-situ STM study (Maurice et al. 2000) has revealed that OH adsorption, i.e. OH-underpotential deposition (UPD) on Cu (111) in 0.1 M NaOH solution proceeds, accompanying a surface reconstruction, in the potential region between −0.65 and −0.25 V, more negative than the oxide formation potential. On the other hand, Kunze et al. (2003b) have reported that OH-UPD on Cu (100) in 0.1 M NaOH solution proceeds in the potential region between −0.81 and −0.21 V.

Figure 2: 
						Voltammogram (A) and volt-stressogram (g vs. E curve) (B) measured during a potentiodynamic polarization at a potential sweep rate of 10 mV/s in deaerated pH 8.4 borate buffer solution at 25 °C for the Cu thin film electrode with a thickness of 300 nm (prepared on a glass plate by magnetron sputtering) (Kim and Seo unpublished data). The value of g in the ordinate of Figure 2B is referred to zero at a cathodic limit potential of −0.8 V.
Figure 2:

Voltammogram (A) and volt-stressogram (g vs. E curve) (B) measured during a potentiodynamic polarization at a potential sweep rate of 10 mV/s in deaerated pH 8.4 borate buffer solution at 25 °C for the Cu thin film electrode with a thickness of 300 nm (prepared on a glass plate by magnetron sputtering) (Kim and Seo unpublished data). The value of g in the ordinate of Figure 2B is referred to zero at a cathodic limit potential of −0.8 V.

The onset potential of the OH-UPD on Cu (100) is more negative by 0.16 V than that of the OH-UPD on Cu (111). In addition, the potential window (0.6 V) of the OH-UPD on Cu (100) is much wider than that (0.4 V) of the OH-UPD on Cu (111) (Kunze et al. 2003a,b; Maurice et al. 2000). As shown in the volt-stressogram of Figure 2B, g decreases toward compressive direction in the anodic potential sweep from −0.80 V, exhibiting a small pause at −0.35 V corresponding to the onset of the OH-UPD, and subsequently decreases rapidly up to 0.15 V. When the anodic potential sweep exceeds 0.15 V, g increases conversely toward tensile direction. The increase in g in the potential region more positive than 0.15 V corresponds to the growth of anodic oxide films on Cu, which will be discussed in the next section. The onset potential (−0.35 V) of the OH-UPD in Figure 2A almost coincides with that (−0.38 V) obtained by the correction of the pH difference for the results of Cu (111) in 0.1 M NaOH solution (Maurice et al. 2000). Furthermore, the potential window (0.5 V) of the OH-UPD in Figure 2A is wider that that (0.4 V) for the results of Cu (111) (Maurice et al. 2000), while it is narrower that that (0.6 V) for the results of Cu (100) (Kunze et al. 2003b).

Okolo et al. (2005) have reported that the magnetron-sputter-deposited Cu thin film (with a thickness of 500 nm) on Si3N4 substrate consists of a columnar microstructure with a more pronounced {111} fiber texture. The above similarity between Figure 2A and the results of Cu (111) (Maurice et al. 2000) suggests that the Cu thin film used for the experiments of Figure 2 is mainly oriented to {111} although no structural investigation was not made for the Cu thin film. The surface stress variation Δg = −0.5 J/m2 in the potential range between −0.35 and 0.15 V in Figure 2B means that the surface stress varies toward compressive direction due to the OH-UPD on Cu (111). The clear STM images of the OH-UPD structure on Cu (111) in 0.1 M NaOH solution (Kunze et al. 2001, 2003a; Marcus and Maurice 2017; Maurice et al. 2000; Strehblow et al. 2001) have indicated that the OH-UPD reconstructs the topmost Cu plane: the adsorbed OH groups occupy the three-fold hollow sites of the reconstructed Cu plane to form a p (2 × 2) superstructure which has a hexagonal lattice with a unit cell parameter of 0.6 ± 0.02 nm and mimics the Cu and O sublattices in Cu2O (111). The nearest neighbor distance between Cu atoms in the sub-lattice of this superstructure is 0.3 ± 0.01 nm, which is larger than that (0.256 nm) in the unreconstructed Cu (111) plane. The chemical bond of OH with surface metal atom induces the surface stress variation toward compressive direction if OH is electronegative with respect to the substrate metal. The increase in the nearest neighbor distance between Cu atoms due to the surface reconstruction, however, induces the surface stress variation toward tensile direction. As a result, the final direction of the surface stress variation due to the OH-UPD will be determined by the competition between the component (compressive direction) due to the chemical bond of OH with surface Cu atom and the component (tensile direction) due to the surface reconstruction. The discussion on magnitude of the above two components is valuable for better understanding of the surface reconstruction.

Figure 3 shows the atomic models of (A) the reconstructed p (2 × 2) superlattice structure of the OH-UPD layer on Cu (111) in 0.1 M NaOH solution and (B) the unreconstructed structure of the bulk Cu (111) plane (Kunze et al. 2003a). Let us estimate roughly the two components of the surface stress variation on basis of the atomic model in Figure 3A. The net surface stress variation Δgnet may be expressed by

(3)Δgnet=Δgad+Δgrec,

where Δgad is the component caused by the change in adsorption energy of OH with an elastic deformation of the substrate Cu (111) surface (Levia et al. 2000), and Δgrec is the component caused by an elastic stretching of the substrate surface due to the reconstruction. The sign of the first term Δgad in the right-hand side of Equation (3) should be negative because an elastic stretching of the substrate metal surface enhances an adsorption energy (Gsell et al. 1998; Jakob et al. 2001; Mavrikakis et al. 1998).

Figure 3: 
						Atomic models of (A) the reconstructed p (2 × 2) superlattice structure of the OH-UPD layer on Cu (111) in 0.1 M NaOH solution and of (B) the unreconstructed structure of the bulk Cu (111) plane (Kunze et al. 2003a). Copyright ©2003. Reproduced with permission from Elsevier. All rights reserved.
Figure 3:

Atomic models of (A) the reconstructed p (2 × 2) superlattice structure of the OH-UPD layer on Cu (111) in 0.1 M NaOH solution and of (B) the unreconstructed structure of the bulk Cu (111) plane (Kunze et al. 2003a). Copyright ©2003. Reproduced with permission from Elsevier. All rights reserved.

Assuming linear elasticity, the second term Δgrecin the right-hand side of Equation (3) is formulated as follows:

(4)Δgrec=Yεdmono=E1νεdmono,

where Y is the biaxial elastic modulus, E is Young’s modulus, ν is Poisson’s ratio, ε is the elastic strain, and dmono is the monolayer thickness (corresponding to the diameter of Cu atom, i.e. dmono = 0.256 nm) for the Cu (111) plane. The values of E = 130 GPa and ν = 0.5 (and thus, Y = 260 GPa) of the Cu (111) plane can be calculated by using the transformation relationships (Brantley 1973; Ibach 2006) between the compliance values (Brandes and Brook 1992) published for the bulk cubic crystal Cu.

The elastic strain ε of the Cu (111) plane is given by

(5)ε=an,ran,oan,o,

where an,r is the nearest neighbor distance between Cu atoms in the sub-lattice of the reconstructed phase and an,o is that in the unreconstructed Cu (111) plane. The tensile elastic strain ε = 0.172 is obtained from the values of an, r = 0.3 nm and an, o = 0.265 nm by using Equation (5). The substitution of Y = 260 GPa, ε = 0.172, and dmono = 0.256 nm into Equation (4) leads to an extraordinarily large value of Δgrec = 11.4 J/m2. In Equation (4), it is supposed that the biaxial elastic stretching of the Cu (111) plane due to the reconstruction proceeds without any ejection of Cu atoms from the lattice sites followed by migration to the step edges.

Atomically resolved STM images in the OH-UPD potential region for Cu (111) in 0.1 M NaOH solution (Kunze et al. 2001, 2003a; Marcus and Maurice 2017; Maurice et al. 2000; Strehblow et al. 2001) have revealed that Cu atoms are ejected during the reconstruction and they are accumulated at the step edges to form Cu ad-islands on the terraces. Such dynamic processes of Cu atoms would reduce the biaxial modulus or induce the downward deviation from the linear elasticity in Equation (4). Molecular dynamic (MD) simulations performed for a Cu (111) plane using embedded atom method potentials (Trimble et al. 2005) have indicated that the surface stress variation associated with the biaxial stretching of a Cu (111) monolayer deviates downward from a linearity when ε exceeds 0.01. The extrapolation of the simulated Δg versus ε curve (Trimble et al. 2005) to ε = 0.172 leads to a value of Δgrec = 3.0 J/m2, which is only one fourth of that (Δgrec = 11.4 J/m2) obtained from the linear elasticity in Equation (4).

Employing Δgnet = −0.5 J/m2 in Figure 2B and Δgrec = 3.0 J/m2, Δgad = −3.5 J/m2 is estimated from Equation (3). The positive value of Δgrec means that the surface stress variation toward compressive direction caused by the OH adsorption (i.e. OH-UPD) on the Cu (111) plane is relaxed by the surface reconstruction. The monotonous decrease in g from −0.8 to −0.35 V prior to the OH-UPD in Figure 2B may result from the surface charging in the electric double layer region of Cu (111) electrode (Umeno et al. 2007; Weissmüller 2013) although not excluding the influence of hydrogen evolution on Cu at potentials more negative than −0.5 V. The potential of zero charge Epzc of Cu (111) in pH 8.4 borate buffer solution would be −0.50 V as expected from Epzc = −0.24 V of Cu (111) in KClO4 solution of pH 4.0 (Lecoeur and Bellier 1985). A surface tension takes a maximum at Epzc, while a surface stress of a solid electrode does not always take a maximum within the measured potential range (Kramer 2008; Schmickler and Levia 1998).

According to in situ AFM study of Cu (100) in 0.1 M H2SO4 and 0.1 M HClO4 solutions (Cruickshank et al. 1993), chemisorbed oxygen or OH occupies the four-fold hollow site to form a c (2 × 2) superlattice, which is same as that for the oxygen adsorption on Cu (100) from gas phase. In contrast, in situ STM study of Cu (100) in 0.1 M NaOH solution (Kunze et al. 2003b) has shown that a surface reconstruction takes place in the OH-UPD potential region of Cu (100): Cu atoms are ejected from the Cu (100) plane to form dimers of superimposed Cu atoms which is stabilized by OH groups probably located in bridging positions, and finally c (2 × 6) and c (6 × 2) superstructure domains are formed as a result of zig-zag arrangement of the dimers. The adsorption energy of oxygen or OH groups on Cu (100) should be larger than that on Cu (111) (Kunze et al. 2003b), which may lead to a large negative value of the first term Δgad in the right-hand side of Equation (3). Although the mechanism of the surface reconstruction of Cu (100) is different from that that of Cu (111), the surface reconstruction of Cu (100) would play an important role in relaxing the surface stress variation toward compressive direction caused by adsorption of OH groups.

In situ STM study of Ni (111) in 10−3 M or 1 M NaOH solution (Marcus and Maurice 2017; Seyeux et al. 2005, 2006) has revealed the local formation of ordered p (2 × 2) islands of limited lateral extension (∼2 nm) on the Ni (111) terraces prior to OH adsorption-induced surface reconstruction followed by the formation of a two-dimensional passive layer at higher potential. The stability of the ordered p (2 × 2) islands on the Ni (111) terraces may be explained in terms of co-adsorption of mobile H2O and OHδ (0 < δ < 1) species because of weak adsorption of OHδ on Ni (111) plane as compared with OH adsorption on Cu (111) plane (Seyeux et al. 2006). The weak adsorption of OHδ on Ni (111) plane and the local formation of p (2 × 2) islands on the Ni (111) terraces may decrease both magnitudes of the first and second terms in the right-hand side of Equation (3). The surface stress variation of (111)-textured Ni thin film in the potential region of OH adsorption has not been measured yet. In future, the experimental and theoretical investigations of the relationships between surface stress variation and OH-induced surface reconstruction on single metal surfaces will be indispensable for better-understanding of the stability of OH adlayer as a two-dimensional precursor for a three-dimensional passive film.

3 Stress variation due to growth of anodic oxide films on metals

3.1 Anodic oxide films on Cu and Ni

The increase in g with the anodic potential sweep in the potential region more positive than 0.15 V in Figure 2B is associated with the growth of anodic oxide films on Cu. The broad anodic current peaks at 0.3–0.4 V and at 0.7–0.8 V in the voltammogram of Figure 2A are responsible for the formation of a duplex oxide film consisting of Cu2O as an inner layer and CuO or Cu (OH)2 as an outer layer, respectively. The volt-stressogram in Figure 2B deviates downward from the linearity at 0.5–0.6 V as shown by dotted lines, which results from the nucleation of CuO or Cu (OH)2 at the Cu2O/solution interface. The linearity in the volt-stressogram, however, is recovered in the potential region more positive than 0.6 V where the growth of inner and outer layers of the anodic oxide film proceeds simultaneously.

As shown in Figure 2B, the film growth in the anodic potential sweep from 0.2 to 1.2 V induces a stress variation of Δg = 1.7 J/m2. A stress variation Δg during film growth is given by

(6)Δg=(σfdf),

where σf (J/m3 or Pa) is the internal stress generated in the film and df is the film thickness. Although σf may depend ondf, Equation (6) is approximated by

(7)ΔgσfΔdf,

where 〈σf〉 is the average internal stress in the film (i.e. film stress) and Δdf is the increase in film thickness. The value of Δdf for the anodic oxide film formed on Cu in the anodic potential sweep from 0.2 to 1.2 V can be estimated from the electric charge densities corresponding to the areas of the cathodic current peaks Q1 and Q2 (see Figure 2A) appeared for the film reduction in the cathodic potential sweep from 1.2 to −0.8 V. At Q1, the outer layer CuO or Cu(OH)2 is reduced to Cu2O, while at Q2, the inner layer Cu2O in addition to Cu2O reduced in the outer layer is reduced to metallic Cu. The electric charge densities are Q1 = 10.0 C/m2 and Q2 = 22.4 C/m2, respectively. The increases in thickness of the outer layer Δdf,1 and of the inner layer Δdf,2 are given by

(8)Δdf,1=Q1M1Fρ1,

and

(9)Δdf,2=(Q2Q1)M22Fρ2,

where M1 or M2 is the molecular weight (g/mol) of the outer or inner layer, ρ1 or ρ2 is the density (g/cm3) of the outer or inner layer, and F is the Faraday constant (96,485 C/mol). Employing M1 = 79.55 or 97.56 g/mol and ρ1 = 6.31 or 3.37 g/cm3 for CuO or Cu(OH)2 (Kunze et al. 2001), respectively, Δdf,1 = 1.31 or 3.00 nm is estimated from Equation (8). Similarly, employing M2 = 143.1 g/mol and ρ2 = 6.04 g/cm3 for Cu2O (Kunze et al. 2001), Δdf,2 = 1.52 nm is estimated from Equation (9). If the outer layer is regarded as a mixture of CuO and Cu(OH)2 (Seo et al. 1988), Δdf,1 = 2.16 nm is pertinent as an average value. Substituting Δg = 1.7 J/m2 and Δdf = 3.68 nm into Equation (7), 〈σf〉 ≈ 460 MPa is approximately estimated as an average internal stress generated in the anodic oxide film on Cu.

Figure 4 shows the anodic current density i versus time t (A) and the stress g versus t (B) curves measured during anodic polarization at 0.8 V in deaerated pH 8.4 borate buffer solution at 25 °C for the electroplated Cu thin film electrode with a thickness of 800 nm (Seo and Hagioi 2007). As shown in Figure 4B, the stress evolves toward tensile direction and the stress variation during the film growth for 1 h is Δg = 1.8 J/m2. Xiang et al. (2002) have reported that electroplated Cu thin films have a pronounced {111} texture. After the anodic oxidation for 1 h at 0.8 V in Figure 4A, the cathodic reduction of the anodic oxide film performed at a constant current density of ic = 0.1 A/m2 provided Q1 = 9.92 C/m2 and Q2 = 20.8 C/m2 from which Δdf,1 = 2.13 nm as an outer layer consisting of CuO and Cu(OH)2 mixture and Δdf,2 = 1.34 nm as an inner Cu2O layer were derived by using Equations (8) and (9). As a result, 〈σf〉 ≈ 520 MPa is eventually obtained as an average internal stress generated in the anodic oxide film on Cu, which is the same orders of magnitude as that (〈σf〉 ≈ 460 MPa) estimated from Figure 2B.

Figure 4: 
						Anodic current density i versus time t (A) and stress g versus t (B) curves measured during anodic polarization at 0.8 V in deaerated pH 8.4 borate buffer solution at 25 °C for the electroplated Cu thin film electrode with a thickness of 800 nm (Seo and Hagioi 2007). Copyright ©2006. Reproduced with permission from Elsevier. All rights reserved.
Figure 4:

Anodic current density i versus time t (A) and stress g versus t (B) curves measured during anodic polarization at 0.8 V in deaerated pH 8.4 borate buffer solution at 25 °C for the electroplated Cu thin film electrode with a thickness of 800 nm (Seo and Hagioi 2007). Copyright ©2006. Reproduced with permission from Elsevier. All rights reserved.

It is known that metal thin films prepared by evaporation or sputtering are usually subjected to residual stresses (D’Heurle and Harper 1989). A residual stress σmof the substrate metal is released due to a decrease in substrate thickness (i.e. consumption of the substrate metal) caused by active dissolution of metal and growth of anodic oxide film during anodic polarization. If σm is tensile, a decrease in substrate thickness leads to an apparent decrease in 〈σf〉, while if σm is compressive, it leads to an apparent increase in 〈σf〉. It has been reported that the residual stresses of Cu thin films electroplated on a Si wafer are tensile: σm ≈ 70 MPa (Hwang et al. 2008) for a thickness of 1 μm and σm = 19.3 MPa (Zhou et al. 2004) for a thickness of 9.4 μm, while the residual stresses of Cu thin films sputtered onto an amorphous SiO2 are compressive: σm = −94 to −134 MPa (Okolo et al. 2005), depending on argon gas pressure for a thickness of 0.5 μm. The effect of σm on 〈σf〉 may be minor because the value of 〈σf〉 for the sputter-deposited Cu thin film is the same orders of magnitude as that for the electroplated Cu thin film, irrespective of the residual stress in the Cu substrate being tensile or compressive.

The growth of a barrier type of anodic oxide films on metals (Cabrera and Mott 1948–1949) is promoted by a high electric field with an order of 108 V/m across the film. The high electric field across the film exerts a stress component normal to the film plane due to coulombic attraction between charges of opposite sign located on both sides of the film. The stress component normal to the film plane is converted to the in-plane stress σfel parallel to the film plane, i.e. the electrostriction stress in the film since the film is mechanically constrained by the metal surface. Consequently, the electrostriction stress σfel is included in 〈σf〉 as far as the high electric field is sustained by the film. The intrinsic film stress σfi in the absence of the electric field can be obtained by subtracting σfel from 〈σf〉:

(10)σfi=σfσfel.

If dealing with an electrostatic force based on a simple-parallel capacitor model (Butler and Ginley 1988; Sahu et al. 1990), σfel is given by

(11)σfel=(νf1νf)(ϵ0ϵf2)(Δϕfdf)2,

where νf is Poisson’s ratio of the film, ϵ0 (8.854 × 10−12 F/m) is the vacuum permittivity, ϵf is the relative dielectric constant of the film, and Δϕf is the potential difference across the film. The term (Δϕfdf) in the right-hand side of Equation (11) corresponds to the electric field E perpendicular to the film plane. It is reminded that the film thickness df in Equation (11) is referred to zero at an onset-potential of film formation. It is assumed that E is uniform through the film. The minus sign of Equation (11) indicates that σfel is compressive in the film geometry due to coulometric attraction along the film thickness. Sato (1971) has proposed that a local breakdown of passive films (i.e. initiation of pitting corrosion) results from mechanical deformation induced by combination of an electrostriction stress with a decrease in surface tension of the film due to chloride adsorption on the film. In Equation (11), however, a dielectrostriction term is neglected. Vanhumbeeck and Proost (2007, 2008b) have pointed out that the large difference between σfel values calculated by using Equation (11) and obtained experimentally for the anodic oxide film on Ti is caused by neglect of a dielectrostriction term in Equation (11). Dielectrostriction is defined as changes of dielectric properties of a material with deformation (Peng et al. 2005). Deformation of the anodic oxide film due to an applied electric field affects the dielectric constant of the film since the diploes in the film are aligned along the direction of the applied electric field.

If the contribution of dielectrostriction is taken into consideration (Lee et al. 2005; Shkel and Klingenberg 1998), σfel can be expressed by

(12)σfel=(νf1νf)(ϵ02)[ϵf(α1+α2)](Δϕfdf)2,

where α1 and α2 are the electrostriction parameters due to changes in film thickness and its volume, respectively. Furthermore, in the case of an isotropic film, α1 and α2 can be expressed as a function of ϵf (Lee et al. 2005; Shkel and Klingenberg 1998):

(13)α1=25(ϵf1)2,

and

(14)α2=13(ϵf1)(ϵf+2)+215(ϵf1)2.

Substituting Equations (13) and (14) into Equation (12), σfel (Vanhumbeeck and Proost 2007, 2008b) is eventually represented by

(15)σfel=(νf1νf)(ϵ02)(0.6ϵf2+0.8ϵf0.4)(Δϕfdf)2.

The value of ϵf = 54 ± 4 calculated from Equation (15) by using the experimental values of σfel for the anodic oxide film (νf = 0.25) on Ti was statistically equal to that (55 ± 6) obtained independently by impedance measurements (Vanhumbeeck and Proost 2008b).

Let us calculate σfel by using Equation (15) in order to evaluate the contribution of σfel to 〈σf〉 for the anodic oxide film on Cu. For the calculation of σfel, νf = 0.454 (Hallberg and Hanson 1970) and ϵf = 7.11 (Switzer et al. 1998) of Cu2O are employed as a representative of the anodic oxide film on Cu since νf and ϵf of CuO or Cu(OH)2 are unknown. The electric field E=Δϕfdf is 2.85 × 108 V/m for the anodic oxide film on sputter-deposited Cu in Figure 2 and it is 1.87 × 108 V/m for the anodic oxide film on electroplated Cu in Figure 4. If E = 2.4 × 108 V/m is employed as an average value, σfel = −1.5 and −7.6 MPa are obtained, respectively, from Equations (11) and (15), indicating that the contribution of σfel to 〈σf〉 for the anodic oxide film on Cu is less than 2 % of 〈σf〉, even if the effect of dielectrostriction on σfel is taken into consideration. From the above discussion, it is drawn that the intrinsic film stress σfi of the anodic oxide film on Cu is around 500 MPa.

Figure 5 shows the anodic current density i versus time t (A) and the stress g versus t (B) curves measured during anodic oxidation at 0.8 V in deaerated pH 8.4 borate buffer solution at 25 °C for the Ni thin film electrode with a thickness of 300 nm (prepared on a glass plate by magnetron sputtering) (Kim and Seo 2003). As shown in Figure 5B, the stress evolves toward tensile direction and the stress variation during anodic oxidation for 1 h is Δg = 1.3 J/m2. The comparison between Figures 4A and 5A indicates that the passivation of Ni is very fast as compared with that of Cu in deaerated pH 8.4 borate buffer solution at 25 °C. According to the ellipsometric study of anodic oxide films on Ni in deaerated pH 8.4 borate buffer solution at 25 °C (Sato and Kudo 1974), the thickness of the anodic oxide film on Ni after 1 h-anodic oxidation at 0.8 V is about 1 nm. If Δdf = 1 nm is employed, 〈σf〉 ≈ 1.3 GPa for the anodic oxide film on Ni is estimated from Equation (7), which is larger by a factor of 2.6 than that for the anodic oxide film on Cu. The residual stress σmof the Ni thin film electrode was not measured in the experiments of Figure 5 (Kim and Seo 2003). Nevertheless, Windischmann (1987) has reported that the residual stress of the Ni thin film prepared with magnetron sputtering is σm ≈ −300 MPa.

Figure 5: 
						Anodic current density i versus time t (A) and stress g versus t (B) curves measured during anodic oxidation at 0.8 V in deaerated pH 8.4 borate buffer solution at 25 °C for the Ni thin film electrode with a thickness of 300 nm (prepared on a glass plate by magnetron sputtering) (Kim and Seo 2003). Copyright ©2003. Reproduced with permission from IOP Publishing Ltd. All rights reserved.
Figure 5:

Anodic current density i versus time t (A) and stress g versus t (B) curves measured during anodic oxidation at 0.8 V in deaerated pH 8.4 borate buffer solution at 25 °C for the Ni thin film electrode with a thickness of 300 nm (prepared on a glass plate by magnetron sputtering) (Kim and Seo 2003). Copyright ©2003. Reproduced with permission from IOP Publishing Ltd. All rights reserved.

By provisionally employing σm ≈ −300 MPa, let us estimate the contribution of σm to 〈σf〉 due to the decrease in thickness (consumption) of the substrate metal caused by the dissolution and film formation during anodic oxidation of Ni. The relationship between σm and 〈σf〉 is represented by

(16)σfΔdf=σf,cΔdf+σmΔdm,

where 〈σf,c〉 is the film stress after the correction of the residual stress and Δdm (<0) is the decrease in thickness of the substrate metal. In Equation (16), it is assumed that σm is uniform in depth-direction of the substrate metal. Furthermore, Δdm is expressed by

(17)Δdm=WmρmΔdfαPB,

where ∆Wm (g/cm2) is the dissolution amount of the substrate metal, ρm (g/cm3) is the density of the substrate metal and αPB is the Pilling–Bedworth ratio (Pilling and Bedworth 1923) that represents the relative volume change per the substrate metal atom due to oxidation of the metal. Furthermore, αPB is given by

(18)αPB=VfxVm=MfρmxMmρf,

where x is the stoichiometric number of the metal component in the film (MexOy), Vf is the molar volume of the film, Vm is the molar volume of the substrate metal, Mf is the molecular weight of the film, Mm is the atomic weight of the substrate metal and ρf is the density of the film.

The second term (ΔdfαPB) in the right-hand side of Equation (17) has been taken into consideration for the stress evolution during anodic oxidation of Ti thin film (Vanhumbeeck and Proost 2008). The first term, however, was not taken into consideration under the assumption of negligibly small dissolution of the substrate Ti (Vanhumbeeck and Proost 2008). In the case of passivation of Ni as well as Fe, the first term (Wmρm) in addition to the second term in the right-hand side of Equation (17) must be taken into consideration for the correction of the residual stress since Ni is subjected to active dissolution prior to passivation and subsequently to slight dissolution through the passive film. If ∆Wm = 3.2 × 10−7 g/cm2 and ρm = 8.845 g/cm3 are employed from the results of anodic oxidation of Ni at 0.8 V for 1 h in pH 8.4 borate buffer solution (Sato and Kudo 1974), the first term in Equation (17) is estimated to be Wmρm = −0.36 nm. In addition, αPB = 1.65 is obtained from Equation (18) by using ρf = 6.83 g/cm3 and ρm = 8.845 g/cm3 (Sato and Kudo 1974) for the anodic oxide film on Ni, Mf = 74.69 g for NiO (x = 1), and Mm = 58.69 g for Ni, respectively: the second term in Equation (17), therefore, is estimated be ΔdfαPB = −0.61 nm from αPB = 1.65 and Δdf = 1 nm. The summation of the first and second terms in Equation (17) is Δdm = −0.97 nm, which leads to σmΔdm ≈ 0.3 J/m2. As a result, 〈σf,c〉 ≈ 1.0 GPa for the anodic oxide film on Ni is derived from Equation (16), suggesting that the contribution of σm to 〈σf〉 is 23 % of 〈σf〉. Nevertheless, the more exact estimation of the contribution of σm to 〈σf〉 needs the preceding measurement of the residual stress of the metal thin film electrode.

Next, let us estimate the magnitude of the electrostriction stress σfel for the growth of the anodic oxide film on Ni in pH 8.4 borate buffer solution. The electric field across the anodic oxide film on Ni is E = 1.6 × 109 V/m (Sato and Kudo 1974), which is larger by a factor of about seven than that for the anodic oxide film on Cu, indicating that the anodic oxide film on Ni can sustain very high electric field as compared with that on Cu. In addition, νf = 0.32 (Robertson and Manning 1990) and ϵf = 13 (Fuschillo et al. 1974) are known for a NiO film. The substitution of the above parameters into Equations (11) and (15) leads to σfel = −60.4 and −520 MPa, respectively, indicating that the effect of dielectrostriction on σfel cannot be neglected for the growth of the anodic oxide film on Ni. It is noted that σfel = −520 MPa comes up to 50 % of 〈σf,c〉 ≈ 1.0 GPa. As a result, the intrinsic film stress of the anodic oxide film on Ni derived from the values of 〈σf,c〉 ≈ 1.0 GPa and σfel = −520 MPa is σfi ≈ 1.5 GPa, which is three times as much as that of the anodic oxide film on Cu.

3.2 Anodic oxide films on Fe

Figure 6 shows the successive potential steps in the passive potential region (A), the corresponding i versus t (B) and g versus t (C) curves in deaerated pH 8.4 borate buffer solution at 25 °C for the Fe thin film electrode with a thickness of 100 nm (evaporated on a glass plate) (Seo and Ueno 2020). The residual stress σm = 1.17 GPa of the evaporated Fe thin film was estimated from the linear relation between the anodic charge density and stress variation in the potential region between −0.6 and −0.3 V where the active dissolution of Fe proceeds (Seo and Ueno 2020) and it is close to that (1.35 GPa) of the Fe thin film (100 nm) evaporated on MgF2 (Thurner and Aberman 1990). In Figure 6A, after an air-formed oxide film on Fe was cathodically reduced at −0.6 V, the successive potential steps (0.2 V → 0.3 V → 0.5 V → 0.7 V → 0.9 V) with each 1 h-interval are applied to grow the anodic oxide film on Fe (Seo and Ueno 2020). The value of g in Figure 6C is referred to zero at −0.6 V. As shown in Figure 6C, at the first potential step (0.2 V), g decreases abruptly toward compressive direction (although not shown in Figure 6C) and then increases toward tensile direction with a decay of the anodic current to attain a steady state after 1 h.

Figure 6: 
						Successive potential steps in the passive potential region (A), the corresponding i versus t (B) and g versus t (C) curves in deaerated pH 8.4 borate buffer solution at 25 °C for the Fe thin film electrode with a thickness of 100 nm (evaporated on a glass plate) (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.
Figure 6:

Successive potential steps in the passive potential region (A), the corresponding i versus t (B) and g versus t (C) curves in deaerated pH 8.4 borate buffer solution at 25 °C for the Fe thin film electrode with a thickness of 100 nm (evaporated on a glass plate) (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.

The abrupt decrease in g results from the relaxation of the tensile residual stress of the substrate Fe due to the active dissolution (i.e. consumption) of Fe (Seo and Ueno 2020). On the other hand, the subsequent increase in g is associated with the anodic oxidation of ferrous ions once dissolved: the hydroxide layer mainly consisting of Fe(OH)3 or FeOOH⋅nH2O (1 ≥ n > 0) is deposited on Fe by the anodic oxidation and simultaneously dehydration of the deposited layer proceeds to induce the volume shrinkage, which leads to the stress variation toward tensile direction (Seo and Ueno 2020). In contrast, g begins to decrease gradually toward compressive direction at the second potential step (0.3 V) and it decreases stepwise from the third potential step (0.5 V). Figure 7 shows the steady state-values of g after 1 h at each potential step in Figure 6C and the thickness values (Sato et al. 1976) of the inner and outer layers measured with ellipsometry as a function of potential for the anodic oxide films formed on Fe for 1 h in deaerated pH 8.4 borate buffer solution at 25 °C.

Figure 7: 
						Steady state-values of g after 1 h at each potential step in Figure 6C and the thickness values (Sato et al. 1976) of the inner and outer layers measured with ellipsometry as a function of potential for the anodic oxide films formed on Fe for 1 h in deaerated pH 8.4 borate buffer solution at 25 °C (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.
Figure 7:

Steady state-values of g after 1 h at each potential step in Figure 6C and the thickness values (Sato et al. 1976) of the inner and outer layers measured with ellipsometry as a function of potential for the anodic oxide films formed on Fe for 1 h in deaerated pH 8.4 borate buffer solution at 25 °C (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.

Sato et al. (1976) have reported that the anodic oxide films formed on Fe at potentials above 0.1 V in deaerated pH 8.4 borate buffer solution at 25 °C consist of the inner barrier layer (γ-Fe2O3) and the outer deposit layer (γ-FeOOH). As shown in Figure 7, the thickness of the inner layer increases with increasing potential, while that of the outer layer keeps nearly constant, indicating that the stepwise decrease in g toward compressive direction caused by the successive potential steps from 0.3 V is associated with the growth of the inner layer. In Figure 7, however, g must be corrected for the residual stress relaxation due to the film growth and dissolution of metal ions through the film. For the estimation of the residual stress relaxation due to the film growth of the inner layer on Fe, it is necessary to know the relationship between Δdm and Δdf in Equation (17).

The ratio of Δdm to Δdf corresponding to reciprocal of αPB (see the second term of the right-hand side of Equation (17)) can be obtained from the linear potential dependencies (Figure 6 in ref. Sato et al. 1971) of the anodic charge density qf for the film formation and of the film thickness df in the passive potential region between 0.3 and 0.9 V for the anodic oxidation of Fe in deaerated pH 8.4 borate buffer solution at 25 °C:

(19)ΔdmΔdf=αPB1=Mm3FρmΔqfΔdf,

where Mm and ρm are the atomic weight and density of Fe, respectively. In Equation (19), it is assumed that the iron component in the anodic oxide film on Fe is Fe3+. The value of ΔdmΔdf = −0.395 (i.e. αPB = 2.53) (Seo and Ueno 2020) was obtained from the values of ΔqfΔdf = 1.61 × 10 C/m3 (Sato et al. 1971) and Mm3Fρm = 2.45 × 10−11 m3/C. As a result, the residual stress relaxation σmΔdm (the second term in the right-hand side of Equation (16)) of the substrate Fe for the film growth of Δdf (nm) is given by

(20)σmΔdm=0.395σmΔdf×109,

where σm = 1.17 GPa is the residual stress of the substrate Fe. Equation (20) indicates that the residual stress relaxation is −0.46 J/m3 per 1 nm-film growth.

Furthermore, the residual stress relaxation due to the dissolution of ferric ions through the anodic oxide film can be estimated from the increase in anodic charge density of dissolved ferric ions Δqd:

(21)σmΔdm=MmσmΔqd3Fρm.

Nevertheless, it is difficult to estimate accurately Δqd from the time-variation of anodic current density during the film growth at each potential step with 1 h-interval since the anodic charge is almost consumed for the film growth. The solution analysis for the Fe component dissolved into solution during anodic oxidation of Fe for 1 h in pH 8.4 borate buffer solution has indicated that no dissolution of Fe takes place in the potential range between 0 and 1.1 V within a detection limit of the solution analysis (Sato and Kudo 1971). The anodic current density after 1 h-anodic oxidation of Fe becomes less than 10−3 A/m2 (Seo and Ueno 2020). If an average dissolution current density of ferric ions is equal to 10−3 A/m2, Δqd = 3.6 C/m2 is estimated for 1 h at each potential step. The substitution of Δqd = 3.6 C/m2 into Equation (21) leads to the residual stress relaxation of −0.1 J/m2 for each potential step with 1 h-interval.

Figure 8 shows the values of g corrected for the residual stress relaxation of the substrate Fe due to the film growth of the inner layer and due to the film growth plus the dissolution which are plotted versus potential for comparison with the uncorrected values. In Figure 8, the corrections were made for g at potentials more positive than 0.3 V by referring to g at 0.3 V in Figure 7. Although the corrected values of g shift upward (toward tensile direction), the decreasing tendency (toward compressive direction) of g with increasing potential does not change. The changes in film stress Δσf due to the growth of the inner layer may be roughly estimated from Figure 8:

(22)Δσf=Δgdf=Δ(σfdf)df=dfΔσfdf,

where Δg′ and df are referred to g and df at 0.3 V, respectively, e.g. Δg′ = g (0.9 V) − g (0.3 V) and df = df (0.9 V) − df (0.3 V) are taken at 0.9 V.

Figure 8: 
						Values of g corrected for the residual stress relaxation of the substrate Fe due to the film growth of the inner layer (solid circles) and due to the film growth plus the dissolution (open triangles) which are plotted versus potential for comparison with the uncorrected values (open circles). The corrections were made for g at potentials more positive than 0.3 V by referring to g at 0.3 V in Figure 7 (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.
Figure 8:

Values of g corrected for the residual stress relaxation of the substrate Fe due to the film growth of the inner layer (solid circles) and due to the film growth plus the dissolution (open triangles) which are plotted versus potential for comparison with the uncorrected values (open circles). The corrections were made for g at potentials more positive than 0.3 V by referring to g at 0.3 V in Figure 7 (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.

Figure 9 shows the estimated values of Δσf as a function of potential, indicating that Δσf decreases toward compressive direction with increasing potential. The estimated values of Δσf at 0.9 V are −2.6 GPa for no correction, −2.2 GPa for the correction of the residual stress relaxation due to the film growth, and −1.9 GPa for that due to the film growth plus the dissolution. The magnitude of Δσf at 0.9 V for the correction of the residual stress relaxation due to the film growth or the film growth plus the dissolution is almost same as that of the film stress 〈σf〉 = −1.88 GPa for the anodic oxide (TiO2) film on evaporated Ti film electrode (Vanhumbeeck and Proost 2008b). It is reminded that Δσf differs from the average film stress 〈σf〉 for the whole thickness of the anodic oxide film on Fe. A tensile stress is generated for the formation of outer (deposit) layer to contribute to 〈σf〉 for the whole thickness of the anodic oxide film on Fe (Seo and Ueno 2020). Nevertheless, the contribution of the tensile stress in the outer layer to 〈σf〉 for the anodic oxide film formed on Fe at 0.9 V seems to be minor because the increase in thickness of the outer layer with potential is insignificant in the potential region between 0.3 and 0.9 V (Sato et al. 1976). The composite film stress of the anodic oxide film on Fe could be estimated by assuming a linear combination of the compressive stress in the inner layer and the tensile stress in the outer layer. At present, however, it seems difficult to estimate the composite film stress of the anodic oxide film on Fe from the stress data obtained experimentally so far (Seo and Ueno 2020).

Figure 9: 
						Changes in film stress Δσf estimated as a function of potential for no correction (open circles) and the corrections of the residual stress due to the film growth (solid circles) and due to the film growth plus the dissolution (open triangles) (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.
Figure 9:

Changes in film stress Δσf estimated as a function of potential for no correction (open circles) and the corrections of the residual stress due to the film growth (solid circles) and due to the film growth plus the dissolution (open triangles) (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.

In order to investigate the contribution of the electrostriction stress σfel to Δσf in the inner layer at 0.9 V, the cathodic polarization with a potential sweep rate of 1 mV/s was performed after 1 h-anodic oxidation at 0.9 V in deaerated pH 8.4 borate buffer solution at 25 °C for the Fe thin film electrode. Figure 10 shows the voltammogram (A) and the corresponding volt-stressogram (B) measured during the cathodic potential sweep for the Fe thin film electrode (Seo and Ueno 2020). In Figure 10B, a small maximum peak of g appears at about 0.5 V where the current density changes from anodic to cathodic although it is difficult to see clearly the maximum of g and the polarity change of the current density from Figures 10A and 10B. The part of Figure 10B in the potential range between 0.25 and 0.9 V is magnified to Figure 11 to confirm the maximum of g at about 0.5 V. The stress variation toward tensile direction due to the cathodic potential sweep from 0.9 to 0.5 V in Figure 11 is Δgel ≈ 0.42 J/m2, which is caused by the relaxation of the electrostriction stress in the inner layer since the electric field in the inner layer decreases with the cathodic potential sweep. Therefore, the electrostriction stress component σfel ≈ −350 MPa in the inner layer at 0.9 V is estimated from Δgel ≈ 0.42 J/m2 and df = df (0.9 V) − df (0.3 V) = 1.2 nm:

(23)σfel=Δgeldf.
Figure 10: 
						Voltammogram (A) and the corresponding volt-stressogram (B) measured during the cathodic polarization at a sweep rate of 1 mV/s from 0.9 V after 1h-anodic oxidation at 0.9 V in deaerated pH 8.4 borate solution at 25 °C for the Fe thin film electrode (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.
Figure 10:

Voltammogram (A) and the corresponding volt-stressogram (B) measured during the cathodic polarization at a sweep rate of 1 mV/s from 0.9 V after 1h-anodic oxidation at 0.9 V in deaerated pH 8.4 borate solution at 25 °C for the Fe thin film electrode (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.

Figure 11: 
						Magnification of Figure 10B in the potential range between 0.25 and 0.9 V (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.
Figure 11:

Magnification of Figure 10B in the potential range between 0.25 and 0.9 V (Seo and Ueno 2020). Copyright ©2020. Reproduced with permission from Springer Nature.

The value of σfel ≈ −350 MPa is 16 % of Δσf = −2.2 GPa for the correction of residual stress relaxation due to the film growth or 18 % of Δσf = −1.9 GPa for the correction of the residual stress relaxation due to the film growth plus the dissolution.

In order to investigate the effect of dielectrostriction on σfel in the inner layer, it is necessary to know the relative dielectric constant ϵf of the inner layer. The substitution of νf = 0.3 as an average value of iron oxides (Chicot et al. 2011), E = 5.7 × 108 V/m (Sato et al. 1976) and σfel ≈ −350 MPa into Equation (15) leads to ϵf ≈ 25, which is close to a mean value of ϵf = 12 (Stimming and Schultze 1979) and 40 (Azumi et al. 1987) reported for the inner layer (γ-Fe2O3). By contrast, the substitution of the above parameters into Equation (11) leads to ϵf ≈ 570, which is extraordinarily large and unrealistic, indicating that the effect of dielectrostriction on σfel in the inner layer must be taken into consideration. Consequently, the intrinsic film stress Δσfi in the growing inner layer at 0.9 V is estimated to be −1.85 GPa for the correction of the residual stress relaxation due to the film growth or −1.55 GPa for the correction of the residual stress due to the film growth plus the dissolution.

The flat band potential Efb of the anodic oxide film on Fe in deaerated pH 8.4 borate buffer solution at 25 °C is located around −0.1 V (Stimming and Schultze 1979), which means that the increase in g toward tensile stress due to the relaxation of the electrostriction stress in the cathodic potential sweep from 0.9 V should continue down to Efb at which no electric field is present in the inner layer. Nevertheless, g begins to decrease toward compressive direction at potentials below 0.5 V. The decrease in g may result from the alteration of the outer deposit layer such as hydration although the cause is not clear. The outer layer would be gradually hydrated by a penetration of H2O at potentials below 0.5 V to induce the volume expansion of the outer layer, which would lead to the decrease in g toward compressive direction (Seo and Ueno 2020). The above speculation may be supported by the ellipsometrical evidence (Azumi 1990) that the thickness of the anodic oxide film on Fe increases gradually after the anodic polarization was interrupted, and simultaneously, the refractive index of the film decreases gradually due to hydration. If the increase in g due to the relaxation of the electrostriction stress in the inner layer is masked with the decrease in g due to the volume expansion caused by hydration at potentials below 0.5 V, the real values of Δσfi in the growing inner layer may be less compressive as compared with the estimated values.

At present, the measurements of stress variations are limited to the anodic oxide films formed on Cu, Ni and Fe in deaerated pH 8.4 borate buffer solution at 25 °C. In future, it is necessary to measure the stress variations under various conditions (electrolyte composition, pH, and temperature) of anodic oxidation in order to accumulate the knowledge of the electro-chemo-mechanical properties of anodic oxide films.

3.3 Main factors controlling the sign of the intrinsic film stress

It seems that the sign of the intrinsic film stress in growing anodic oxide films on metals depends on semiconductive properties of the films. The anodic oxide films on Cu (Maurice et al. 1999; Strehblow et al. 2001) and Ni (Seyeux et al. 2006; Zuili et al. 2000) have p-type semiconductive properties. The major point defects in p-type semiconductive oxides are metal cation vacancies. In contrast, the anodic oxide films on Fe (Azumi et al. 1987; Stimming and Schultze 1979) have n-type semiconductive properties. The major point defects in n-type semiconductive oxides are oxygen vacancies and interstitial metal cations. Macdonald (1992, 1999, 2011 has proposed a point defect model (PDM) for the growth of passive films on metals. In the PDM model, it is assumed that the barrier (inner)layer of passive film is the stoichiometric oxide MOχ/2 and the precipitated (outer) layer is formed by hydrolysis of metal cations transmitted through the barrier layer and by reaction with anions in solution.

The metal cation vacancy VMχ with a negative charge is moved toward the metal/film interface by an applied electric field and reacts with the substrate metal atom m to form the metal cation MM in the cation site of the film and the vacancy Vm in the metal phase:

(24)m+VMχMM+Vm+χe.

Although VMχ is consumed at the metal/film interface, the dissolution of MM from the film into solution at the film/solution interface creates VMχ:

(25)MMMaqδ++VMχ+(δχ)e,

where Maqδ+ is the dissolved metal cation species. In addition, the interstitial metal cation Miχ+ is created at the metal/film interface:

(26)mMiχ++Vm+χe.

The created Miχ+ is moved toward the film/solution interface and dissolves as Maqδ+ into solution:

(27)Miχ+Maqδ++(δχ)e.

On the other hand, the oxygen vacancy VO with a positive charge is created at the metal/film interface:

(28)mMM+χ2VO+χe.

The created VO is moved toward the film/solution interface by an applied electric field and is consumed by reaction with H2O:

(29)VO+H2OOO+2H+,

where OO is the oxide ion in the anion site of the film. Furthermore, MOχ/2 reacts with H+ to dissolve as Maqδ+ into solution:

(30)MOχ/2+χH+Maqδ++χ2H2O+(δχ)e.

In p-type semiconductive oxides such as anodic oxide films on Cu and Ni, VMχ is continuously supplied to the metal/film interface since VMχ is created at the film/solution interface during film growth. Consequently, the reaction of Equation (24) proceeds to bring the increase in Vm as far as VMχ is supplied. If Vm does not sink sufficiently in the bulk metal, the increase in concentration of Vm leads to the formation of a void space (i.e. free space) in the metal side, by which the cantilever electrode is concaved to induce a tensile stress in the film (Macdonald 1992, 1999, 2011; Nelson and Oriani 1993). As a result, the intrinsic tensile stress in growing anodic oxide films on Cu and Ni may result from the void formation in the metal side. By the way, the PDM does not take the transport of metal cations via vacancy and interstitial sites, and of oxygen anions via vacancy sites into account (Seyeux et al. 2013). In p-type semiconductive oxides, the metal cations transport outward via cation vacancy sites and react with water at the film/solution interface to form a new oxide. The film/solution interface would not be mechanically constrained since the solution side is plastically deformed and thus the volume expansion due to the formation of the new oxide could not exert significantly the intrinsic film stress.

In n-type semiconductor oxides such as a barrier (inner) layer of the anodic oxide film on Fe, the oxygen anions transport inward via oxygen vacancy sites and react with metal atoms at the metal/film interface to form a new oxide in the metal side. The formation of the new oxide in the metal side leads to a volume expansion because of αPB > 1 (e.g. αPB = 2.53 for the inner layer (γ-Fe2O3) of the anodic oxide film on Fe). Since the metal/film interface is mechanically constrained, the bending of the cantilever electrode becomes convex due to the volume expansion, which induces a compressive stress in the film. On the other hand, interstitial metal cations Miχ+ in n-type semiconductor oxides transport outward via interstitial sites to react with water at the film/solution interface to form a new oxide. The formation of the new oxide at the film/solution interface could not exert significantly the intrinsic film stress since the film/solution interface would not be mechanically constrained. As a result, the compressive stress in the inner layer of the anodic oxide film on Fe may result mainly from the volume expansion due to the formation of the new oxide in the metal side at the metal/film interface (Seo and Ueno 2020).

3.4 Stress variations during cathodic reduction of anodic oxide films once formed on metals

As shown in Figures 2B and 10B, g changes toward compressive direction and then recovers toward tensile direction during cathodic reduction of anodic oxide films on Cu and Fe. In Figure 2B, two minima of g appear at about 0 V and −0.25 V corresponding to the cathodic current Q1 and Q2 peaks in Figure 2A, respectively. At the Q1 peak, the reduction of the outer (CuO/Cu(OH)2) layer to Cu2O proceeds and at the Q2 peak, the reduction of the converted Cu2O and inner (Cu2O) layers to metallic copper proceeds. The initial decrease in g results from the penetration of H+ into the film followed by its participation in the reduction of the film to produce H2O in the film as an intermediate stage (Seo and Hagioi 2007), which induces the volume expansion of the reducing film leading to the temporary decrease in g toward compressive direction. The subsequent desorption of H2O from the reducing film into solution brings the recovery of g toward tensile direction. The reduction reaction process of the outer layer of the anodic oxide film on Cu may be represented by

(31)2CuOx(OH)22x+2H++2e[Cu2O(1+x)H2O]*Cu2O+(1+x)H2O,

where the outer layer CuOx(OH)2−2x is partially hydrated (i.e. 1 > x > 0) (Seo et al. 1988) and [Cu2O(1+x)H2O]* means the intermediate species. In addition, the reduction reaction process of the produced Cu2O and inner Cu2O layers may be also represented by

(32)Cu2O+2H++2e[2CuH2O]*2Cu+H2O,

where [2CuH2O]* means the intermediate species. On the other hand, it seems that the reduction reaction process of the anodic oxide film on Fe is different from that of the anodic oxide film on Cu.

As shown in Figure 10A, the cathodic current peaks of the outer layer (γ-FeOOH) and inner layer (γ-Fe2O3) on Fe at −0.2 and −0.3 V are not completely separated, become doublet unlike those on Cu and the concomitant changes of g in Figure 10B exhibit one broad minimum. The one broad minimum of g observed during the cathodic reduction of the outer and inner layers on Fe may be explained in terms of the formation of Fe(OH)2 as an intermediate stage followed by its dissolution as Feaq2+ into solution or its deposition as metallic Fe (Seo and Ueno 2020). The formation of the intermediate Fe(OH)2 for the cathodic reduction of the outer and inner layers is represented by

(33)γFeOOH+H++e[Fe(OH)2]*

and

(34)γFe2O3+2H++H2O+2e2[Fe(OH)2]*,

where [Fe(OH)2]* means the intermediate species. The molar volume Vf = 2.64 × 10−5 m3/mol (West 1973–1974) for Fe(OH)2 is larger than Vf = 2.08 × 10−5 m3/mol (West 1973–1974) for α-FeOOH and Vf = 1.67 × 10−5 m3/mol (Seo and Ueno 2020) for γ-Fe2O3, indicating that the formation of [Fe(OH)2]* induces the volume expansion leading to the decrease in g toward compressive direction.

Furthermore, [Fe(OH)2]* reacts with H+ to dissolve as Feaq2+ into solution or deposit as metallic Fe (Seo and Ueno 2020):

(35)[Fe(OH)2]*+2H+Feaq2++2H2O

or

(36)[Fe(OH)2]*+2H++2eFe+2H2O.

The reaction of Equations (35) or (36) leads to the increase in g, i.e. g recovers toward tensile direction. Sato et al. (1971) have reported that the reductive dissolution as Feaq2+ does not proceed with a current efficiency of 100 % for the cathodic reduction of the anodic oxide film on Fe in pH 8.4 borate buffer solution after its formation in the same pH borate buffer solution, suggesting that the partial contribution of the reductive deposition as metallic Fe is not excluded in pH 8.4 borate buffer solution. In Figure 10, at potentials lower than −0.5 V where hydrogen evolution proceeds on Fe, g decreases rapidly toward compressive direction with increasing cathodic current density, which is attributed to the volume expansion in the surface region of the Fe thin film electrode due to the hydrogen adsorption or absorption (Seo and Ueno 2020).

In contrast to the anodic oxide films on Cu and Fe, the anodic oxide films on Ti (Ohtsuka et al. 1987) and Ni (MacDougall and Cohen 1974, 1976a,b) are not completely reduced in neutral solutions. The following compositional changes of the anodic oxide film (TiO2) on Ti during cathodic polarization in pH 6.9 phosphate solution have been proposed from the ellipsometrical results (Ohtsuka et al. 1987):

(37)TiO2+xH++xeTiO2x(OH)x,

where x varies from 0 to 1 with changing cathodic potential from −0.66 to −1.31 V. Concomitantly, the complex refractive index of the film, Nf = nki (n: refractive index, k: extinction index) changes from Nf = 2.1 − 0.03i to Nf = 1.85 − 0.35i (Ohtsuka et al. 1987). If the charge neutrality during hydrogen ingress in the film is taken into consideration, Equation (37) can be written by introducing the valence change of Ti ions (Kim et al. 2003):

(38)TiO2+xH++xe[xTi3+(1x)Ti4+]O2x(OH)x.

The stress variations (Kim et al. 2003) have been measured during cyclic potential steps between 0 and −1.0 V for the anodic oxide film formed on sputter-deposited Ti thin film electrode for 1 h at 5.2 V in pH 8.4 borate buffer solution: g decreases toward compressive direction at the downward potential step (0 V → −1.0 V), while it increases toward tensile direction at the upward potential step (−1.0 V → 0 V), and the changes in g during cyclic downward and upward potential steps are reversible, indicating that the reaction of Equation (37) or (38) proceeds reversibly in the potential region between 0 and −1.0 V.

The molar volume Vf = 29.42 cm3/mol of TiOOH (x = 1) (Michaelis et al. 1995) is larger than Vf = 20.8 cm3/mol of anatase type of TiO2 (x = 0) (Samsonov 1973). Consequently, the changes in g toward compressive direction at the downward potential step result from the volume expansion due to the formation of TiO2−x(OH)x or [xTi3+(1x)Ti4+]O2x(OH)x, and the reversible changes in g toward tensile direction at the upward potential step are caused by the volume contraction due to the regeneration of TiO2. The difference Δσfi = 320 MPa (Kim et al. 2003) in film stress between TiOOH and TiO2 was calculated from the linear relationship between Δg and anodic charge Qa (for hydrogen desorption) obtained during upward potential steps from various cathodic potentials (between −1.2 and −0.7 V) to 0 V after the formation of TiO2 film (thickness: 1.8 nm) on Ti for 1 h at 5.2 V in pH 8.4 borate buffer solution.

MacDougall and Cohen (1976a) have reported that the anodic oxide films formed on Ni in Na2SO4 solutions of pH 2.0−8.4 cannot be completely reduced at pH > 2.8. Moreover, the multiple-angle-of-incidence reflectivity measurements (Ohtsuka and Heusler 1979) have indicated that the film thickness increases initially, passing through a maximum and then decreases during cathodic reduction in 1 M H2SO4 solution after the formation of the anodic oxide film on Ni for 1 h in 1 M H2SO4 solution: concomitantly the complex refractive index of the film changes from Nf ≈ 2.1 − 0.5i to Nf ≈1.6 − 0.1i. The above changes of the thickness and complex refractive index of the film suggest that the anodic oxide film on Ni is cathodically reduced in 1 M H2SO4 solution via the formation of Ni(OH)2 as an intermediate stage followed by dissolution of Ni(OH)2 as Ni2+ into solution (Ohtsuka and Heusler 1979). The formation of Ni(OH)2 leads to the increase in thickness i.e. the volume expansion since the molar volume of Ni(OH)2 is larger than that of NiO (Ohtsuka and Heusler 1979).

The stress variations (Kim and Seo 2003) have been measured during the downward (from 0.8 or 0 V to −0.75 V) and upward potential steps (from −0.75 V to 0 or 0.8 V) after the formation of the anodic oxide film on sputter-deposited Ni thin film electrode at 0.8 V in pH 8.4 borate buffer solution: the downward and upward potential steps induce the changes in g toward compressive and tensile directions, respectively. The changes in g toward compressive direction during the downward potential step were mainly attributed to the volume expansion due to the formation of Ni(OH)2 (Kim and Seo 2003). The cathodic reduction of the anodic oxide film on Ni in pH 8.4 borate buffer solution may be represented by

(39)Ni1xO+2xH++2xe+(12x)H2O(1x)Ni(OH)2,

where the anodic oxide film on Ni is regarded as a p-type of semiconductive oxide with cation vacancies, i.e. Ni1−xO (1/3 >x > 0) and x depends on the film formation potential. In Equation (39), Ni(OH)2 is stable at −0.75 V where hydrogen evolution proceeds in pH 8.4 borate solution. The changes in g toward tensile direction during the upward potential step were mainly attributed to the reverse reaction of Equation (39), i.e. the volume contraction due to the formation of Ni1−xO. At −0.75 V, hydrogen atoms are absorbed into the substrate Ni through the produced Ni(OH)2, which induces a compressive stress. Nevertheless, its contribution to the changes in g toward compressive direction was about one fourth (Kim and Seo 2003).

4 Summary

The electro-chemo-mechanical (ECM) properties of the anodic oxide films on metals have been comprehensively discussed to provide a new insight into passivity:

  1. Oxygen adsorption on single crystal Cu or Ni from gas phase induces the changes in surface stress toward compressive direction, accompanying the formation of a superlattice structure of an oxygen overlayer as a two-dimensional precursor for a three-dimensional oxide film. The surface stress of the sputter-deposited (mainly (111)-textured) Cu thin film electrode in deaerated pH 8.4 borate buffer solution at 25 °C varies also toward compressive direction due to OH adsorption (i.e. OH-UPD) at potentials more negative than the oxide formation potential, accompanying the reconstruction of the Cu sublattice to form a p (2 × 2) oxygen overlayer. The compressive stress induced by the chemical bond between OH (or O) and Cu is relaxed by the tensile stress induced by the expansion of the Cu sublattice that caused by the reconstruction.

  2. The stress varies with growth of three-dimensional oxide films during anodic oxidation of Cu, Ni or Fe thin film electrode in deaerated pH 8.4 borate buffer solution at 25 °C. In the presence of a residual stress in the metal substrate, the residual stressis released due to the decrease in substrate thickness caused by the growth of the oxide film and the metal dissolution through the oxide film into solution. Furthermore, the dielectrostriction caused by a high electric field in the growing oxide film induces a compressive stress in the oxide film. The magnitude of the intrinsic film stress of the anodic oxide film itself can be determined by subtracting the effects of the residual stress and dielectrostriction from the measured stress variations.

  3. The sign (plus: tensile or minus: compressive) of the intrinsic film stress of the anodic oxide film depends on semiconductive properties of the film with point defects. The tensile intrinsic film stress of the anodic oxide film (with p-type semiconductive properties) on Cu and Ni is explained in terms of the void formation in the metal side due to the accumulation of the metal vacancies formed by the reaction of the substrate metal atoms with the cation vacancies at the metal/film interface. In contrast, the compressive intrinsic film stress of the inner barrier layer (with n-type semiconductive properties) of the anodic oxide film on Fe is explained in terms of the volume expansion due to the oxide formed in the metal side by the reaction of the substrate metal atoms with the oxygen anions at the metal/film interface.

  4. The stress variations observed during cathodic reduction of the anodic oxide films once formed on metals in deaerated pH 8.4 borate buffer solution at 25 °C are associated with the cathodic reduction mechanism. In the case of the anodic oxide films on Cu and Fe, the stress varies toward compressive direction due to the volume expansion caused by the production of H2O (for Cu) and the formation of Fe(OH)2 (for Fe) in reducing film as an intermediate stage. The subsequent stress recovery toward tensile direction is due to the desorption of H2O into solution followed by the complete reduction up to metallic Cu and due to the dissolution of Fe(OH)2 into solution as Fe2+ and further reduction up to metallic Fe. In the case of the anodic oxide (Ni1−xO) film on Ni and the anodic oxide film (TiO2) on Ti, the stress also varies toward compressive direction due to the volume expansion caused by the formation of intermediate species Ni(OH)2 and TiOOH. The stress, however, does not subsequently recover toward tensile direction because of no dissolution of Ni(OH)2 and TiOOH into solution and no further reduction up to metallic states.


Corresponding author: Masahiro Seo, Faculty of Engineering, Hokkaido University, Kita-13 Jo, Nishi-8 Chome, Kita-ku, Sapporo060-8628, Japan, E-mail:

About the author

Masahiro Seo

Masahiro Seo is an emeritus professor at Hokkaido University, Sapporo. He received his PhD degree in applied chemistry in 1971 from Hokkaido University. He joined as a research associate in 1972 at the Fontana Corrosion Center of the Ohio State University. Dr. Seo was a professor (from 1990 to 2005) at Hokkaido University. His research interests are focused on aqueous corrosion and passivation processes of metal thin films, measurement of stresses generated during anodic oxidation of transition metals, and inhibition effects of underpotential deposition on aqueous corrosion of metals. He is an ECS fellow and an honorary member of Japan Society of Corrosion Engineering.

  1. Research ethics: Not applicable.

  2. Author contributions: The author has accepted responsibility for the entire content of this submitted manuscript and approved submission.

  3. Competing interests: The author declares no conflicts of interest regarding this article.

  4. Research funding: None declared.

  5. Data availability: Not applicable.

References

Azumi, K. (1990). Study for the semiconductive property of the passive film on iron. Ph.D. thesis. Sapporo, Hokkaido University.Search in Google Scholar

Azumi, K., Ohtsuka, T., and Sato, N. (1987). Mott-Schottky plot of the passive film formed on iron in neutral borate and phosphate solutions. J. Electrochem. Soc. 134: 1352–1357, https://doi.org/10.1149/1.2100672.Search in Google Scholar

Bradhurst, D.H. and Leach, J.S.L. (1966). The mechanical properties of thin anodic films on aluminum. J. Electrochem. Soc. 113: 1245–1249, https://doi.org/10.1149/1.2423797.Search in Google Scholar

Brandes, E.A. and Brook, G.B. (1992). Smithells metals reference book, 7th ed. Butterworth-Heinemann, Oxford, pp. 15–16.Search in Google Scholar

Brantley, W.A. (1973). Calculated elastic constants for stress problems associated with semiconductor devices. J. Appl. Phys. 44: 534–535, https://doi.org/10.1063/1.1661935.Search in Google Scholar

Butler, M.A. and Ginley, D.S. (1988). New technique for measurements of electrode strain during electrochemical reactions. J. Electrochem. Soc. 135: 45–51, https://doi.org/10.1149/1.2095586.Search in Google Scholar

Cabrera, N. and Mott, N.F. (1948–1949). Theory of the oxidation of metals. Rep. Progr. Phys. 12: 163–184, https://doi.org/10.1088/0034-4885/12/1/308.Search in Google Scholar

Chicot, D., Mendoza, J., Zaoui, A., Louis, G., Lepingle, V., Roudet, F., and Lesage, J. (2011). Mechanical properties of magnetite (Fe3O4), hematite (α-Fe2O3) and geothite (α-FeOOH) by instrumented indentation and molecular dynamic analysis. Mater. Chem. Phys. 129: 862–870, https://doi.org/10.1016/j.matchemphys.2011.05.056.Search in Google Scholar

Cruickshank, B.J., Sneddon, D.D., and Gewirth, A.A. (1993). In situ observation of oxygen adsorption on a Cu (100) substrate using atomic force microscopy. Surf. Sci. Lett. 281: L308–L314, https://doi.org/10.1016/0167-2584(93)91171-j.Search in Google Scholar

D’Heurle, F.M. and Harper, J.M.E. (1989). Note on the origin of intrinsic stresses in films deposited via evaporation and sputtering. Thin Solid Films 171: 81–92, https://doi.org/10.1016/0040-6090(89)90035-7.Search in Google Scholar

Feibelman, P.J. (1997). First-principles calculations of stress induced by gas adsorption on Pt (111). Phys. Rev. B 56: 2175–2182, https://doi.org/10.1103/physrevb.56.2175.Search in Google Scholar

Fuschillo, N., Lalevic, B., and Leung, B. (1974). Dielectric properties of NiO and NiO (Li). Thin Solid Films 24: 181–192, https://doi.org/10.1016/0040-6090(74)90263-6.Search in Google Scholar

Gokhshtein, A.Y. (1976). Surface tension of solids and adsorption. Nauka, Moscow.Search in Google Scholar

Grossmann, A., Erley, W., and Ibach, H. (1995). Adsorbate-induced surface stress and surface reconstruction: oxygen, sulfur, and carbon on Ni (111). Surf. Sci. 337: 183–189, https://doi.org/10.1016/0039-6028(95)00615-x.Search in Google Scholar

Gsell, M., Jakob, P., and Menzel, D. (1998). Effect of substrate strain on adsorption. Science 280: 717–720, https://doi.org/10.1126/science.280.5364.717.Search in Google Scholar PubMed

Haiss, W., Nichols, R.J., Sass, J.K., and Charle, K.P. (1998). Linear correlation between surface stress and surface charge in anion adsorption on Au (111). J. Electroanal. Chem. 452: 199–202, https://doi.org/10.1016/s0022-0728(98)00137-5.Search in Google Scholar

Hallberg, J. and Hanson, R.C. (1970). The elastic constants of cuprous oxide. Phys. Status Solidi B 42: 305–310, https://doi.org/10.1002/pssb.19700420131.Search in Google Scholar

Harrison, M.J., Woodruff, D.P., Robinson, J., Sander, D., Pan, W., and Kirschner, J. (2006). Adsorbate-induced surface reconstruction and surface-stress changes in Cu (100)/O: experiment and theory. Phys. Rev. B74: 165402-1-7, https://doi.org/10.1103/physrevb.74.165402.Search in Google Scholar

Hong, S., Kara, A., Rahman, T.S., Heid, R., and Bohnen, K.P. (2004). Ab initio calculations of adsorbate-induced stress on Ni (100). Phys. Rev. B69: 195403-1-8, https://doi.org/10.1103/physrevb.69.195403.Search in Google Scholar

Hwang, S.-J., Joo, Y.-C., and Koike, J. (2008). Stress relaxation during isothermal annealing in electroplated Cu films. Thin Solid Films 516: 7588–7594, https://doi.org/10.1016/j.tsf.2008.03.030.Search in Google Scholar

Ibach, H. (1997). The role of surface stress in reconstruction, epitaxial growth, and stabilization of mesoscopic structures. Surf. Sci. Rep. 29: 193–263, https://doi.org/10.1016/s0167-5729(97)00010-1.Search in Google Scholar

Ibach, H. (2006). Physics of surfaces and interfaces. Springer-Verlag, Berlin.Search in Google Scholar

Jakob, P., Gsell, M., and Menzel, D. (2001). Interactions of adsorbates with locally strained substrate lattices. J. Chem. Phys. 114: 10075–10085, https://doi.org/10.1063/1.1369161.Search in Google Scholar

Jensen, F., Besenbacher, F., Laegsgaard, E., and Stensgaard, I. (1990). Dynamics of oxygen-induced reconstruction of Cu (100) studied by scanning tunneling microscopy. Phys. Rev. B42: 9206–9212, https://doi.org/10.1103/physrevb.42.9206.Search in Google Scholar PubMed

Kim, J.-D., Pyun, S.-I., and Seo, M. (2003). Effect of hydrogen on stresses in anodic oxide film on titanium. Electrochim. Acta 48: 1123–1130, https://doi.org/10.1016/s0013-4686(02)00823-x.Search in Google Scholar

Kim, J.-D. and Seo, M. (2003). Effect of applied potential steps on stresses of anodic oxide film on nickel. J. Electrochem. Soc. 150: B193–B198, https://doi.org/10.1149/1.1561630.Search in Google Scholar

Kittel, M., Polcik, M., Terborg, R., Hoeft, J.-T., Baumgärtel, P., Bradshow, A.M., Toomes, R.L., Kang, J.-H., Woodruff, D.P., Pascal, M., et al.. (2001). The structure of oxygen on Cu (100) at low and high coverages. Surf. Sci. 470: 311–324, https://doi.org/10.1016/s0039-6028(00)00873-6.Search in Google Scholar

Kramer, D. (2008). Dependence of surface stress, surface energy and surface tension on potential and charge. Phys. Chem, Chem. Phys. 10: 168–177, https://doi.org/10.1039/b710065e.Search in Google Scholar PubMed

Kunze, J., Maurice, V., Klein, L.H., Strehblow, H.-H., and Marcus, P. (2001). In situ tunneling microscopy of the anodic oxidation of Cu (111) in 0.1 M NaOH. J. Phys. Chem. B 105: 4263–4269, https://doi.org/10.1021/jp004012i.Search in Google Scholar

Kunze, J., Maurice, V., Klein, L.H., Strehblow, H.-H., and Marcus, P. (2003a). In situ STM study of the effect of chlorides on the initial stages of anodic oxidation on Cu (111) in alkaline solutions. Electrochim. Acta 48: 1157–1167, https://doi.org/10.1016/s0013-4686(02)00826-5.Search in Google Scholar

Kunze, J., Maurice, V., Klein, L.H., Strehblow, H.-H., and Marcus, P. (2003b). In situ STM study of the anodic oxidation of Cu (001) in 0.1 M NaOH. J. Electronal. Chem. 554–555: 113–125, https://doi.org/10.1016/s0022-0728(03)00115-3.Search in Google Scholar

Láng, G.G. and Barbero, C.A. (2012). Laser techniques for the study of electrode processes. Springer, Berlin.10.1007/978-3-642-27651-4Search in Google Scholar

Lecoeur, J. and Bellier, J.P. (1985). Potentials de charge nulle d’electrodes monocristallines de cuivre d’orientation (111) et (100) au contact de solutions aqueuses de perchlorate de potassium. Electrochim. Acta 30: 1027–1033, https://doi.org/10.1016/0013-4686(85)80168-7.Search in Google Scholar

Lee, H.Y., Peng, Y., and Shkel, Y.M. (2005). Strain-dielectric response of dielectrics as foundation for electrostriction stresses. J. Appl. Phys. 98: 074104-1-9, https://doi.org/10.1063/1.2073977.Search in Google Scholar

Levia, E.P.M., Del Pópolo, M.G., and Schmickler, W. (2000). Changes in surface stress caused by the adsorption of an epitaxial metal monolayer. Chem. Phys. Lett. 320: 393–397, https://doi.org/10.1016/s0009-2614(00)00240-2.Search in Google Scholar

Macdonald, D.D. (1992). The point defect model for the passive state. J. Electrochem. Soc. 139: 3434–3449, https://doi.org/10.1149/1.2069096.Search in Google Scholar

Macdonald, D.D. (1999). Passivity-the key to our metals-based civilization. Pure Appl. Chem. 71: 951–978, https://doi.org/10.1351/pac199971060951.Search in Google Scholar

Macdonald, D.D. (2011). The history of the point defect model for the passive state: a brief review of film growth aspects. Electrochim. Acta 56: 1761–1722, https://doi.org/10.1016/j.electacta.2010.11.005.Search in Google Scholar

MacDougall, B. and Cohen, M. (1974). Anodic oxidation of nickel in neutral solution. J. Electrochem. Soc. 121: 1152–1159, https://doi.org/10.1149/1.2402003.Search in Google Scholar

MacDougall, B. and Cohen, M. (1976a). Anodic oxide films on nickel in acid solution. J. Electrochem. Soc. 123: 191–197, https://doi.org/10.1149/1.2132784.Search in Google Scholar

MacDougall, B. and Cohen, M. (1976b). Mechanism of the anodic oxidation of nickel. J. Electrochem. Soc. 123: 1783–1789, https://doi.org/10.1149/1.2132697.Search in Google Scholar

Marcus, P. and Maurice, V. (2017). Atomic level characterization in corrosion studies. Phil. Trans. R. Soc. A 375: 20160414, https://doi.org/10.1098/rsta.2016.0414.Search in Google Scholar PubMed

Maurice, V., Strehblow, H.-H., and Marcus, P. (1999). In-situ scanning tunneling microscope study of the passivation of Cu (111). J. Electrochem. Soc. 146: 524–530, https://doi.org/10.1149/1.1391638.Search in Google Scholar

Maurice, V., Strehblow, H.-H., and Marcus, P. (2000). In situ STM study of the initial stages of oxidation of Cu (111) in aqueous solution. Surf. Sci. 458: 185–194, https://doi.org/10.1016/s0039-6028(00)00442-8.Search in Google Scholar

Mavrikakis, M., Hammer, B., and Nørskov, J.K. (1998). Effect of strain on the reactivity of metal surfaces. Phys. Rev. Lett. 81: 2819–2822, https://doi.org/10.1103/physrevlett.81.2819.Search in Google Scholar

Michaelis, A., Delplancke, J.L., and Schultze, J.W. (1995). Ellipsometric determination of density of TiO2 passive films on Ti single crystals: combination of ellipsometry and coulometry. Mater. Sci. Forum 185–188: 471–480, https://doi.org/10.4028/www.scientific.net/msf.185-188.471.Search in Google Scholar

Nelson, J.C. and Oriani, R.A. (1993). Stress generation during anodic oxidation of titanium and aluminum. Corros. Sci. 34: 307–326, https://doi.org/10.1016/0010-938x(93)90009-6.Search in Google Scholar

Ohtsuka, T. and Heusler, K.E. (1979). Potential-dependent properties of the passive film on nickel in sulfuric acid solution from multiple-angle-of-incidence reflectivity measurements. J. Electroanal. Chem. 102: 175–187, https://doi.org/10.1016/s0022-0728(79)80389-7.Search in Google Scholar

Ohtsuka, T., Masuda, M., and Sato, N. (1987). Cathodic reduction of anodic oxide films on titanium. J. Electrochem. Soc. 134: 2407–2410.10.1149/1.2100212Search in Google Scholar

Okolo, B., Lamparter, P., Welzel, U., Wagner, T., and Mittemeijer, E.J. (2005). The effect of deposition parameters and substrate surface condition on texture, morphology and stress in magnetron-sputter-deposited Cu thin films. Thin Solid Films 474: 50–63, https://doi.org/10.1016/j.tsf.2004.08.006.Search in Google Scholar

Olatunji-Ojo, O. and Taylor, C.D. (2013). Changes in valence, coordination and reactivity that occur upon oxidation of fresh metal surfaces. Phil. Mag. 93: 4286–4310, https://doi.org/10.1080/14786435.2013.827801.Search in Google Scholar

Peng, Y., Shkel, Y.M., and Kim, G.-H. (2005). Stress dielectric response in liquid polymers. J. Rheol. 49: 297–311, https://doi.org/10.1122/1.1835340.Search in Google Scholar

Pilling, N.B. and Bedworth, R.E. (1923). The oxidation of metals at high temperatures. J. Inst. Met. 29: 529–591.Search in Google Scholar

Robertson, J. and Manning, M.J. (1990). Limits to adherence of oxide scales. Mater. Sci. Technol. 6: 81–91, https://doi.org/10.1179/mst.1990.6.1.81.Search in Google Scholar

Robinson, I.K., Vlieg, E., and Ferrer, S. (1990). Oxygen-induced missing-row reconstruction of Cu (001) and Cu (001)-vicinal surfaces. Phys. Rev. B42: 6954–6964, https://doi.org/10.1103/physrevb.42.6954.Search in Google Scholar PubMed

Sahu, S.N., Scarminio, J., and Decker, F. (1990). A laser beam deflection system for measuring stress variations in thin film electrodes. J. Electrochem. Soc. 137: 1150–1154, https://doi.org/10.1149/1.2086618.Search in Google Scholar

Samsonov, G.V. (Ed.) (1973). The oxide handbook. IFI/Plenum, New York.10.1007/978-1-4615-9597-7Search in Google Scholar

Sato, N. (1971). A theory for breakdown of anodic oxide films on metals. Electrochim. Acta 16: 1683–1692, https://doi.org/10.1016/0013-4686(71)85079-x.Search in Google Scholar

Sato, N. and Kudo, K. (1971). Ellipsometry of the passivation film on iron in neutral solution. Electrochim. Acta 16: 447–462, https://doi.org/10.1016/0013-4686(71)85182-4.Search in Google Scholar

Sato, N. and Kudo, K. (1974). An ellipsometric study of anodic passivation of nickel in borate buffer solution. Electrochim. Acta 19: 461–470, https://doi.org/10.1016/0013-4686(74)87025-8.Search in Google Scholar

Sato, N., Kudo, K., and Nishimura, R. (1976). Depth analysis of passive films on iron in neutral borate solution. J. Electrochem. Soc. 123: 1419–1423, https://doi.org/10.1149/1.2132612.Search in Google Scholar

Sato, N., Kudo, K., and Noda, T. (1971). The anodic oxide film on iron in neutral solution. Electrochim. Acta 16: 1909–1921, https://doi.org/10.1016/0013-4686(71)85146-0.Search in Google Scholar

Schmickler, W. and Levia, E. (1998). A note on the surface stress and tension of solid electrodes. J. Electroanal. Chem. 453: 61–67, https://doi.org/10.1016/s0022-0728(98)00133-8.Search in Google Scholar

Schwennicke, C. and Pfnür, H. (1996). Structural analysis of a short range ordered layer with several adsorption sites: O/Ni (111). Surf. Sci. 369: 248–264, https://doi.org/10.1016/s0039-6028(96)00926-0.Search in Google Scholar

Seo, M. (2020). Electro-chemo-mechanical properties of solid electrode surfaces. Springer Nature, Singapore.10.1007/978-981-15-7277-7Search in Google Scholar

Seo, M. and Hagioi, M. (2007). Changes in mass and stress during anodic oxidation and cathodic reduction of the Cu/Cu2O multilayer film. Corros. Sci. 49: 176–185, https://doi.org/10.1016/j.corsci.2006.05.016.Search in Google Scholar

Seo, M., Jiang, X.C., and Sato, N. (1988). Photoacoustic study on cathodic reduction of anodic oxide films formed on copper in borate solution. Mater. Corros. (Werk. Korr.) 39: 583–588, https://doi.org/10.1002/maco.19880391205.Search in Google Scholar

Seo, M. and Ueno, K. (2020). Stress variations during polarization of iron thin film electrode in pH 8.4 borate buffer solution. J. Solid State Electrochem. 24: 929–940, https://doi.org/10.1007/s10008-020-04544-y.Search in Google Scholar

Seyeux, A., Maurice, V., Klein, L.H., and Marcus, P. (2005). In situ scanning tunneling microscopic study of the initial stages of growth and of the structure of the passive film on Ni (111) in 1 mM NaOH (aq). J. Solid State Electrochem. 9: 337–346, https://doi.org/10.1007/s10008-004-0627-5.Search in Google Scholar

Seyeux, A., Maurice, V., Klein, L.H., and Marcus, P. (2006). In situ STM study of the effect of chloride on passive film on nickel in alkaline solution. J. Electrochem. Soc. 153: B453–B463, https://doi.org/10.1149/1.2337768.Search in Google Scholar

Seyeux, A., Maurice, V., and Marcus, P. (2013). Oxide film growth kinetics on metals and alloys, 1. Physical model. J. Electrochem. Soc. 160: C189–C196, https://doi.org/10.1149/2.036306jes.Search in Google Scholar

Shkel, Y.M. and Klingenberg, D.J. (1998). Electrostriction of polarizable materials: comparison of models with experimental data. J. Appl. Phys. 83: 7834–7843, https://doi.org/10.1063/1.367958.Search in Google Scholar

Shuttleworth, R. (1950). The surface tension of solids. Proc. Phys. Soc. London A63: 444–457, https://doi.org/10.1088/0370-1298/63/5/302.Search in Google Scholar

Stimming, U. and Schultze, J.W. (1979). A semiconductor model of the passive layer on iron electrodes and its application to electrochemical reactions. Electrochim. Acta 24: 859–869, https://doi.org/10.1016/0013-4686(79)87009-7.Search in Google Scholar

Strehblow, H.-H., Maurice, V., and Marcus, P. (2001). Initial and later stages of anodic oxide formation on Cu, chemical aspects, structure and electronic properties. Electrochim. Acta 46: 3755–3766, https://doi.org/10.1016/s0013-4686(01)00657-0.Search in Google Scholar

Switzer, J.A., Hung, C.-H., Huang, L.-Y., Miller, F.S., Zhou, Y., Raub, E.R., Shumsky, M.G., and Bohannan, E.W. (1998). Potential oscillations during the electrochemical self-assembly of copper/cuprous oxide layered nanostructures. J. Mater. Res. 13: 909–916, https://doi.org/10.1557/jmr.1998.0124.Search in Google Scholar

Thurner, G. and Aberman, R. (1990). Internal stresses and structure of ultrahigh vacuum evaporated chromium and iron thin films and their dependence on substrate temperature and oxygen partial pressure during deposition. Thin Solid Films 192: 277–285, https://doi.org/10.1016/0040-6090(90)90072-l.Search in Google Scholar

Trimble, T., Tang, L., Vasiljevic, N., Dimitrov, N., van Schilfgaarde, M., Friesen, C., Thompson, C.V., Seel, S.C., Floro, J.A., and Sieradzki, K. (2005). Anion adsorption induced reversal of coherency strain. Phys. Rev. Lett. 95: 166106-1-4, https://doi.org/10.1103/physrevlett.95.166106.Search in Google Scholar PubMed

Ueno, K., Pyun, S.-I., and Seo, M. (2000). Stresses of a titanium thin-film electrode generated during anodic oxidation by a beam-bending method. J. Electrochem. Soc. 147: 4519–4523, https://doi.org/10.1149/1.1394095.Search in Google Scholar

Umeno, Y., Elsässer, C., Gumbsch, B., Meyer, P., Nothacker, M., Weissmüller, J., and Evers, F. (2007). Ab initio study of surface stress response to charging. Europhys. Lett. 78: 13001-1-5, https://doi.org/10.1209/0295-5075/78/13001.Search in Google Scholar

Vanhumbeeck, J.-F. and Proost, J. (2007). In-situ monitoring of the dielectric and electrostrictive properties of anodised thin films for biochip applications. Colloid Surf., Biointerf. 56: 163–169, https://doi.org/10.1016/j.colsurfb.2006.10.021.Search in Google Scholar PubMed

Vanhumbeeck, J.-F. and Proost, J. (2008a). On the relation between growth instabilities and internal stress evolution during galvanostatic Ti thin film anodization. J. Electrochem. Soc. 155: C506–C514, https://doi.org/10.1149/1.2958306.Search in Google Scholar

Vanhumbeeck, J.-F. and Proost, J. (2008b). On contribution of electrostriction to charge-induced stress in anodic oxide films. Electrochim. Acta 53: 6165–6172, https://doi.org/10.1016/j.electacta.2007.11.028.Search in Google Scholar

Van Overmeere, Q. and Proost, J. (2010). Stress-induced breakdown during galvanostatic anodising of zirconium. Electrochim. Acta 55: 4653–4660, https://doi.org/10.1016/j.electacta.2010.03.042.Search in Google Scholar

Weissmüller, J. (2013). Electrocapillarity of solids and its impact on heterogeneous catalysis. In: Alkire, R.C., Kolb, D.M., Kibler, L.A., and Lipkoeski, J. (Eds.), Advances in electrochemical science and engineering, Vol. 14. Wiley-VCH, Weinheim.10.1002/9783527680436.ch5Search in Google Scholar

West, R.C. (Ed.) (1973–1974). Handbook of chemistry and physics, 54th ed. CRC Press, Cleveland, pp. B98.Search in Google Scholar

Windischmann, H. (1987). An intrinsic stress scaling law for polycrystalline thin films prepared by ion beam sputtering. J. Appl. Phys. 62: 1800–1807, https://doi.org/10.1063/1.339560.Search in Google Scholar

Xiang, Y., Chen, X., and Vlassak, J.J. (2002). The mechanical properties of electroplated Cu thin films measured by means of the bulge test technique. Mat. Res. Soc. Symp. Proc. 65: L4.9.1–L4.9.6.Search in Google Scholar

Zhou, Y., Yang, C.-S., Chen, J.-A., Ding, G.-F., Ding, L., Wang, L., Wang, M.-J., Zhang, Y.-M., and Zhang, T.-H. (2004). Measurement of Young’s modulus and residual stress of copper film electroplated on Si wafer. Thin Solid Films 460: 175–180, https://doi.org/10.1016/j.tsf.2004.01.088.Search in Google Scholar

Zuili, D., Maurice, V., and Marcus, P. (2000). Surface structure of nickel in acid solution studied by in situ scanning tunneling microscopy. J. Electrochem. Soc. 147: 1393–1400, https://doi.org/10.1149/1.1393367.Search in Google Scholar

Received: 2023-05-27
Accepted: 2023-09-08
Published Online: 2023-12-04
Published in Print: 2024-02-26

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 6.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2023-0073/html
Scroll to top button