0% found this document useful (0 votes)
164 views27 pages

Role of Zirconium Conversion Coating in Corrosion Performance of Aluminum Alloys

A variety of chromate-free conversion coatings are being actively investigated to improve the corrosion performance of light-weight alloys for aerospace and defense applications. Advancing conversion coating, however, requires an in-depth understanding of the underlying corrosion mechanisms in order to rationally design sustainable coatings. Here, we present a multiscale modeling approach to predict corrosion performance of metallic materials, with a focus on localized corrosion of Cu.

Uploaded by

bmalki68
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
164 views27 pages

Role of Zirconium Conversion Coating in Corrosion Performance of Aluminum Alloys

A variety of chromate-free conversion coatings are being actively investigated to improve the corrosion performance of light-weight alloys for aerospace and defense applications. Advancing conversion coating, however, requires an in-depth understanding of the underlying corrosion mechanisms in order to rationally design sustainable coatings. Here, we present a multiscale modeling approach to predict corrosion performance of metallic materials, with a focus on localized corrosion of Cu.

Uploaded by

bmalki68
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 27

Role of Zirconium Conversion Coating in Corrosion Performance of Aluminum Alloys:

An Integrated First-Principles and Multiphysics Modeling Approach


Arash Samaei a, Santanu Chaudhuri a, b*

a
Civil, Materials, and Environmental Engineering Department, University of Illinois at Chicago, IL, 60607, United States

b
Argonne National Laboratory, Lemont, IL, 60439, United States

*
Corresponding author’s email address: [email protected]

Abstract

A variety of chromate-free conversion coatings are being actively investigated to improve the corrosion
performance of light-weight alloys for aerospace and defense applications. Advancing conversion coating,
however, requires an in-depth understanding of the underlying corrosion mechanisms in order to rationally
design sustainable coatings. Here, we present a multiscale modeling approach to predict corrosion performance
of metallic materials, with a focus on localized corrosion of Cu-containing aluminum alloys coated with ZrO2
layer. First-principles and transition-state theory are used to implement the kinetics model, which includes
electrolyte-metal interfacial reactions. The modeling framework systematically characterizes and couples
multiple electrochemical and physical (e.g., transport) phenomena to explore interrelationships between pit
morphology, surface chemistry, and local environment. This multiscale model can quantitatively link the
corrosion rate of ZrO2-coated aluminum alloys with the evolution of interfacial reactions during immersion,
which is very difficult to establish using in situ experiments. We have evaluated the presented multiscale model
using available experimental data. The rate of corrosion and pit stability were quantitatively assessed for various
environmental parameters and applied potentials. Results show that Zr-based conversion coating strongly
enhances the corrosion performance of aluminum alloys due to zirconium involvement in interfacial kinetics.

Keywords: Multiscale Simulation; Density Functional Theory; Transition State Theory; Finite Element
Method; Multiphysics; Corrosion Conversion; Aluminum Alloys; Zirconia; Localized Pitting Corrosion

1. Introduction

Light-weight aluminum alloys with excellent mechanical properties have been extensively used in defense and
aerospace applications. Yet, these alloys are prone to localized pitting corrosion, which often takes place on the
surface of passivated metals with a native oxide layer, due to the presence of heterogeneous microstructures [1,
2]. When aggressive anions (e.g., 𝐶𝑙 − ) attack aluminum alloys, severe localized pits with sizes ranging from
nano- to millimeters can form, which are difficult to diagnose and frequently initiate cracks that eventually lead
to catastrophic failures [3]. The complexity of this issue is primarily due to two factors: the microstructure of
the alloy and the corrosion environment, both of which vary dynamically with time and exhibit significant
heterogeneity [4].

Aluminum alloys possess complex microstructures due to the presence of intermetallic particles (IMPs) with a
broad range of compositions and sizes, as well as periphery phases around composite particles and clustering
[5]. The IMPs, which have a variety of compositions and phases, are believed to be critical sites for the
initiation of localized corrosion due to particle dealloying, etching of the alloy matrix, and Cu redistribution on
the surface. The latter phenomena drastically alter the local electrochemical activity on the surface of aluminum

1
alloys, hastening the progression of localized pitting corrosion [6, 7]. Extensive efforts have been made to
characterize the electrochemical and dissolution behaviors, such as pitting potential, corrosion potential, and
polarization curves, of all forms of IMPs (e.g., Al2Cu and Al2CuMg) within aluminum alloys under various
corrosion environments [8-14]. These studies demonstrated that the local corrosion environment governs the
main phenomena at the alloy-solution interface, such as IMP dealloying, aluminum oxide formation and
dissolution, matrix etching, and Cu redistribution. Moreover, pit morphology, electrical potential, electrolyte
chemistry, and microstructure of aluminum alloys, all of which are space-time variant features, exhibit coupled
effects on localized pitting corrosion. However, differentiating the effects of each individual feature on pitting
corrosion through experiments is very challenging. In particular, the distinct role of galvanic coupling and
electrolyte chemistry in the initiation and propagation of localized pits around copper-rich IMPs still need
investigations [15].

In order to improve the corrosion performance of aluminum alloys, numerous surface modification techniques,
such as organic (or polymeric), inorganic (e.g., anodization, conversion coating, physical or chemical vapor
deposition, and electroplating), and hybrid organic-inorganic coatings (e.g., sol-gel coating), have been broadly
investigated [16]. Conversion coating is used as a stronger pretreatment coating in a variety of coating
applications. Chromate conversion coatings (CCCs) are the most important coatings used in a wide range of
applications in the aircraft and defense industries due to their exceptional corrosion performance and great
adhesion properties [17]. CCCs are chromium compounds in the 6+ oxidation state that provide self-healing
corrosion inhibition due to the reduction of Cr(VI) within the coating to insoluble Cr(III) compounds [18]. The
ability of such coatings to reform a Cr-containing protective barrier after the CCC layer is broken during the
chemical process is referred to as self-healing.

Yet, recent studies have shown that hexavalent chromium poses significant environmental and health hazards
[19, 20]. CCCs threat workers’ life due to the high degree of toxicity and carcinogenicity of chromate
compounds, thus their usage is largely prohibited within the United States (by the Occupational Safety and
Health Administration) and the European Union (by the Restriction of Hazardous Substances Directive). These
issues have motivated researchers to develop chromate-free conversion coatings with equal or better corrosion
performance over the last two decades [21-24]. Several chromate-free conversion coatings, including rare-earth
metal (REM), Ti, and Zr-based coatings, have been introduced to improve the corrosion performance of
aluminum alloys. Among these alternatives, Zr-based conversion coating (ZrCC) has the advantages of having a
low environmental impact while also being abundant and economically viable [25]. Extensive experimental
efforts have been devoted to improving corrosion performance of ZrCC for aluminum alloys [22, 24, 26-30].
However, development of an effective ZrCC for long-term corrosion prevention requires a thorough
understanding of the underlying interfacial kinetics and surface chemistry, as well as optimization of coating
formulations and compositions. Since characterizing the above-mentioned space-time variant features is
experimentally very challenging, multiscale modeling and simulations offer chemo-physics-based tools for
investigating the controlling factors for corrosion prevention.

To date, various modeling approaches, including finite element method (FEM) and first-principles, have been
reported to investigate corrosion of metallic materials such as steel (or iron), magnesium, and aluminum alloys
[3, 4, 15, 31-47]. Aiming at energetic calculation and phase stability evaluation, ab initio was used to investigate
surface phenomena such as dissolution, adsorption, and oxygen reduction reaction with metal/alloy at the
nanoscale [48, 49]. Yet, such simulations are limited to very small systems that cannot properly describe the
localized pitting process and dynamics related to those phenomena, all of which may take months or years. FE
models, on the other hand, have been widely used for corrosion studies of different materials due to the ability

2
to simulate localized pits with real-world dimensions and environmental conditions. However, according to
Xiao and Chaudhuri [4], most of these models were developed based on simple and fixed pit geometry and
simplified solution chemistry, with critical assumptions such as zero concentration gradient in electrolyte and
time-invariant interfacial kinetics. Furthermore, they used simple metal compositions, which are inapplicable to
aluminum alloys with heterogeneous microstructures and compositions [4, 5]. Current FE models rely heavily
on experimental-based corrosion mechanism parametrization, and thus the availability of sufficient and accurate
experimental data severely limits the predictability of such models.

Incorporating first-principles into a multiscale modeling framework is believed to significantly improve


predictive corrosion models [4, 50]. Yet, to the best of our knowledge, no multiscale modeling has been
conducted on the localized pitting corrosion of light-weight alloys coated with conversion coatings. This work
develops a multiscale model that fully describes the coupled effects of electrolyte chemistry, interfacial kinetics,
alloy microstructure, and pit morphology on the corrosion rate and stability of pits. Using density functional
theory (DFT) and FEM, this model explores the interrelationships between these features to promote a
fundamental understanding of localized pitting corrosion in aluminum alloys with Zr-based conversion
coatings. The proposed kinetic model for electrolyte and surface chemistry was completed using DFT and
transition-state theory. This research focuses on the propagation and repassivation of a localized pit in vicinity
of a cathodic IMP. This work contributes to reducing the number of unknown parameters for underlying
reaction mechanisms and hastening the development of more comprehensive models for corrosion resistance of
conversion coatings. Section 2 will present a generic multiscale modeling framework with details of the DFT
calculations and multiphysics model, Section 3 will discuss the results, and Section 4 will summarize the
implications of this research.

2. Multiscale Modeling of Localized Corrosion

Figure 1 shows the system investigated in this work, which involves the exposure of an aluminum alloy to a
sodium chloride (NaCl) solution. The alloy microstructure was obtained from SEM images to serve as model
geometry for simulations that mimics a real system. The model considers a system comprised of a pure
aluminum matrix and an IMP region, i.e., -phase (Al2Cu). Two layers of metal oxides protect the Al matrix: a
thin inner layer of native aluminum oxide (Al2O3) and a thick deposited layer of zirconium oxide (ZrO2). This
study focuses on pit propagation and repassivation near a cathodic IMP, with repassivation triggered by the
formation of a passive layer of aluminum oxide and other insoluble precipitates on the surface [11, 12]. The
model considers four main domains, as shown in Figure 1(a): aluminum matrix Ω𝐴𝑙 as anode, -phase (Al2Cu)
Ω𝜃 as cathode, passive oxide layer (i.e., including both Al2O3 and ZrO2 layers) Ω𝑃 , and electrolyte (NaCl
solution) Ω𝑒 . The electrolyte-metal interface is defined as ∂Ω𝑒−𝑚 = ∂Ω𝑒 ∩ (∂Ω𝐴𝑙 ∪ ∂Ω𝜃 ∪ ∂Ω𝑃 ).

Figure 1(b) presents a generic multiscale modeling framework for investigating localized corrosion in light-
weight alloys. It includes three layers of modeling blocks that are distinguished by their dependencies and
relationships with one another and are linked to an experimental data block. The modeling process starts with
inputs from experimental data such as pit surface geometry and corrosion environment. In the first modeling
layer, first-principles and transition state theory are used to calculate required thermodynamic parameters,
including internal energy, enthalpy, and Gibbs free energy of activation for electrochemical reactions. The
Eyring-Polanyi relation, which will be explained in the following section, aids in obtaining rate constants for
dissolution reactions in order to complete the kinetics model. The reaction kinetics serves as an input in the
second layer, where the mass transport model is coupled with electrochemical reactions to characterize the
chemical (i.e., material concentrations) and electrical (i.e., electrostatic potential) environments. In the third

3
layer, the corrosive environment is used to determine both reactive species fluxes through the electrolyte-metal
interface and changes in interface location. Subsequently, the pit boundary location and condition for the
specified electrolyte model are obtained.

(a) (b)

Figure 1: (a) Schematic showing the geometry used for corrosion models. A pit is located adjacent to the cathodic region, which is composed of
Al2Cu IMPs (θ-phase). A layer of Zr-based conversion coating covers the aluminum matrix, beneath which is a very thin layer of native alumina. The
coating layer has a thickness of 1 𝜇𝑚. The models incorporate real-world pit geometry. The electrolyte is a sodium chloride (NaCl) solution. (b) The
scheme for multiscale modeling of localized corrosion of alloys/metals. First-principles in conjunction with transition-state theory are used for
energetic calculations to eventually determine the values of reaction parameters (e.g., barrier energy and rate constants). Coupling the proposed
kinetics with a mass transport model, the corrosive environment over the alloy is determined. Finally, the growth of pit and its stability are
quantitatively predicted.

2.1 First-Principles Calculations

Among the various mechanisms for metal hydrolysis reaction in aqueous solution reported in the literature, this
work used the reaction pathway for first-order hydrolysis introduced by Dong et al. [51]. The solvation model
density (SMD) – a continuum model – was used to simulate the reaction pathways [52]. The frequency
calculations and structure optimizations of reactants, transition states (TS), and products were performed using
DFT at the B3LYP/6-311+G(d,p) level [53]. By optimizing the structures achieved through calculations of
intrinsic reaction coordinate, the stable configurations of the reactants and products without imaginary
frequencies were obtained. However, the structures of TS with an imaginary frequency were determined using
Berny optimizations [54]. All DFT calculations were carried out at 0 K. Supplementary information includes
the Cartesian coordinates of all reaction species for reactants and transition states. The electronic energies
𝐸𝐸𝑙𝑒𝑐(0𝐾) for aqueous reaction species were computed using single-point calculations using the DFT MP2/6-
311+G(d,p) method. Subsequently, the thermodynamic variables total energy 𝐸298 , Gibbs free energy 𝐺298 , and
enthalpy 𝐻298 for those species were obtained by adding enthalpy correction 𝐻𝐶𝑜𝑟𝑟 , thermal correction 𝐸298 ,
entropy correction −𝑇𝑆298, and zero-point energy 𝐸298(0𝐾) onto 𝐸𝐸𝑙𝑒𝑐(0𝐾) [55]. The zero-point energy can be
obtained from frequency calculations. 𝐻𝐶𝑜𝑟𝑟 , 𝐸298 , and −𝑇𝑆298 are determined using thermochemistry analyses
carried out at 1 atm and 298.15 K. The Eyring equation yields the transition-state rate constant:

4
𝑘𝐵 𝑇 ∆𝐺298,𝑎𝑐𝑡 𝑘𝐵 𝑇 ∆𝑆298,𝑎𝑐𝑡 ∆𝐻298,𝑎𝑐𝑡
𝑘 𝑇𝑆𝑇 = ∙ exp (− )= ∙ exp (− − ) (1)
ℎ 𝑅𝑇 ℎ 𝑅𝑇 𝑅𝑇

where 𝑘𝐵 , ℎ, 𝑇, 𝑅, ∆𝐺298,𝑎𝑐𝑡 , ∆𝑆298,𝑎𝑐𝑡 , and ∆𝐻298,𝑎𝑐𝑡 denote the Boltzmann’s constant, Planck’s constant,
temperature, gas constant, activation Gibbs free energy at 298 K, activation entropy, and activation enthalpy,
respectively.

2.2 Multiphysics Model

The multiphysics model used in this work is similar to one introduced and detailed previously by Xiao and
Chaudhuri [4]. The following explains its key features, including certain variations in reaction kinetics
compared to the one presented in [4].

2.2.1 Material Balance

Both mass transport and reaction kinetics have a direct impact on the dynamical changes in mass for each
chemical species 𝑗 in the electrolyte solution, which can be expressed mathematically as follows:
𝜕𝐶𝑗
= −∇ ∙ ⃗⃗⃗
𝑁𝑗 + 𝑅𝑗 (2)
𝜕𝑡

where 𝐶, 𝑁 ⃗ , and 𝑅 represent the concentration (𝑚𝑜𝑙 ⁄𝑚3 ), flux density (𝑚𝑜𝑙 ⁄(𝑚2 𝑠)), and reaction rate
(𝑚𝑜𝑙 ⁄(𝑚3 𝑠)), respectively. Since species mass transfer is driven by differences in electrical or chemical
⃗ for the dilute solution can be expressed in
potential between two sites in an aqueous solution, the flux density 𝑁
terms of convection, migration, and diffusion:
𝐷𝑗
⃗⃗⃗
𝑁𝑗 = −𝐷𝑗 ∇𝐶𝑗 − 𝑧𝑗 𝐹 𝐶 ∇𝜑 + 𝐶𝑗 𝑣 (3)
𝑅𝑔 𝑇 𝑗

where 𝐷 is the diffusion coefficient (𝑚2 /𝑠); 𝑧 is the charge number; 𝑅𝑔 is the gas constant (8.314 𝐽/(𝑚𝑜𝑙 𝐾));
𝜑 denotes the electrostatic potential (𝑉); and 𝑣 is the flow velocity, which is zero in the case of a stagnant
electrolyte solution.

The general form of the chemical production or consumption rate can be given as [56]
𝑁𝑟
𝑚 𝑚
𝑅𝑗 = ∑ {−𝜔𝑗𝑚 (𝑘𝑓𝑚 ∏ (𝐶𝑙 )𝜔𝑙 − 𝑘𝑏𝑚 ∏ (𝐶𝑙 )−𝜔𝑙 )} (4)
𝑚=1 ∀𝜔𝑙𝑚 >0 ∀𝜔𝑙𝑚 <0

where 𝑁 𝑟 , 𝜔𝑗𝑚 , 𝑘𝑓𝑚 , and 𝑘𝑏𝑚 are the total number of reactions, the stoichiometric coefficient of species 𝑙 in the
𝑚th reaction, the forward reaction rate constant, and the backward reaction rate constant, respectively. For
reaction products, the stoichiometric coefficient is negative, while for reactants, it is positive. Historically, all
homogeneous bulk reactions were assumed to be in equilibrium at any given position in solution and time. This
assumption is avoided in our work because it could lead to significant errors in calculating chemical species
concentrations at high dissolution current densities [3].

5
The electroneutrality condition given below allows us to solve equation (2) to obtain both the electrostatic
potential and the species concentration for the dilute solution:

∑ 𝑧𝑗 𝐶𝑗 = 0 (5)
𝑗

2.2.2 Surface Reaction Mechanism

This section focuses on interfacial phenomena so that interfacial kinetics, material-solution chemistry, and
surface morphology can be mathematically correlated.

2.2.2.1. Anodic Kinetics

The pit state transition resulting from rivalry between passive oxide layer formation and dissolution is used in
this multiphysics model [4, 57]. The total current density (𝐴⁄𝑚2 ) associated with the anode can be expressed as
follows:
𝑖𝑎 = (𝑖𝑎,1 + 𝑖𝑎,2 )(1 − 𝜃0 ) (6)

where 𝑖𝑎,1 and 𝑖𝑎,2 denote the magnitudes of active dissolution and oxide formation current densities,
respectively. 𝜃0 is the oxide coverage fraction:
𝑖𝑎,2
𝜃0 = (7)
𝑖𝑎,2 + 𝑖𝑎,𝑃

in which 𝑖𝑎,𝑃 is the current density of oxide dissolution. Based on Anderko et al. [58], adsorption of aggressive
and inhibitive species is respectively responsible for active dissolution and oxide formation. The current
densities in equation (6) can be quantitatively correlated to temperature 𝑇, electrostatic potential 𝜑, applied
potential 𝜓, electrolyte chemistry, and materials as expressed below [4, 57]:
0
∗ 𝜔𝑗 𝛼𝑗 𝐹(𝜓 − 𝜑 − 𝐸𝑀,𝑗 )
𝑖𝑎,1 = ∑ 𝑖1,𝑗 (𝐶𝑗𝑠 ) ∙ exp ( ) (8)
𝑅𝑔 𝑇
𝑗

0
∗ 𝜔𝑗 𝛼𝑗 𝐹(𝜓 − 𝜑 − 𝐸𝑀𝑂,𝑗 )
𝑖𝑎,2 = ∑ 𝑖2,𝑗 (𝐶𝑗𝑠 ) ∙ exp ( ) (9)
𝑅𝑔 𝑇
𝑗

where 𝑖 ∗ , 𝐶 𝑠 , and 𝜔 are the concentration-independent part of exchange current density [57], species
concentration around the electrode surface [58], and reaction order, respectively. 𝐸 0 and 𝛼 are the reversible
potential and electrochemical transfer coefficient, respectively. The subscript 𝑗 represents an aggressive ion for
𝑖𝑎,1 and an inhibitive ion for 𝑖𝑎,2 . The reaction kinetics, which is strongly dependent on the solution chemistry,
can be written as:
𝜔𝑗
𝑖𝑎,𝑃 = ∑ 𝑘𝑃,𝑗 (𝐶𝑗𝑠 ) (10)
𝑗

The corrosion rate can be obtained from the following relationship:

6
𝑀𝑚 ∫𝜕Ω𝑒−𝑎 𝑖𝑎 𝑑𝑆
𝑅𝐶𝑜𝑟𝑟 = (11)
𝑧𝑚 𝐹 ∫ 𝑑𝑆
𝜕Ω𝑒−𝑎

in which 𝜕Ω𝑒−𝑎 = 𝜕Ω𝑎 ∩ 𝜕Ω𝑒−𝑚 .

2.2.2.2 Cathodic Kinetics

The general format of the kinetic expression for electrochemical reduction reactions on cathodes is as follows:

∗ 𝜔𝑗 𝛼𝑗 𝐹(𝜓 − 𝜑 − 𝐸𝑗0 )
𝑖𝑐 = − ∑ 𝑖𝑐,𝑗 (𝐶𝑗𝑠 ) ∙ exp (− ) (12)
𝑅𝑔 𝑇
𝑗

where the subscript 𝑗 represents a dissolved ion to be reduced.

2.2.2.3 Interface Boundary Evolution

The following relation can be used to track the position of the metal-electrolyte interface, Λ, during pit growth:

𝜕Λ 𝑀𝑚
= −𝑖𝑒−𝑚,𝑚 𝑛⃗ ⃗ ∈ 𝜕Ω𝑒−𝑚 (𝑡)
∀Λ (13)
𝜕𝑡 𝜌𝑚 𝑧𝑚 𝐹

where 𝑀𝑚 , 𝜌𝑚 , 𝑧𝑚 , and 𝑛⃗ denote the metal molecular weight, density, effective valence, and outward-pointing
unit normal vector on the electrolyte domain, respectively. 𝑖𝑒−𝑚,𝑚 can be determined using the following
function, which is time and location dependent:

𝑖𝑎 ⃗ ∈ 𝜕Ω𝑎 (𝑡) ∩ 𝜕Ω𝑒−𝑚 (𝑡)


∀Λ
𝑖𝑒−𝑚,𝑚 = { 𝑖𝑐,𝑚 ⃗ ∈ 𝜕Ω𝑐 (𝑡) ∩ 𝜕Ω𝑒−𝑚 (𝑡)
∀Λ (14)
0 ⃗ ∈ 𝜕Ω𝑃 (𝑡) ∩ 𝜕Ω𝑒−𝑚 (𝑡)
∀Λ

where 𝑖𝑐,𝑚 denotes the metal-ion-reduction component of total cathodic current density [4]. According to
Equations (13) and (14), 𝑖𝑎 determines the speed at which the pit propagates through the anode. There is no
material transfer between the oxide layer and electrolyte solution.

2.2.2.4 Initial and Boundary Conditions

The multiphysics model is based on mass transport phenomena in electrolytes with moving pit boundaries. In
this study, the following relationships were considered for the initial conditions:
⃗ , 𝑡 = 0) = 𝐶𝑗∞
𝐶𝑗 (Λ ⃗ ∈ Ω𝑒
∀Λ (15)

⃗ , 𝑡 = 0) = 𝜑∞
𝜑(Λ ⃗ ∈ Ω𝑒
∀Λ (16)

in which 𝐶𝑗∞ and 𝜑 ∞ denote the bulk concentration of species 𝑗 and the electrostatic potential, respectively.
Other conditions that can be applied to the boundaries far from the pit are as follows:

7
⃗ , 𝑡 > 0) = 𝐶𝑗∞
𝐶𝑗 (Λ ⃗ ∈ . ∂Ω𝑒 ⁄∂Ω𝑒−𝑚
∀Λ (17)

⃗ , 𝑡 > 0) = 𝜑∞
𝜑(Λ ⃗ ∈ . ∂Ω𝑒 ⁄∂Ω𝑒−𝑚
∀Λ (18)

While the initial boundary ∂Ω𝑒−𝑚 (𝑡 = 0) is associated with the surface morphology obtained from
experiments, equations (13) and (14) determine the boundary location ∂Ω𝑒−𝑚 (𝑡 > 0) during the corrosion
process. The flux of chemical species involved in interfacial reactions can be obtained using the following
relationship:
−𝑖𝑒−𝑚,𝑗
⃗⃗⃗ ⃗ , 𝑡 > 0) =
𝑁𝑗 (Λ 𝑛⃗ ⃗ ∈ ∂Ω𝑒−𝑚
∀Λ (19)
𝑧𝑗 𝐹

where 𝑖𝑒−𝑚,𝑗 denotes the current density resulting from an electrochemical reaction involving species 𝑗 (see
equation (14) to obtain this term).

3. Case Study: Aluminum Alloy with Zr-Based Conversion Coating

In this work, the presented multiscale modeling approach was used to investigate localized pitting corrosion in
aluminum alloys coated with a zirconia layer and immersed in NaCl solution (Figure 1(a)). To better understand
the critical phenomena that occur during corrosion, some assumptions were considered in order to simplify the
current problem. The following are the assumptions: 1) dealloying of Al2Cu is inhibited by covering a passive
layer on the θ-phase, as found in experiments [10], 2) Cu-plating on the θ-phase is ignored after covering this
phase with an insoluble CuCl passive layer, 3) the zirconia layer (i.e., Ω𝑃 ) remains intact during the simulations,
and 4) Al(OH)3 precipitation as a corrosion product on the pit is ignored due to its low concentration when
𝑝𝐻 < 4 [3]. Because of hydrolysis reactions, the solution in an active pit is very acidic [59].

3.1 Reaction Mechanisms

The kinetics model associated with the localized corrosion of an aluminum alloy coated with a zirconia layer
consists of homogeneous bulk reactions involving both Al and Zr species. Table 1 contains a list of these
reactions, which were compiled from various works of literature.

In anodic region, a layer of oxide film can form as a result of the following reactions, which causes
repassivation of the active pit in aluminum alloys:
≡ 2Al + 3𝐻2 𝑂 → ≡ Al2 𝑂3 + 6𝐻 + + 6𝑒 − for acidic to neutral solution (20)

≡ 2Al + 6OH − → ≡ Al2 𝑂3 + 3𝐻2 𝑂 + 6𝑒 − for alkaline solution (21)

On the other hand, dissolution reactions of the oxide film at different pH levels are as follows:
≡ Al2 𝑂3 + 6𝐻 + → 2Al3+ + 3𝐻2 𝑂 for acidic to neutral solution (22)

≡ Al2 𝑂3 + 2OH − → 2AlO2 − + 𝐻2 𝑂 for alkaline solution (23)

In the cathodic region, however, oxygen reduction is assumed to be the only cathodic reaction on the Cu
surface:

8
𝑂2 + 2𝐻2 𝑂 + 4𝑒 − → 4𝑂𝐻 − (24)

3.2 Simulation Inputs

The material properties used in the modeling for aluminum and zirconium are: 𝑀𝑤,𝐴𝑙 = 26.98 𝑔/𝑚𝑜𝑙, 𝜌𝐴𝑙 =
2700 𝑘𝑔/𝑚3, 𝑀𝑤,𝑍𝑟 = 91.22 𝑔/𝑚𝑜𝑙, and 𝜌𝑍𝑟 = 6506 𝑘𝑔/𝑚3 . The following dissolved species were
considered in the simulations: Zr4+, ZrO2+, ZrOH3+, Zr(OH)22+, ZrCl3+, Zr(OH)Cl2+, ZrOCl+, Zr(OH)3Cl,
ZrO(OH)Cl, Al3+, Al(OH)2+, Al(OH)2+, Al2(OH)24+, AlCl2+, Al(OH)Cl+, Al(OH)2Cl, Cl ̄, H+, OH ̄, Na+, and O2.
The diffusion coefficients for these species were collected from the literature [3, 60]. The diffusion coefficient
for Zr-containing species was considered to be the same and equal to 2.36 × 10−11 𝑚2 /𝑠 [60].

The component was immersed in a NaCl solution at room temperature, 298.15 𝐾. The concentration of the
solution was considered to have a range from 10−4 𝑀 to 1 𝑀. A bulk concentration of 60 𝜇𝑀 was set for the
dissolved oxygen in the electrolyte [61]. The applied potentials were changed from −1 𝑉 to −0.55 𝑉 (vs.
saturated calomel electrode (SCE)). Note that this study aims at investigating the full anodic polarization
behavior by changing the applied potentials. Furthermore, the free corrosion case, in which the applied
potentials equal the open circuit potential (OCP), is not explored.

4. Results and Discussion

4.1 Evaluation of Kinetics and Multiscale Models

To complete the kinetics model for multiphysics simulations, DFT-TST calculations were carried out to
determine the missing reaction constants and energy barriers for first-order reactions involving Zr. However,
reaction constants for a few reactions containing Al were obtained for evaluating the calculations due to a lack
of experimental data for the rate constants and energy barriers of Zr-containing reactions. Table 1 shows the
results of calculations that include the forward-reaction rate constant and the energy barrier. Comparisons of
computed and experimental values for the kinetic parameters indicate that for reactions containing Al,
calculations overestimated the rate constants by less than 50%. However, the results of the calculations may fall
within the measurement error, which has not been reported in the literature. In addition, the difference in
reaction rate constants between experiments and DFT-TST calculations can be attributed to the variation of
conditions under which the data are obtained [62]. The rate constants used in the simulations are a combination
of data from both experiments and DFT-TST calculations, as shown in the last two columns of Table 1. Due to
very expensive computations, the values for the second-order reaction rates containing Zr were assumed in this
study.

To assess the multiscale models, the corrosion rates of bare aluminum and zirconium obtained from simulations
were compared to the experimental results in Table 2. The simulation inputs were set to match the experimental
conditions. Model results are in good agreement (almost in the same ballpark) with measurements reported in
the literature. Differences in corrosion rates between simulations and experiments can be explained by
differences in the conditions under which the data is obtained. Along with the results for Al 6061 and Zr
705/702, the corrosion rate of Al 6061 with Zr-based conversion coating is reported for the same conditions as
bare Al 6061 to better understand the effects of Zr-based coatings on the corrosion performance of aluminum
alloys. However, there is no experimental data for Al coated with ZrO2 in the literature. According to the
simulation results in Table 2, Zr-based conversion coating improves the corrosion performance of aluminum

9
alloys immersed in NaCl solution, which is in line with experimental findings [17, 63-65]. The reasons will be
explained in the following section.

10
Table 1: Homogeneous bulk reactions proposed for the dissolution of aluminum alloy with Zr-based conversion coating immersed in chloride solution. The energy barrier ΔG and rate constants for
forward reaction Kf and backward reaction Kb is used to compare DFT calculations performed in this work with experimental data. The rate constants selected for experiments are a combination of
experimental and DFT data.

DFT – Transition State Selected Rate Constants for


Experiment
Theory Simulations
Reaction
Reaction
Type
ΔG
Kf Kf Kb Ref. a Kf Kb
(Cal/mol)

4.49×102 [66, 4.49×102 1×107


Zr4+ + H2O ↔ ZrOH3+ + H+ 1.37×104 - -
s−1 67] s−1 M−1 s−1 c

1.24×10-6 1.24×10-6 1×102


Zr4+ + H2O ↔ ZrO2+ + 2H+ 2.53×104 - - [68]
s−1 s−1 M−1 s−1 c

1.01×107 1.01×107 1×109


ZrOH3+ + H2O ↔ Zr(OH)22+ + H+ 7.81×103 - - [66]
s−1 s−1 M−1 s−1 c
Hydrolysis
1.58×105 1.09×105 4.4×109 1.09×105 4.4×109
Al3+ + H2O ↔ Al(OH)2+ + H+ 1.03×104 [69]
s−1 s−1 M−1 s−1 s−1 M−1 s−1

4.87×107 4.87×107 4.4×109


Al(OH)2+ + H2O ↔ Al(OH)2+ + H+ c 6.89×103 - - [3]
s−1 s−1 M−1 s−1 c

10-2 108 10-2 108


2Al3+ + 2H2O ↔ Al2(OH)24+ + 2H+ - - [69]
M−1 s−1 M−1 s−1 M−1 s−1 M−1 s−1

100 10
Zr4+ + Cl- ↔ ZrCl3+ - - - - [70]
M−1 s−1 c
s−1 c

10 1000
ZrOH3+ + Cl- ↔ Zr(OH)Cl2+ - - - - [70]
Reactions M−1 s−1 c
s−1 c
with
Cl− ion 226 75 − 𝑘𝑓 . (𝑐𝐴𝑙3+ − 226 75 − 𝑘𝑓 . (𝑐𝐴𝑙3+ −
Al3+ + Cl- ↔ AlCl2+ - - [71]
M−1 s−1 𝑐𝐴𝑙𝑂𝐻 2+ ) s−1 M−1 s−1 𝑐𝐴𝑙𝑂𝐻 2+ ) s−1

19 5700 − 𝑘𝑓 . 𝑐𝐴𝑙𝑂𝐻 2+ 19 5700 − 𝑘𝑓 . 𝑐𝐴𝑙𝑂𝐻 2+


Al(OH)2+ + Cl- ↔ Al(OH)Cl+ - - [71]
M s−1
−1
s−1 M s−1
−1
s−1

11
3.62×10−4 3.62×10−4
ZrCl3+ + 3H2O → Zr(OH)3Cl + 3H+ 2.19×104 - - b
-
s−1 s−1

5.26×10−4 5.26×10−4
Zr(OH)Cl2+ + 2H2O → Zr(OH)3Cl + 2H+ 2.17×104 - - b
-
s−1 s−1

Product 1.44×10−6 1.44×10−6


ZrOCl+ + H2O → ZrO(OH)Cl + H+ 2.52×104 - - b
-
Formation s−1 s−1

5.94×10−6 4×10−6 4×10−6


AlCl2+ + 2H2O → Al(OH)2Cl + 2H+ 2.43×104 - [72] -
s−1 s−1 s−1

1.74×10−3 4×10−6 4×10−6


Al(OH)Cl+ + H2O → Al(OH)2Cl + H+ 2.10×104 - [72] -
s−1 s−1 c s−1

Water 4.37×10−5 2.6×10−5 1.3×1011 2.6×10−5 1.3×1011


H2O ↔ OH- + H+ 2.32×104 [73]
Dissociation s−1 s−1 M−1 s−1 s−1 M−1 s−1

a
References are associated with both reactions and experimental data for reaction constants.
b
Reactions are proposed in this study.
c
Assumed.

12
Table 2: Corrosion rate of bare aluminum alloy 6061, bare zirconium 705/702, and aluminum alloy 6061 with Zr-based conversion coating obtained
from simulations in this work and reported experimental measurements. Certain conditions for simulations and experimental measurements are also
mentioned for the purpose of comparison.

Simulations Experiment
Metal
RCorr RCorr
Condition Condition Ref.
(mm/year) (mm/year)

NaCl (1%),
NaCl, 30 °C,
30 - 50 °C,
Al 6061 pH = 6, 0.020899 0.0305 [74]
pH = 5.8 – 6.0,
Steady
after 480 hr

NaCl (25%),
NaCl, 30 °C,
35-100 °C,
Zr 705/702 pH = 1, 0.0005506 < 0.0007 [75, 76]
pH = 1,
Steady
after 504 hr

NaCl, 30 °C,
Al with ZrO2 NaCl, 30 °C, 0.005588 ±
pH = 6, 0.002254 [77]
Coating pH = 6.5 – 7.2 0.0038
Steady

4.2 Environmental Effects on the Growth of Pit

To understand the effects of chloride ions (i.e., 𝐶𝑙 − ) as aggressive species on localized pitting corrosion, the
change in corrosion rate of bare Al 6061, bare Zr 705, and Al 6061 coated with Zr-based conversion coating as
a function of the bulk concentration of chloride ions in the solution is plotted in Figure 2(a). Here, the alloy was
immersed in a neutral NaCl solution and was potentiostatically controlled with a −0.55 𝑉 applied potential. A
fixed pit geometry corresponding to a particular instant of time during pit propagation was used. The figure
shows that the corrosion rate of the systems increases as the bulk concentration of 𝐶𝑙 − was increased. This is
because increasing the bulk concentration of 𝐶𝑙 − in the solution drives more chloride ions to the pit surface,
resulting in increased acidity in the pit (see Figure 2(b)) and thus increased corrosion rate of the system.
Experimental observations for localized corrosion of aluminum alloys indicated the same pattern [78].

Furthermore, Figure 2(a) demonstrates that Zr-based conversion coatings promote the corrosion resistance of
aluminum alloys due to the involvement of Zr in interfacial kinetics. This is due to the suppression of both the
oxygen reduction reaction (ORR) and anodic dissolution, with the ORR being more strongly inhibited than bare
Al 6061 due to the impediment of oxygen chemisorption on IMPs [79]. The corrosion rate of Zr is very low in
comparison to the other two systems. Figure 2(b) depicts the pH distribution along the pit surface, specifically
close to the electrolyte-metal interface. With a constant pH, the pit status changes from active to passive as the
concentration of 𝐶𝑙 − in the solution decreases, which is in line with the experimental observations [80]. Al3+,
AlCl2+, Zr4+, ZrCl3+, and Zr(OH)Cl2+ are the main species in neutralizing the chloride ion charges in the
solution. The present model assumes that 1) Al- and Zr-containing salts are sufficiently soluble during the early
stages of pit growth, and 2) continuous wet environments, similar to salt fog chambers, can allow for very high
concentrations of ionic species.

13
(a) (b)
Figure 2: (a) The effect of bulk concentration of 𝐶𝑙 − on the corrosion rate of bare Al 6061, bare Zr 705, and Al 6061 coated with Zr-based
conversion coating. It shows that the Zr-based conversion coating significantly improves the corrosion performance of aluminum alloys in NaCl
solution. (b) Chemical conditions near the pit surface showing how solution pH changes close to the pit surface for different concentrations of
chloride ions.

Figure 3(a-c) depicts the anodic polarization curves obtained by using a series of constant concentrations of 𝐶𝑙 −
and decreasing the applied potential 𝜓 in 10 mV steps. This plot aids in determining pit stability as the
repassivation potential can be quantified using the current approach. The repassivation potential, as described
by Frankel et al. [81], is the potential that the pit current density falls below a critical value, which is the
minimum current density required to preserve the critical pit environment and prevent repassivation. Figure 3(a-
c) shows that the corrosion rate of the systems decreases at a fixed 𝐶𝑙 − concentration as the applied potential
decreases from −0.55 𝑉. A sharp decrease in corrosion rate indicates pit repassivation as applied potentials are
reduced. The sudden growth of an aluminum oxide layer on the metal surface accurately described pit
repassivation.

Furthermore, Figure 3(a-c) demonstrates that the repassivation potential of a system (e.g., Al 6061) decreases as
the concentration of chloride ions in the solution increases. It implies that the pit surface becomes less stable
when surrounded by less concentrated aggressive ions (e.g., 𝐶𝑙 − ). Our model successfully predicted the
experimentally observed behavior for aluminum alloys [78, 80]. Figure 3(d) compares the polarization curves of
different systems (Al 6061, Zr 705, and Al 6061 with Zr-based conversion coating) at a fixed 𝐶𝑙 − concentration
of 1𝑀. In comparison to the other two systems depicted in Figure 3(d), Al 6061 has a very high repassivation
potential and a much higher corrosion rate, which is consistent with experimental observations.

14
(a) (b)

(c) (d)
Figure 3: Anodic polarization curves for (a) bare Al 6061, (b) bare Zr 705, and (c) Al 6061 with Zr-based conversion coating at different
concentrations of 𝐶𝑙 − . It shows that the repassivation potential decreases as the concentration of chloride ions in the solution increases. (d) A
comparison of anodic polarization curves at 𝑐𝐶𝑙− = 1𝑀 for different systems.

To study the effects of pH on the corrosion rate of the systems, the concentration of chloride ions in the solution
was kept constant at 0.01M. Either hydrochloric acid (HCl) or sodium hydroxide (NaOH) was used to adjust the
bulk pH of the solution. Experiments have shown that when aluminum alloys are exposed to alkaline solutions,
general corrosion occurs rather than localized pitting corrosion because uniform thinning of the passive oxide
film (due to OH ion attack) overwhelms pitting corrosion [78]. Because localized corrosion is the focus of this
research, acidic or near-neutral solutions (i.e., pH range 1 to 8) have been examined.

Figure 4(a) compares the corrosion rate of Al 6061 and Al 6061 with Zr-based conversion coating as a function
of solution pH at a fixed applied potential of −0.55 𝑉. The decrease in pH results in an increase in corrosion
rate, which is in line with the experimental findings [82]. According to Xiao and Chaudhuri [4], there is a

15
critical pH above which the corrosion rate rapidly decreases to zero. Such a significant decrease shows that the
pit status has changed from active to passive. The critical pH values for bare Al 6061 and Al 6061 coated with
ZrO2 are approximately 6 and 5, respectively, indicating that the presence of Zr changes the pit status in Al
6061 from active to passive at lower pH. The presence of Zr near the pit surface, on the other hand, indicates
that pit stabilization is accelerated.

Figure 4(b) and (c) depict how the corrosion rate of Al 6061 with Zr-based conversion coating changes with pH
at various applied potentials. The corrosion rate of the system increases as the solution pH decreases for
different potentials due to the more acidic environment around the pit surface. At a certain pH, the corrosion
rate increases as the applied potential increases. Furthermore, changes in pH between 4 and 8 have almost no
effect on the corrosion rate. Figure 4(d) shows a 3D pH-potential diagram that can be used to quickly determine
the corrosion rate and pit stability of Al 6061 with Zr-based conversion coating under any corrosive conditions.
This 3D diagram expands on the concept of Pourbaix diagram, which shows the equilibrium phases of specific
metal-electrolyte systems at various pH and potentials, by adding a new dimension of information, namely,
corrosion rate. According to Figure 4(d), a more acidic environment near the pit surface results in the
continuous growth of an active pit at a lower applied potential.

(a) (b)

16
(c) (d)
Figure 4: The effects of pH and applied potential on the performance of localized corrosion: (a) a comparison of the corrosion rates of bare Al 6061
and Al 6061 coated with ZrO2 as a function of pH at a fixed applied potential of −0.55 𝑉, (b) and (c) pH effects on the corrosion rate of Al 6061
coated with ZrO2 at different applied potentials, and (d) a 3D corrosion rate-pH-applied potential diagram for Al 6061 coated with ZrO2. Because of
the presence of Zr near the pit surface, the corrosion rate of Al 6061 coated with ZrO 2 becomes nearly zero at lower pH values compared to that of
bare Al 6061.

To analyze corrosive environment within the active pit in the Al 6061 coated with ZrO2, Figure 5(a-h) shows
2D contour plots of pH and electrostatic potential for different concentrations of chloride ions in the system.
The pH contour plots indicate that the solution aggressivity increases monotonically from the pit mouth to the
pit bottom. Figure 5(a, c, e, g) shows that as the concentration of chloride ions increases, the electrolyte within
the pit becomes more acidic (i.e., has a higher H+ concentration). This event accelerates the corrosion of
systems. Figure 5(b, d, f, h) demonstrates that the electrostatic potentials increase from the pit mouth to the pit
bottom, indicating that the ohmic potential drop in the system increases. Both ohmic potential drop and solution
aggressivity affect the corrosion rate of the alloy, with the former contributing negatively and the latter
positively. It should be noted that the ohmic potential effect becomes negligible as the applied potential
approaches the repassivation potential [4].

(a) (b)

17
(c) (d)

(e) (f)

(g) (h)
Figure 5: Corrosive environment within an active pit of Al 6061 coated with ZrO2: (a, c, e, and g) 2D contour plots of solution pH, and (b, d, f, and
h) 2D contour plots of electrostatic potential. Plots are shown for four different concentrations of chloride ions: 0.001, 0.01, 0.1, and 1 M. The unit

18
of the pit dimension in r and z directions is meter. Figures indicate that ohmic and chemistry effects both contribute to the corrosion rate of a system,
which must be taken into account as a coupled effect.

4.3 Pit Growth Dynamics

To study the dynamics of pit growth around IMPs (θ-phase), the pit depicted in Figure 1(a) is set to propagate in
a neutral sodium chloride electrolyte (0.01 M) at an applied potential of −0.55 𝑉. Figure 6 shows that the
corrosion rate of both systems decreases with time throughout the corrosion process, which is consistent with
experimental observations [59, 83]. The corrosion rate of Al 6061 coated with ZrO2 approaches a near constant
value during pit propagation much earlier than that of bare Al 6061. This is due to the presence of positively
charged Al- and Zr-containing species in the pit that neutralize the charges and modify the local chemical
environment close to the pit surface. The electrolyte solution becomes more aggressive over time (i.e., higher
concentration of chloride ions and lower pH), and the ohmic potential drop increases in the system. Thus, the
net trend of the change in corrosion rate is determined by a competition between chemistry and ohmic effects.

Figure 6: Dynamics of corrosion rate during pit growth for bare Al 6061 and Al 6061 with Zr-based conversion coating.

Figure 7(a-f) shows 2D contour plots of the growing pit on Al 6061 coated with ZrO2 in the r and z – directions
at different time instances. Because of the heterogeneous microstructure of aluminum alloys and moving
boundary, the nature of electrochemical reactions in various locations of the pit surface was distinct and
dynamically varied with time. As the pit grows in the system, more θ-phase surface is exposed. Thus, the ORR
occurs on the cathodic regions (i.e., IMPs) and produces 𝑂𝐻 − ions, which neutralizes the acidic solution in the
pit, resulting in a higher pH value. Consequently, the chemistry effect (i.e., solution aggressivity) on the
corrosion rate of system decreases as time proceeds. The progressively exposed Al2O3/ZrO2 protection causes
the pit to expand and form an occluded geometry, which contributes to a higher potential drop within the pit. As
a result, while the pit grows in size, the ohmic effect on corrosion of Al 6061 coated with ZrO2 becomes more
pronounced. It can be concluded that the ohmic effect outweighed the chemistry effect (which was hampered by
the cathodic reaction on IMPs), resulting in an overall decrease in corrosion rate during the pit propagation.
Figure 7(a-f) also presents the 2D contour plots of the growing pit on Al 6061 in r and z – directions at 0.45
seconds. At 2.494 s, this instance exhibits the same pit development as the Al 6061 coated with ZrO2 instance,
indicating that corrosion progress is much slower in Al 6061 coated with ZrO2 than in bare Al 6061.

19
Aluminum 6061 Coated with ZrO2

(a) (b)

(c) (d)

(e) (f)

20
Bare Aluminum 6061

(g) (h)
Figure 7: 2D contour plot of the growing pit in Al 6061 coated with ZrO2 (a) Ur at 0.45 s, (b) Uz at 0.45 s, (c) Ur at 1.5 s, (d) Uz at 1.5 s, (e) Ur at
2.494 s, and (f) Uz at 2.494 s, and in bare Al 6061 (g) Ur at 0.45 s and (h) Uz at 0.45 s. Ur and Uz are the displacement in r and z directions,
respectively. The unit of pit dimension in the r and z directions is meter. Comparisons show that corrosion progress in Al 6061 coated with ZrO2 is
much slower than in bare Al 6061.

5. Conclusion

A multiscale modeling framework was developed to predict the corrosion rate and pit stability of aluminum
alloys with Zr-based conversion coatings. The kinetics model, which included bulk interfacial reactions, was
completed using first-principles calculations and transition-state theory. By incorporating the kinetics model
into multiphysics simulations, the corrosion performance of aluminum alloys with Zr-based conversion coatings
was quantified for various local corrosive environments within pits, which is very difficult to establish using in
situ experiments. The simulation results were compared to the results of experiments reported in the literature,
and they showed excellent agreements for corrosion rates of different alloys. Furthermore, good agreement was
found between the DFT-TST results and experimental data for the rate constants of bulk first-order reactions
occurring in the electrolyte. Later, the model was employed to quantitatively explore the effects of local
corrosive environements on the corrosion performance of aluminum alloys coated with ZrO2. The local
corrosive fields (i.e., chemical and electrical conditions) were determined within the localized pits at different
solution conditions. The following are some of the key findings from simulation results for certain
electrochemical conditions as elucidated in this work: 1) a Zr-based conversion coating significantly improves
the corrosion performance of aluminum alloys due to the presence of zirconium in interfacial kinetics, 2) the
repassivation potential of a system (e.g., Al 6061) decreases as the concentration of chloride ions in the solution
increases, leading to a less stable pit, 3) the critical pH value for Al 6061 coated with ZrO2 is lower than for
bare Al 6061 because the presence of Zr changes the pit status from active to passive at lower pH, and 4) the
corrosion rate of Al 6061 coated with ZrO2 reaches an almost constant value during pit propagation, much
earlier than that of bare Al 6061, due to the attendance of Al- and Zr-containing species in the pit that modify
the local chemical environment close to the pit surface.

Data Availability

All the required data that support the findings of this study are available in the paper.

21
Acknowledgments

This work was supported by the Strategic Environmental Research and Development Program (SERDP) under
project WP-2742. We would like to thank the experimental team at the Center for Plasma Materials Interactions
(CPMPI) of the University of Illinois at Urbana-Champaign for their sustained effort in developing these
coatings using atmospheric pressure plasma and providing us with corrosion data of the components.

Reference

[1] A. Boag, R. Taylor, T. Muster, N. Goodman, D. McCulloch, C. Ryan, B. Rout, D. Jamieson, A. Hughes,
Stable pit formation on AA2024-T3 in a NaCl environment, Corrosion Sci. 52(1) (2010) 90-103.

[2] A. Hughes, A. Boag, A. Glenn, D. McCulloch, T. Muster, C. Ryan, C. Luo, X. Zhou, G. Thompson,
Corrosion of AA2024-T3 Part II: Co-operative corrosion, Corrosion Sci. 53(1) (2011) 27-39.

[3] O. Guseva, P. Schmutz, T. Suter, O. von Trzebiatowski, Modelling of anodic dissolution of pure aluminium
in sodium chloride, Electrochim. Acta 54(19) (2009) 4514-4524.

[4] J. Xiao, S. Chaudhuri, Predictive modeling of localized corrosion: an application to aluminum alloys,
Electrochim. Acta 56(16) (2011) 5630-5641.

[5] A. Boag, A. Hughes, N. Wilson, A. Torpy, C. MacRae, A. Glenn, T. Muster, How complex is the
microstructure of AA2024-T3?, Corrosion Sci. 51(8) (2009) 1565-1568.

[6] R. Buchheit, R. Grant, P. Hlava, B. McKenzie, G. Zender, Local dissolution phenomena associated with S
phase (Al2CuMg) particles in aluminum alloy 2024‐T3, Journal of the electrochemical society 144(8) (1997)
2621.

[7] Y. Yoon, R. Buchheit, Dissolution behavior of Al2CuMg (S phase) in chloride and chromate conversion
coating solutions, Journal of The Electrochemical Society 153(5) (2006) B151.

[8] N. Birbilis, R. Buchheit, Investigation and discussion of characteristics for intermetallic phases common to
aluminum alloys as a function of solution pH, Journal of the Electrochemical Society 155(3) (2008) C117.

[9] K.D. Ralston, T.L. Young, R.G. Buchheit, Electrochemical evaluation of constituent intermetallics in
aluminum alloy 2024-T3 exposed to aqueous vanadate inhibitors, Journal of the electrochemical society 156(4)
(2009) C135.

[10] R. Grilli, M.A. Baker, J.E. Castle, B. Dunn, J.F. Watts, Localized corrosion of a 2219 aluminium alloy
exposed to a 3.5% NaCl solution, Corrosion Sci. 52(9) (2010) 2855-2866.

[11] S.-I. Pyun, E.-J. Lee, Effect of halide ion and applied potential on repassivation behaviour of Al-1 wt.% Si-
0.5 wt.% Cu alloy, Electrochim. Acta 40(12) (1995) 1963-1970.

[12] G. Williams, A.J. Coleman, H.N. McMurray, Inhibition of Aluminium Alloy AA2024-T3 pitting corrosion
by copper complexing compounds, Electrochim. Acta 55(20) (2010) 5947-5958.

[13] J. Li, N. Birbilis, R.G. Buchheit, Electrochemical assessment of interfacial characteristics of intermetallic
phases present in aluminium alloy 2024-T3, Corrosion Sci. 101 (2015) 155-164.

[14] Y. Zhu, K. Sun, G. Frankel, Intermetallic phases in aluminum alloys and their roles in localized corrosion,
Journal of The Electrochemical Society 165(11) (2018) C807.

22
[15] R. Oltra, B. Malki, F. Rechou, Influence of aeration on the localized trenching on aluminium alloys,
Electrochim. Acta 55(15) (2010) 4536-4542.

[16] A.S.H. Makhlouf, Handbook of smart coatings for materials protection, Elsevier2014.

[17] I. Milošev, G. Frankel, conversion coatings based on zirconium and/or Titanium, Journal of The
Electrochemical Society 165(3) (2018) C127.

[18] J. Zhao, G. Frankel, R.L. McCreery, Corrosion protection of untreated AA‐2024‐T3 in chloride solution by
a chromate conversion coating monitored with Raman spectroscopy, Journal of the Electrochemical Society
145(7) (1998) 2258.

[19] H.J. Gibb, P.S. Lees, P.F. Pinsky, B.C. Rooney, Lung cancer among workers in chromium chemical
production, American journal of industrial medicine 38(2) (2000) 115-126.

[20] P.L. Bidstrup, R. Case, Carcinoma of the lung in workmen in the chromates-producing industry in Great
Britain, British journal of industrial medicine 13(4) (1956) 260.

[21] J. Nordlien, J. Walmsley, H. Østerberg, K. Nisancioglu, Formation of a zirconium-titanium based


conversion layer on AA 6060 aluminium, Surface and Coatings Technology 153(1) (2002) 72-78.

[22] F. George, P. Skeldon, G. Thompson, Formation of zirconium-based conversion coatings on aluminium


and Al–Cu alloys, Corrosion Sci. 65 (2012) 231-237.

[23] X. Zhong, X. Wu, Y. Jia, Y. Liu, Self-repairing vanadium–zirconium composite conversion coating for
aluminum alloys, Appl. Surf. Sci. 280 (2013) 489-493.

[24] S.S. Golru, M. Attar, B. Ramezanzadeh, Morphological analysis and corrosion performance of zirconium
based conversion coating on the aluminum alloy 1050, J. Ind. Eng. Chem. 24 (2015) 233-244.

[25] J. Gray, B. Luan, Protective coatings on magnesium and its alloys—a critical review, Journal of alloys and
compounds 336(1-2) (2002) 88-113.

[26] J. Cerezo, I. Vandendael, R. Posner, K. Lill, J. De Wit, J. Mol, H. Terryn, Initiation and growth of modified
Zr-based conversion coatings on multi-metal surfaces, Surface and Coatings Technology 236 (2013) 284-289.

[27] J. Cerezo, P. Taheri, I. Vandendael, R. Posner, K. Lill, J.H.W. de Wit, J. Mol, H. Terryn, Influence of
surface hydroxyls on the formation of Zr-based conversion coatings on AA6014 aluminum alloy, Surface and
coatings technology 254 (2014) 277-283.

[28] X. Liu, D. Vonk, H. Jiang, K. Kisslinger, X. Tong, M. Ge, E. Nazaretski, B. Ravel, K. Foster, S. Petrash,
Environmentally friendly Zr-based conversion nanocoatings for corrosion inhibition of metal surfaces evaluated
by multimodal x-ray analysis, ACS Applied Nano Materials 2(4) (2019) 1920-1929.

[29] T. Lostak, A. Maljusch, B. Klink, S. Krebs, M. Kimpel, J. Flock, S. Schulz, W. Schuhmann, Zr-based
conversion layer on Zn-Al-Mg alloy coated steel sheets: insights into the formation mechanism, Electrochim.
Acta 137 (2014) 65-74.

[30] J. Cerezo, I. Vandendael, R. Posner, J. De Wit, J. Mol, H. Terryn, The effect of surface pre-conditioning
treatments on the local composition of Zr-based conversion coatings formed on aluminium alloys, Appl. Surf.
Sci. 366 (2016) 339-347.

[31] S. Sharland, A review of the theoretical modelling of crevice and pitting corrosion, Corrosion Sci. 27(3)
(1987) 289-323.

23
[32] M. Verhoff, R. Alkire, Experimental and modeling studies of single pits on pure aluminum in ph 11 nacl
solutions i. Laser initiated single pits, Journal of the Electrochemical Society 147(4) (2000) 1349.

[33] N. Murer, R. Oltra, B. Vuillemin, O. Néel, Numerical modelling of the galvanic coupling in aluminium
alloys: A discussion on the application of local probe techniques, Corrosion Sci. 52(1) (2010) 130-139.

[34] K.B. Deshpande, Validated numerical modelling of galvanic corrosion for couples: Magnesium alloy
(AE44)–mild steel and AE44–aluminium alloy (AA6063) in brine solution, Corrosion Sci. 52(10) (2010) 3514-
3522.

[35] L. Abodi, J. DeRose, S. Van Damme, A. Demeter, T. Suter, J. Deconinck, Modeling localized aluminum
alloy corrosion in chloride solutions under non-equilibrium conditions: steps toward understanding pitting
initiation, Electrochim. Acta 63 (2012) 169-178.

[36] L. Yin, Y. Jin, C. Leygraf, J. Pan, A FEM model for investigation of micro-galvanic corrosion of Al alloys
and effects of deposition of corrosion products, Electrochim. Acta 192 (2016) 310-318.

[37] O. Guseva, J. DeRose, P. Schmutz, Modelling the early stage time dependence of localised corrosion in
aluminium alloys, Electrochim. Acta 88 (2013) 821-831.

[38] W. Mai, S. Soghrati, New phase field model for simulating galvanic and pitting corrosion processes,
Electrochim. Acta 260 (2018) 290-304.

[39] W. Mai, S. Soghrati, R.G. Buchheit, A phase field model for simulating the pitting corrosion, Corrosion
Sci. 110 (2016) 157-166.

[40] L. Yin, Y. Jin, C. Leygraf, N. Birbilis, J. Pan, Numerical simulation of micro-galvanic corrosion in Al
alloys: effect of geometric factors, Journal of The Electrochemical Society 164(2) (2017) C75.

[41] S.R. Cross, S. Gollapudi, C.A. Schuh, Validated numerical modeling of galvanic corrosion of zinc and
aluminum coatings, Corrosion Sci. 88 (2014) 226-233.

[42] L. Yin, Y. Jin, C. Leygraf, J. Pan, Numerical simulation of micro-galvanic corrosion of Al alloys: Effect of
chemical factors, Journal of The Electrochemical Society 164(13) (2017) C768.

[43] A. Sumer, S. Chaudhuri, A first principles investigation of corrosion chemistry of common elemental
impurities in Mg-Al alloys, Corrosion 73(5) (2017) 596-604.

[44] S.R. Cross, C.A. Schuh, Modeling localized corrosion with an effective medium approximation, Corrosion
Sci. 116 (2017) 53-65.

[45] N. Li, C. Dong, C. Man, X. Li, D. Kong, Y. Ji, M. Ao, J. Cao, L. Yue, X. Liu, Insight into the localized
strain effect on micro-galvanic corrosion behavior in AA7075-T6 aluminum alloy, Corrosion Sci. 180 (2021)
109174.

[46] Y. Zhu, J.D. Poplawsky, S. Li, R.R. Unocic, L.G. Bland, C.D. Taylor, J.S. Locke, E.A. Marquis, G.S.
Frankel, Localized corrosion at nm-scale hardening precipitates in Al-Cu-Li alloys, Acta Materialia 189 (2020)
204-213.

[47] L. Bailly-Salins, L. Borrel, W. Jiang, B.W. Spencer, K. Shirvan, A. Couet, Modeling of High-Temperature
Corrosion of Zirconium Alloys Using the eXtended Finite Element Method (X-FEM), Corrosion Sci. (2021)
109603.

24
[48] S.K. Sankaranarayanan, S. Ramanathan, Molecular dynamics simulation study of nanoscale passive oxide
growth on Ni-Al alloy surfaces at low temperatures, Physical review b 78(8) (2008) 085420.

[49] B. Diawara, M. Legrand, J.-J. Legendre, P. Marcus, Use of quantum chemistry results in 3D modeling of
corrosion of iron-chromium alloys, Journal of the Electrochemical Society 151(3) (2004) B172.

[50] D. Gunasegaram, M. Venkatraman, I. Cole, Towards multiscale modelling of localised corrosion, Int.
Mater. Rev. 59(2) (2014) 84-114.

[51] S. Dong, W. Shi, J. Zhang, S. Bi, DFT studies on the water-assisted synergistic proton dissociation
mechanism for the spontaneous hydrolysis reaction of Al3+ in aqueous solution, ACS Earth and Space
Chemistry 2(3) (2018) 269-277.

[52] A.V. Marenich, C.J. Cramer, D.G. Truhlar, Universal solvation model based on solute electron density and
on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions, The
Journal of Physical Chemistry B 113(18) (2009) 6378-6396.

[53] A.D. Beck, Density-functional thermochemistry. III. The role of exact exchange, J. Chem. Phys 98(7)
(1993) 5648-6.

[54] W. Shi, X. Jin, S. Dong, S. Bi, Theoretical investigation of the thermodynamic structures and kinetic
water-exchange reactions of aqueous Al (III)–salicylate complexes, Geochim. Cosmochim. Acta 121 (2013) 41-
53.

[55] F.P. Rotzinger, The Water‐Exchange Mechanism of the [UO2 (OH2) 5] 2+ Ion Revisited: The Importance
of a Proper Treatment of Electron Correlation, Chemistry–A European Journal 13(3) (2007) 800-811.

[56] D. Macdonald, G. Engelhardt, Predictive modeling of corrosion, Shreir’s Corros, Elsevier BV 2 (2010)
1630-1679.

[57] A. Anderko, N. Sridhar, D. Dunn, A general model for the repassivation potential as a function of multiple
aqueous solution species, Corrosion Sci. 46(7) (2004) 1583-1612.

[58] G. Engelhardt, D.D. Macdonald, Unification of the deterministic and statistical approaches for predicting
localized corrosion damage. I. Theoretical foundation, Corrosion Sci. 46(11) (2004) 2755-2780.

[59] K.P. Wong, R.C. Alkire, Local chemistry and growth of single corrosion pits in aluminum, Journal of the
Electrochemical Society 137(10) (1990) 3010.

[60] H. Groult, A. Barhoun, H. El Ghallali, S. Borensztjan, F. Lantelme, Study of the electrochemical reduction
of Zr4+ ions in molten alkali fluorides, Journal of the electrochemical society 155(2) (2007) E19.

[61] M. Sosna, G. Denuault, R.W. Pascal, R.D. Prien, M. Mowlem, Development of a reliable microelectrode
dissolved oxygen sensor, Sensors and Actuators B: Chemical 123(1) (2007) 344-351.

[62] T.H. Osterheld, M.D. Allendorf, C.F. Melius, Unimolecular decomposition of methyltrichlorosilane:
RRKM calculations, The Journal of Physical Chemistry 98(28) (1994) 6995-7003.

[63] Y. Liu, Y. Yang, C. Zhang, T. Zhang, B. Yu, G. Meng, Y. Shao, F. Wang, L. Liu, Protection of AA5083
by a zirconium-based conversion coating, Journal of The Electrochemical Society 163(9) (2016) C576.

[64] G. Šekularac, J. Kovač, I. Milošev, Prolonged protection, by zirconium conversion coatings, of AlSi7Mg0.
3 aluminium alloy in chloride solution, Corrosion Sci. 169 (2020) 108615.

25
[65] Y.-S. Kim, J.G. Park, B.-S. An, Y.H. Lee, C.-W. Yang, J.-G. Kim, Investigation of zirconium effect on the
corrosion resistance of aluminum alloy using electrochemical methods and numerical simulation in an acidified
synthetic sea salt solution, Materials 11(10) (2018) 1982.

[66] C. Baes, R. Mesmer, The Hydrolysis of Cations, John Wiley and Sons, New York, NY, 1976.

[67] E. Curti, C. Degueldre, Solubility and hydrolysis of Zr oxides: a review and supplemental data, Radiochim.
Acta 90(9-11) (2002) 801-804.

[68] M. Pourbaix, Atlas of electrochemical equilibria in aqueous solution, NACE 307 (1974).

[69] L.P. Holmes, D.L. Cole, E.M. Eyring, Kinetics of aluminum ion hydrolysis in dilute solutions, The Journal
of Physical Chemistry 72(1) (1968) 301-304.

[70] T.L. Brown, H.E. LeMay Jr, B.E. Bursten, Chemical thermodynamics, Chemical Thermodynamics of
Zirconium. NEA OECD, Elsevier, ISBN-13 (2005) 978-0.

[71] T. Nguyen, R. Foley, The chemical nature of aluminum corrosion: II. The initial dissolution step, Journal
of the Electrochemical Society 129(1) (1982) 27.

[72] R. Turner, G. Ross, Conditions in solution during the formation of gibbsite in dilute Al salt solutions. 4.
Effect of Cl concentration and temperature and a proposed mechanism for gibbsite formation, Canadian Journal
of Chemistry 48(5) (1970) 723-729.

[73] M. Eigen, L. De Maeyer, Untersuchungen über die Kinetik der Neutralisation. I, Zeitschrift für
Elektrochemie, Berichte der Bunsengesellschaft für physikalische Chemie 59(10) (1955) 986-993.

[74] M. Starostin, G. Shter, G. Grader, Corrosion of aluminum alloys Al 6061 and Al 2024 in ammonium
nitrate‐urea solution, Materials and Corrosion 67(4) (2016) 387-395.

[75] W. Ferrando, Zirconium as a structural material for naval systems, Naval surface weapons center silver
spring MD, 1985.

[76] T. Yau, R. Webster, Corrosion of zirconium and hafnium, Metals handbook ninth edition. Volume 13.
Corrosion1987.

[77] This data are provided by our experimental collaborator in the Center for Plasma Materials Interactions
(CPMPI) under Professor David N Ruzic at the University of Illinois at Urbana-Champaign, which will be
published in the near future. .

[78] B. Zaid, D. Saidi, A. Benzaid, S. Hadji, Effects of pH and chloride concentration on pitting corrosion of
AA6061 aluminum alloy, Corrosion Sci. 50(7) (2008) 1841-1847.

[79] Y. Guo, G.S. Frankel, Characterization of trivalent chromium process coating on AA2024-T3, Surface and
Coatings Technology 206(19-20) (2012) 3895-3902.

[80] S. Pride, J. Scully, J. Hudson, Metastable pitting of aluminum and criteria for the transition to stable pit
growth, Journal of the Electrochemical Society 141(11) (1994) 3028.

[81] G. Frankel, J. Scully, C. Jahnes, Repassivation of pits in aluminum thin films, Journal of the
Electrochemical Society 143(6) (1996) 1834.

[82] V. Branzoi, F. Golgovici, F. Branzoi, Aluminium corrosion in hydrochloric acid solutions and the effect of
some organic inhibitors, Mater. Chem. Phys. 78(1) (2003) 122-131.

26
[83] D.W. Buzza, R.C. Alkire, Growth of corrosion pits on pure aluminum in 1M NaCl, Journal of the
Electrochemical Society 142(4) (1995) 1104.

27

You might also like