Corrosion can lead to mechanical damage near the material surface and reduce the material’s strength. It is essential to understand and simulate corrosion damage evolution for predicting the residual service life of engineering structure, reliability analysis, and corrosion-resistant design of materials. Several major novel corrosion simulation methods in the past 10 y are mainly introduced: cellular automata method, finite element method, phase field model, and peridynamics model. The computational modeling of localized corrosion is discussed and the advantages and disadvantages are compared. Finally, some difficulties in practical engineering applications such as dynamic interface tracking, multiscale and multiphysical field corrosion simulation, and standardization of corrosion simulation are proposed, and the future investigation direction is explored. With the rapid development of software science and computer technology, the operation speed and accuracy of numerical simulation will be greatly improved. The application advantages of numerical simulation in the field of corrosion will be more prominent.

Most metals exist in natural environments in the form of compounds, such as oxides, carbonates, and sulfides. Through smelting technology, people consume energy to extract metal materials with excellent performance from these compounds, and people utilize metallic materials in all aspects of life. The extracted metal materials are comparatively unstable with higher energy than their natural compound. Corrosion is a process in which a chemical reaction occurs between a metal and its environment (usually the natural environment) and slowly returns to its stable state from an unstable state. In this process, the structural materials may deteriorate, which can cause severe harm to society’s development.

First of all, corrosion will directly lead to a decline in the life of metal materials, resulting in tremendous economic losses. The annual economic loss due to corrosion in China reached 2.1 trillion yuan (2014), accounting for about 3.34% of the gross domestic product (GDP) of that year.1  The annual economic loss caused by corrosion in the world is about 2.5 trillion US dollars (2013).2  The impact of corrosion is not just measurable as a share of GDP. Corrosion causes enormous waste of resources and economic losses, pollutes the environment seriously, and even leads to fire and explosion, resulting in heavy casualties. A typical instance is the leakage of radioactive water from the Fukushima nuclear power plant in 2013 due to corrosion.3  It will take 40 y to completely repair the environmental contamination caused by the incident. On November 22, 2013, the Donghuang oil pipeline of Sinopec in Qingdao leaked and exploded, resulting in 62 deaths, 136 injuries, and a direct economic loss of nearly 752 million yuan. The subsequent investigation found that the direct cause of the accident was the corrosion at the junction of the oil pipeline and the drainage canal, and the pipeline was thinned and cracked, leading to crude oil leakage.4  On August 1, 2014, a gas explosion occurred in Kaohsiung, Taiwan due to the corrosion of old pipelines.5  In addition, corrosion will also affect the appearance and use of civil architecture, scenic spots, and daily necessities, and poses a serious threat to human health and travel safety. Corrosion also has a serious impact on cultural relics, historical sites, and national security. With the rapid development of China’s infrastructure, especially the continuous advancement of the national marine strategy, China has paid increasing attention to the problem of material corrosion. In fact, corrosion has become a stumbling block for the implementation of China’s “One Belt, One Road” strategy, and the regions along the “21st Century Maritime Silk Road” such as Southeast Asia and the coast of the Indian Ocean are facing serious corrosion problems. In a recently published report, Wu summarized and analyzed in detail the intense impact of corrosion on China’s economy, environment, and national defense.6  In May 2015, the “Made in China 2025” issued by the State Council of China clearly pointed out 10 key development areas, many of which involved corrosion research, such as new materials, marine engineering equipment and high-tech ships, energy-saving and environmental protection equipment and resource recycling, etc.7 

Benefiting from the rapid development of various high-precision material characterization methods, a large number of experimental observations have been made on the evolution of various corrosion damages at different scales. Due to the complexity of the corrosion process, the current cognizance of the corrosion evolution process and the simulation and prediction of corrosion damage are still challenging. The National Research Council of the United States published “Research Opportunities in the Field of Corrosion Science and Engineering” in 2011. The book summarizes four major scientific research opportunities related to corrosion that can also be said to be four major challenges:8  (1) development of low-cost and environmentally friendly corrosion-resistant materials and coatings; (2) high-fidelity modeling of corrosion evolution in service environments; (3) quantitatively link accelerated corrosion tests under laboratory conditions with long-term corrosion observations in service environments; and (4) accurately predict the remaining service life. Among the four challenges, items (2) and (4) are in connection with corrosion modeling directly. The establishment of corrosion models will also greatly accelerate the development of corrosion-resistant materials and coatings. Additionally, they will provide insight into the viability and advantages of various corrosion mitigation methods.

In this paper, several modeling methods of localized corrosion in the last 10 y are introduced, including cellular automata (CA), finite element method (FEM), phase field (PF) model, and peridynamics (PD) model. The strengths and weaknesses of each method are analyzed in this present review. Additionally, several difficulties that need to be solved are pointed out, providing guidance for further research on localized corrosion simulation.

The establishment of corrosion models is required to combine the understanding of the corrosion mechanism, the knowledge of the materials in the specific application environment, and the calibration of the model parameters with the corrosion test data. The purpose of corrosion modeling is to predict the corrosion evolution process of materials in a specific application environment and the resulting changes in other properties and to reasonably estimate the expected corrosion behavior of materials. Over the past two decades, researchers’ progress in corrosion models has made the prediction of corrosion damage evolution a realistic possibility in the near future. In recent years, the experimental characterization of corrosion at different scales has deepened the understanding of corrosion mechanisms. At the same time, with the further enhancement of computing power, modeling research, and numerical prediction of corrosion have received more and more attention and development. The following introduces some important localized corrosion models and their corresponding results.

Cellular Automata Model

CA, proposed by mathematicians Stanislaw M. Ulam and John von Neumann, is a mathematical model to describe complex phenomena in nature.9  CA technique has been used in simulation studies in various fields, including public transport systems,10  the spread of forest fires,11  and the growth of cells.12  Because the CA model is on the basis of the localized reaction rules of the system, it can reflect the influence of complex physical and chemical systems at different scales, including the molecular or atomic scale, and the properties of these systems can be qualitatively described in the macroscopic scale. It is more intuitive and convenient to model the corrosion evolution process using this technology. In recent years, a growing number of scholars have used CA to conduct microscopic modeling of corrosion and realize the numerical simulation of the metal corrosion process. The application of CA in the corrosion field has experienced three stages: simple CA model, complex CA model, and three-dimensional (3D) CA model. We first introduce the mathematical basis of CA, and then summarize the application and research status of CA method in the field of corrosion.

Mathematical Basis of Cellular Automata Model

The mathematical definition of CA is:13  Σ is the set of cellular states, K = |Σ| is the number of states, N is the scale of CA, i = 0, 1, … , N − 1 is the serial number of each cell, the state of the cell i at the moment t is marked as , is neighbor of cell i, represents the scale of the neighbor cells, is renewal rule of cellular, is the configuration of CA. Then CA {ΣN, φ} can be regarded as a mapping on configuration space, that is φ: ΣN → ΣN, where and the rules of CA are also often represented by f. CA is mainly composed of elements such as cells, cell space, cell neighbor, and cell rules. In the CA model, each cell has a specific state at a given time, and the evolutions of the cells and the updating of the internal states follow certain rules. A cell is the most basic unit of the CA model. As a state point, a cell is discretely distributed in one-dimensional or multidimensional space. Each cell is in a specific state at every moment, which can be binary {0,1} or discrete set form {S0, S1, … , Sn}. The cell space is a collection of distribution points of cells on a spatial grid. In one-dimensional space, there is only one way to divide cells. However, in high-dimensional space, there can be multiple ways to divide cells. In common two-dimensional (2D) space, the main types of cell division are triangular cells, quadrangular cells, and hexagonal cells.

When using a CA model for simulation, the simulated objects are limited, and the selected model space is bounded. This means that the evolution rules at the boundary cannot be defined based on the processing methods used in other regions of the cellular space. Therefore, special processing must be done at the boundary of the cell space. One possible solution is to extend the neighbor of the cell at the boundary. The most common boundary conditions include periodic boundary conditions, fixed boundary conditions, adiabatic boundary conditions, and mapped boundary conditions.13  During the evolution of a CA model, certain rules must be followed, and these rules are based on the local range of the cellular space. This means that the state of a cell at the next moment is determined by the state of the cell itself and its neighbors at that moment. Therefore, we must specify which cells belong to the neighbor by defining the neighbor rules of the cell. In general, the neighborhood of a one-dimensional CA model is determined by a radius. For 2D and high-dimensional CA, the neighbors are related to the cellular spatial structure of the model. For example, in the case of a two-dimensional quadrilateral space structure, the most commonly used cellular neighbors include the Von Neumann type, Moore type, and extended Moore type.13  Cellular rules are the core aspect of the CA model that must be formulated according to the physical nature of the research object. The evolution rules are applicable to every cell in the cellular space, making them global in nature.

Simple Cellular Automata Model

In 1993, Meakin, et al.,14  established a 2D model of passivation and depassivation by pitting based on the study of pitting morphology, as shown in Figure 1. The model consists of four types of cells: corrosive and noncorrosive solution cells, and active and inactive metal cells. Corrosive solution cells could freely migrate, and when they came into contact with active metal cells, the metal cells dissolved. This model describes corrosion processes in which the diffusion of corrosive electrolytes or corrosion products is impeded, and although relatively simple, it shows the pitting kinetics as well as current fluctuations. Taleb, et al.,15-16  used the CA method to simulate and study the phenomenon of insoluble corrosion products generated in the metal corrosion process depositing on the metal surface to form a film. They simulated the processes involved in the metal corrosion process, such as the early metal corrosion, the redistribution of the intermediate corrosion products penetrating the previously generated film, and the deposition of corrosion products on the film surface. Li, et al.,17-18  also used the CA method to simulate the metastable pitting evolution process. The CA model includes several rules of evolution that govern electrochemical reactions, solution reactions, and diffusion. The effects of cationic diffusion, the degree of fracture of protective films, the radius of pits, and the presence of salt films on the morphology of pits were investigated. The simulated kinetic results of metastable pit growth were compared to experimental data cited in Ren, et al.,19  defined some simple localized rules of evolution that describe the basic physical and chemical processes. By characterizing the CA, temperature, electrolyte concentration, and dissolution probability for metal exposed to corrosive environments in aircraft structures, the crevice corrosion morphology was obtained through simulation with good consistency with the experiment. The simulation shows that the change of dissolution current with time follows the power function relationship. Lishchuk, et al.,20  built a CA model for the intergranular corrosion (IGC) of AA2024 (UNS A92024(1)) aluminum alloy. Based on a series of evolutionary rules and random walk processes, the model defined corrosion probabilities of grain boundaries and surface layers. The simulation parameters were calibrated through the differential evolution algorithm to adapt to specific experiments. The model can reproduce and predict the growth of IGC of AA2024 aluminum alloy.
FIGURE 1.

The corrosive solution dissolves the active metal and the metal passivates and depassivates with certain probability14  (draw based on Meakin, et al.14 ).

FIGURE 1.

The corrosive solution dissolves the active metal and the metal passivates and depassivates with certain probability14  (draw based on Meakin, et al.14 ).

Close modal

Complex Cellular Automata Model

The simple CA model simplifies the corrosion model by converting the electrochemical behavior of the corrosion process into chemical behavior. However, the electrochemical behavior of metal corrosion is the main behavior, and the anode and cathode reactions occur in different positions. In order to make the corrosion model closer to realistic situations, Taleb, et al.,21  adopted a CA model to define the anode and cathode reactions in different positions as shown in Figure 2. The diffusion of reaction products was simulated as the random walking of cells, and the evolution process of the size, morphology, and chemical composition of the pits over time was examined. The variation of the pit shape with the diffusion rate was further explored.
FIGURE 2.

Model of separation of anode and cathode reaction sites21  (draw based on Taleb, et al.21 ).

FIGURE 2.

Model of separation of anode and cathode reaction sites21  (draw based on Taleb, et al.21 ).

Close modal

Stafiej, et al.,22  established a CA model of the metal corrosion-passivation process based on the anode and cathode reactions of space separation and simulated the evolution process of corrosion pit morphology when the protective layer of the metal surface was locally destroyed or the protective layer was not covered on the metal surface. Reis, et al.,23  defined metal cells in active and passivated states, acidic, neutral, and alkaline solution cells, and then described the formation and dissolution process of the passivated layer on the metal surface. Pidaparti, et al.,24-26  established two stages of pitting initiation and propagation to simulate multipit corrosion. In the initiation stage, it is assumed that all cells do not corrode, and the chaos degree of cells increases with each time step. When the chaos degree reaches a certain value, the cells become corrosive cells and enter the pitting propagation stage, with corresponding rules for cellular transformation. The changes in the cellular state depend on the surrounding cellular and the pH, dissolution parameters, temperature, and potential difference between the metal and the solution. Taleb, et al.,27  and Vautrin-Ul, et al.,28-29  established probabilistic CA models based on electrochemical and diffusion mechanisms and simulated the generation of the “island” phenomenon in the corrosion process by replacing the actual electrochemical reaction with probabilistic events in the CA model. This deviation between the electrochemical reaction and Faraday’s law in real solutions is explained. Saunier, et al.,30-31  simulated the growth process of the corrosion product layer at the solid/solution interface in a metal corrosion system with the diffusion process as the control factor and further explored the diffusion behavior of cations in the corrosion product layer. The thickness of the corrosion product layer, the concentration distribution of the cation, and the intensity and position of the diffusion front were compared under different reaction rates. Fatoba, et al.,32  used CA to describe the corrosion components and utilized finite element analysis (FEA) with a stress concentration effect to simulate the damage caused by the interaction between corrosion and mechanics, which occurs in the initial environment-induced cracking. This provides a physical framework for modeling damage evolution under the combination of the corrosive environment and mechanical load, which is applicable to stress corrosion cracking (SCC) and corrosion fatigue mechanisms.

Three-Dimensional Cellular Automata Model

To present the corrosion microstructure and initiation process more intuitively, researchers have used the CA method to establish a 3D model to simulate the corrosion process of metal. Van der Weeën, et al.,33  developed a 3D model that considered the role of corrosive chloride ions in solution. The model focused on mass transport, IR drop, pit initiation, metal dissolution, and cathodic protection, but ignored special corrosion behaviors such as passivation and bimetallic corrosion. Pérez-Brokate, et al.,34-35  established a 3D CA model based on the electrochemical half reaction of spatial separation in the corrosion process, ion diffusion, acid-base neutralization of the solution, and passivation characteristics of the oxide layer, and studied the influence of pitting morphology and solution acidity on the pitting process. Zenkri, et al.,36  explored the evolution process of metal corrosion damage under three different initial conditions using a 3D random CA model. The model considered the local acidification of the solution, and the research revealed that the autonomous spatial separation of the anode and cathode regions is not only related to the metal matrix and surface passivation film but also to the local acidification of the solution. Compared with 2D simulation, CA simulation in 3D space can describe the morphology of pits and the evolution process of corrosion damage more intuitively, but the amount of calculation considerably increases. Xuefeng, et al.,37  presented a 3D CA model that efficiently simulates multipit corrosion, which considers the breakdown of the passive layer that plays a crucial role in the corrosion process. The results can contribute to simulating the corrosion process and interactions of multipit corrosion covered by a protective coating.

Finite Element Method

The FEM is a popular numerical technique utilized to solve problems in engineering and mathematical physics that involve behaviors described by differential equations. These differential equations can describe a wide range of physical phenomena, from electrical and mechanical systems to chemical and fluid flow problems. The FEM uses different types of discretization methods to divide the domain of interest into smaller elements and approximate the differential equations with numerical model equations to provide numerical solutions for a given set of boundary and initial conditions. It is important to note that the selection of the appropriate governing equation is a critical step in any FEM-based computational study. In electrochemical and corrosion systems, there are two main modeling approaches using the FEM: those that use the Nernst-Planck equation (N-P) and those that use the Laplace equation.38-40  In this context, we will first introduce the mathematical basis of these two approaches and then discuss their application to local corrosion simulation.

Governing Equation

The Nernst-Planck equation is the most complete means to describe the materials balance of each charged species (j) in an electrochemical system. It is formulated as a summation of diffusion, migration, convection, and homogeneous reaction terms:
formula
where Nj is the mass flux, Cj is the concentration, zj is the number of charges, Dj is the diffusivity, φl is the electrolytic potential, F is the Faraday’s constant, v is the fluid velocity, and Rj is the homogeneous production of species j. If the number of charged species is n, a system of n equations in the form of Equation (1) can be obtained. Importantly, it should be noted that there are (n + 1) variables in this system of equations: all of the Cj (j from 1 to n) and φl. Thus, in order to find the solution of this system, an (n + 1)th equation is needed, which in electrochemical systems is most often implemented by the application of the electroneutrality assumption in the electrolyte domain:41 
formula

This assumption simply states that any bulk material (the electrolyte in this case) cannot sustain a net charge. Application of this assumption provides the (n +1)th equation needed and, as a result, the electrolyte potential as well as the concentration of each charged species are obtained.

As for the current density distribution, in the electrolyte domain, the total current density il can be expressed as:
formula
The last term (convection) in Equation (3) is canceled out due to the electroneutrality assumption, then Equation (3) is reduced to:
formula
By combining conservation of charge (Equation [5]) with Equation (4) the electrolyte current density il can be obtained eventually.
formula

Using the N-P equation provides complete transient descriptions of the distributions of potential, current density, and concentrations of important species. However, this approach is computationally expensive due to its modeling complexity and execution time. It is important to consider the difference in time scales when modeling both fast steps (such as Faradaic reaction at the interface and ion migration under electrostatic force, etc.) and slow steps (such as species diffusion under a concentration gradient). These varying time scales, along with highly nonlinear electrochemical kinetics as boundary conditions, require highly refined spatial meshing at the relevant boundaries and smaller time steps during the transient study. Furthermore, N-P equation-based modeling generally assumes electroneutrality, and the selection of reference ions may result in discrepancies in the modeling results.

An alternative approach is to use the Laplace equation as the governing equation. In this approach, it is assumed that the solution is well mixed and there is no bulk motion of the fluid such that the first and last terms in Equation (3) are canceled out, as a result, Equation (3) is reduced to Equation (6):
formula
where conductivity . By applying charge conservation (Equation [5]), one can obtain the Laplace Equation:42 
formula
instead of solving for a transient equation like Equation (1), the use of the Laplace equation approach depends highly on the electrolyte characteristics (primarily conductivity) as well as the boundary conditions (electrochemical kinetics), as electrochemical kinetics are always functions of position and/or some other external variables (e.g., electrode surface property, solution chemistry, electrolyte layer thickness, etc.). However, when a physical system involves important contributions from diffusion and convection, the Laplace equation-based approach will fail to accurately capture the situation.

Both approaches use boundary conditions to describe the electrochemical kinetics of reactions, which are generally highly nonlinear. These nonlinear boundary conditions can be described mathematically in some cases, such as via the Butler-Volmer equation for cathodic/anodic reactions under either charge-transfer control, mass-transfer control, or mixed control. However, in many cases, the polarization behavior measured does not follow any prescribed law, such as in a system containing an active-passive transition. In these and other cases, a precise description of the kinetics is required via numerical fits to the polarization behavior for both anodic and cathodic reactions. Regardless of the means by which the electrochemical kinetics are described, they serve as the most critical aspect of the problem statement.

Pitting

Sharland, et al.,43-44  proposed several mathematical approximations to solve the steady-state mass-transport equations governing the concentration of species and potential of the electrode in crevice corrosion or pitting. In their study, they developed a mechanistic formula for the propagation stage of pitting or crevice corrosion based on the representation of physical mechanisms controlling the process. The model considered the time evolution of solution chemistry and electrochemistry within an active cavity, as shown in Figure 3. To solve the complex mass-conservation equations describing the system, they utilized established a foundation for FEM of corrosion damage.
FIGURE 3.

Schematic view of processes involved in the pit/crevice propagation.44 

FIGURE 3.

Schematic view of processes involved in the pit/crevice propagation.44 

Close modal

Engelhardt, et al.,45  presented a simplified method to estimate pit propagation rates in steel in dilute sodium chloride solutions. They assumed that if the rate of an electrode reaction depends only on the potential, the pit growth rate depends only on the concentration of those species that determine the potential distribution near the metal within the cavity. They compared analytical expressions for calculating the propagation rates of cylindrical and hemispherical pits and explored the influence of aggressive anions on the propagation rate using experimental data. Laycock, et al.,46  proposed an accurate model, incorporating a minimum cation concentration for active dissolution, to explore the observation of porous covers made of metal and oxide over the pit mouth and the resultant pit growth morphology for stainless steel. The lace-like porous structure of the pit cover in stainless steel arises from the strong dependence of the dissolution rate on dissolved metal ion concentration. Later, they developed a mathematical model to simulate the propagation of pitting in stainless steel.42  The FEM was used to simulate 2D pit growth, which showed good agreement with experimental results concerning pit morphology and propagation stability. In situ synchrotron x-ray radiography was used to verify the simulation results.47 

Galvele and Gravano48-49  built a pitting model based on the assumption that metal ions hydrolyze inside the pits and that corrosion products are transported by diffusion. They found that the dominant reason for passivity breakdown at the initial stages of pit growth is localized acidification due to metal ion hydrolysis. Malki, et al.,50  developed a corrosion pitting model for the polarization mode and made more detailed calculations to compare the behavior of ferritic and austenitic stainless steel. They obtained the formulation of critical conditions for pit stabilization in terms of pit geometry and applied potential. Amri, et al.,51  proposed a numerical model of the steady-state behavior of single pitting, which provided an interesting view of the behavior of the chemistry and electrochemistry inside the active pit.

The influence of mass transport can also play an important role in the development of pitting. Alkire, et al.,52-53  investigated the mass transfer between the interior surface of small rectangular cavities and the solution flowing past the cavity opening using both theoretical and experimental techniques. This approach can be a powerful tool for better understanding phenomena involving cavities, such as polishing, corrosion, rinsing, and roughness. Recently, Srinivasan, et al.,54-55  examined the flux from a one-dimensional artificial pit electrode corroding under a salt film using experimental and modeling techniques. Their work provided new insight into the origin of the dependence of the measured repassivation potential with pitting depth and contributed toward a quantitative framework relating the various critical influences controlling pitting.

The passive film formed on a metal surface can have a direct impact on the size and life of the pitting. The formation of the passive film depends on the critical potential or current density criterion. Marshall, et al.,56  adopted the FEM to simulate and compare the critical pit radius of stainless steel predicted by thermodynamic and kinetic repassivation criteria. They utilized experimental electrochemical boundary conditions to obtain the kinetics of active pits and incorporated geometric and environmental parameters, such as the shape and size of the pit, solution concentration, and thickness of the water layer, to evaluate their impact on the repassivation criteria of pitting. The results showed that parameters, such as lower chloride concentration, smaller cathode diameter, and thinner water layers, lowered the stability of a pit and increased the probability of repassivation, independent of pit shape.

Crevice Corrosion

Xu and Pickering57  developed a computational model to calculate the distribution of potential and current in the electrolyte phase on the electrode surface for a system. A boundary variation and a trial and error technique were incorporated into the FEM to evaluate the values of the critical distance into a crevice for pure iron in buffered ammoniacal (pH = 9.7) and acetic acid (pH = 4.6) solutions. The results were in good agreement with the available literature values.58  The presented model can also be used in the prediction of IR-induced crevice corrosion under open-circuit conditions and in the prediction of precracked samples, critical temperature of SCC, design of anodic protection systems, and design of some corrosion-resistant alloys and coatings.

Vankeerberghen59  explored the effect of the critical characteristic dimension on crevice corrosion. This critical dimension determines whether crevice corrosion occurs or not. Lee, et al.,60  proposed a controlling scaling factor to relate the corrosion depth and crevice geometry and explored the impact of crevice geometry on the susceptibility of a crevice. Percheron, et al.,61  measured the pH gradients over the electrode by total internal reflection fluorescence microscopy during cathodic polarization under potentiostatic and galvanostatic conditions to validate the calculated pH gradients within an occluded cell. The results were consistent with those from a pseudo-2D computational model of transport. Malki, et al.,62  investigated the crevice corrosion of ferritic stainless steel in chlorine-containing media by combining experimental data with calculation. One unique aspect of their study was the use of a three-electrode device to observe the current and potential during the initiation and propagation stages. Figure 4 shows the potential distribution inside the crevice. The technique used in this study can highlight several physical and chemical effects, such as chloride concentration, pH, temperature, and the effect of surface pretreatment. One limitation of the model is that there is no description of crevice chemistry in the propagation stage, which made it difficult to better monitor what happens when corrosion extends beyond the restricted area.
FIGURE 4.

Potential distributions inside the crevice calculated as a function of the counter electrode area, note that the ohmic drop remains almost constant (∼120 mV).62 

FIGURE 4.

Potential distributions inside the crevice calculated as a function of the counter electrode area, note that the ohmic drop remains almost constant (∼120 mV).62 

Close modal

On the basis of crevice corrosion simulation of single stainless steel and single titanium, Ding, et al.,63  conducted crevice corrosion simulation of 304 stainless steel and titanium overlapping, which is based on a multiphysical model. This work provides a lot of favorable information for the study of crevice corrosion mechanism that is difficult to obtain in experiments, which performs a significant function in the prevention and control of crevice corrosion.

Galvanic Corrosion

Galvanic corrosion refers to a kind of electrochemical corrosion, sometimes also known as contact corrosion, which occurs when metal components form corrosive batteries in an electrolyte solution and generate galvanic current, which speeds up the dissolution of metals with lower corrosion potential and slows down the dissolution of metals with higher corrosion potential. A new numerical model has been developed to capture the evolution of cross-sectional microstructure by incorporating a moving mesh technique into COMSOL MultiPhysics.64  The model used an arbitrary Lagrangian Eulerian (ALE) method to determine the moving boundary of the corroding phase (α phase) and velocity of the interface. It is found that a configuration with a continuous network of β phase around the α phase provides the best corrosion resistance. The corrosion rate is found to be higher during the initial stages due to increase in the β phase fraction. However, corrosion is halted after the α phase is preferentially dissolved in the electrolyte solution, exposing the β phase which is electrochemically less active. This trend is obtained from both the numerical model and scanning vibrating electrode technique experiments.

This method is commonly used in computational mechanics and provides a powerful tool to predict the evolution of the microstructure during the corrosion process. Xiao and Chaudhuri65  presented a new corrosion model based on FEM, specifically tailored for pitting corrosion of aluminum alloys. This model is different from existing ones in that it has strong predictive power and high generality. The corrosion rate as well as pitting stability can be quantitatively estimated for a wide range of systems, including complex pitting morphology, heterogeneous alloy microstructure, and versatile solution chemistry. The model takes the corrosive environment as input for the bottom layer, where the flux of reactive species across the interface and the changes of the interface location depend on surface reactions. As a consequence, electrolyte domain models with boundary location and condition can be provided (as shown in Figure 5). Yin, et al.,66-67  and Wang, et al.,68  studied the deposition of Al(OH)3 on the corrosion pit and its blocking impact on the corrosion of the Al matrix during pitting morphology development in a single-intermetallic compounds (IMC) (Al3Fe)/AA7075-matrix microgalvanic coupling. A 2D FEM model was proposed to describe the chemical changes in front of the ultramicroelectrode positioned with a scanning electrochemical microscope device, which confirmed that the initiation of the dissolution of aluminum is controlled by chemical reaction.69  These studies demonstrate the potential of FEM in understanding the mechanism of corrosion and developing effective strategies to prevent and control it.
FIGURE 5.

Model system for pitting corrosion: (a) system domain decomposition and (b) model framework.65 

FIGURE 5.

Model system for pitting corrosion: (a) system domain decomposition and (b) model framework.65 

Close modal

In addition to the study of microgalvanic corrosion, the FEM can also contribute to the overall corrosion distribution of many engineering structures. Deshpande39,70  presented a Laplace equation-based numerical model to predict the corrosion rate of a galvanic couple, which was capable of tracking the moving boundary of the corroding components. The corrosion rates simulated are in good agreement with those tested from the experimental data. In addition, the FEM model can provide a view on the performance of sacrificial metallic coatings which protect the underlying materials substrate.71-76 

Stress Corrosion Cracking

SCC refers to corrosion damage that can promote crack initiation and propagation through continuous dissolution in the anode area of the crack tip under a corrosive environment and tensile stress. Currently, the simulation of stress corrosion processes using FEM is not widely used due to the environmental impact.77-79  Some studies instead concentrate on analyzing the pitting or stress distribution of the entire structure,80-81  while others focus on examining the correlation between various influencing factors such as chloride concentration, temperature, pH, cation species, and SCC.82-83 

Turnbull, et al.,80  used FEM to investigate the stress-strain distribution near the pit of cylindrical samples. The study presented the maximum principal stress and strain distribution of the pit as shown in Figure 6. It reveals that the pit’s bottom experiences the maximum stress, and the shoulder of the pit experiences the highest strain. As the size of the pit increases, the plastic strain decreases. The SCC behavior of SS304 was analyzed in a study by utilizing the Lemaitre damage model.82  The proposed model was modified by incorporating the effect of temperature and chloride concentration as damage parameters. The outcomes indicate that increasing temperature and chloride concentration can augment the number of cracks, and also reduce the time needed to initiate them. Moreover, compared to the chloride ion concentration within the test range, temperature has a greater impact on the onset of cracking during SCC.
FIGURE 6.

Maximum principal stress and maximum principal strain in cylindrical samples viewed in cross section: (a) 100 μm hemispherical pit and (b) 500 μm U-shaped pit. The applied stress was 90% of σ0.2.80 

FIGURE 6.

Maximum principal stress and maximum principal strain in cylindrical samples viewed in cross section: (a) 100 μm hemispherical pit and (b) 500 μm U-shaped pit. The applied stress was 90% of σ0.2.80 

Close modal

Phase-Field Model

PF is another important method for simulating moveable interface moveable problems. In the PF method, the free energy of the system is assumed to be a function of the field variables and their gradients of each material point. The derivation of the governing equation of the PF variables follows thermodynamics, that is, the dynamic simulated process follows the minimization of the system’s free energy. The position of the phase interface is implicitly solved by the control equation, and it is not necessary to apply boundary conditions on the interface, which also contributes to the application of this method to the morphological evolution problem. The PF model has been used for many years to simulate the evolution of interfaces between phases in various problems such as solidification,84  microstructural evolution,85  phase transitions in ferroelectric86  and ferromagnetic87  materials, etc. PF models have been recently used in simulating corrosion through modeling the evolution of the metal/electrolyte interface.88  We first introduce the mathematical basis of PF model, and then summarize the application and research status of the PF model in the field of corrosion.

Governing Equations

A PF approach for modeling corrosion allows for the implicit approximation of interface evolution by evaluating the distribution of an auxiliary field variable φ over the entire system.88  In contrast to sharp interface models (SIMs), which rely on explicit interfaces, a PF model assumes constant values in the bulk of each phase with φ = 0 for the electrolyte and φ = 1 for the metal. The PF variable φ varies continuously across the diffuse interface with a finite thickness l, allowing for a more accurate representation of the pitting boundary. Thus, in a PF model, the auxiliary PF variable φ is utilized to track pitting interface evolution, with a thickness of l representing the diffuse pitting boundary.

In a PF model for corrosion, the free energy of the system is evaluated by considering the distributions of the normalized molar concentration c′ and the PF variable φ. The driving force for any phase transformation, including corrosion, is the reduction of the system’s free energy. For a PF model, the free energy consists of both the homogeneous bulk energy and the interface energy 88 
formula
where f(c′, φ) is the local free energy density, which is a function of c′ and φ, while fint is the excessive energy associated with the diffuse interface. The excessive interface energy arises from the inhomogeneity within the interface region, which can be written as a function of the field variable gradients 88 
formula
The gradient energy coefficients associated with the concentration and PFs in Equation (9) are αc and αφ, respectively. To ensure that the free energy of the system reduces during the corrosion process, the governing equations for the PF are derived by minimizing the free energy functional via variational differentiation88 
formula
formula

where L is the interface kinetics parameter and M is the diffusion mobility for mass transport. The two equations are known as the Allen-Cahn and Cahn-Hilliard equations, respectively. It is worth noting that in practice, only one of the gradient terms (∇c′ or ∇φ) is needed to approximate the energy contribution from the diffuse interface. For simplicity, the concentration gradient energy coefficient αc is assumed to be 0 in this case. Although Equations (10) and (11) represent the general form of a PF model, the definition of the local free energy density f(c′, φ) is what differentiates various models. By using an appropriate local free energy density f(c′, φ) with the proper assumptions, one can simulate the propagation of pitting corrosion by solving Equations (10) and (11), respectively.

Pitting

Mai, et al.,88  presented a PF model to simulate the activation-controlled and diffusion-controlled pitting phenomena in metallic materials. They demonstrated that the proposed model can reproduce different portions of the polarization curve associated with the activation-controlled, diffusion-controlled, and mixed-controlled corrosion kinetics by calibrating the interface kinetics parameter. In addition, the model has significant application in the simulation of the electrolytic polishing process, interactions between multiple growing pits and pitting corrosion in a metal matrix composite and polycrystalline stainless steel. Figure 7 shows the pitting of SiC particle-reinforced aluminum composite. Other PF corrosion models89-91  are generally similar to the above models but may differ in some details from the one presented above.
FIGURE 7.

Examples for pitting corrosion simulations with PF models: (a) microstructure of the composite specimen before corrosion, (b) through (d) evolution of the corrosion pit morphology at different times.88 

FIGURE 7.

Examples for pitting corrosion simulations with PF models: (a) microstructure of the composite specimen before corrosion, (b) through (d) evolution of the corrosion pit morphology at different times.88 

Close modal

Brewick92  proposed a computational modeling framework to study the effect of crystallographic orientation on corrosion pits by relying on corrosion potentials and elastic constants. However, the model is constrained by assumptions and is currently limited to 2D analysis. In addition, the current model does not contain potential distribution, which also affects the accuracy and fidelity of the model to a certain extent. A 3D PF model was established to study the influence of substrate orientation on the morphology of crystal pits, which shows that the morphology of crystallographic pit changes significantly with the orientation of the matrix, and has the same symmetry as the orientation of the matrix.93  This is the first 3D PF model of crystallographic pits, which is helpful to predict complex shapes of pits, thus promoting the forefront of pitting modeling.

The PF model currently lacks standardized commercial software, leaving researchers to write programs from scratch, requiring a significant amount of computation. As a result, the development of efficient PF calculation methods at both spatial and temporal scales is of utmost importance. Gao, et al.,94  proposed a pitting PF model which adopted spatiotemporal adaptive FEM to solve two main challenges. The first challenge is the need for an intensely small time step size. To solve this problem, they combined the Rosenbrock-Euler exponential integrator with the Crank-Nicolson scheme for time discretization. In addition, the adaptive time stepping formula is derived by using the characteristic that the corrosion interface velocity decreases with time. The second one is the requirement for very fine mesh generation. To solve this problem, they propose a simple and effective adaptive mesh generation strategy, which greatly reduces the computational cost. In future work, this method may be extended to other types of corrosion models.

Crevice Corrosion

Xiao, et al.,95  proposed a quantitative PF to study crevice corrosion. The overpotential through the Butler-Volmer equation was considered as well as the potential distribution and chemical reaction between different kinds of ions in the electrolyte. The system is the same as the system studied by Sharland, et al.43-44  It should be noted that Sharland’s work was on the basis of steady-state corrosion kinetics, while Xiao’s model is time dependent. They considered six kinds of ions and related chemical reactions. In addition, some physicochemical properties related to corrosion under different metal potentials including overpotential, pH, and corrosion rate, were also examined. The 2D and 3D PF corrosion simulation can be realized by adjusting the modeling area of ion species in the quantitative PF model. This investigation provides a sound framework for further study on 2D and 3D local corrosion.

Galvanic Corrosion

A new PF model was proposed to simulate galvanic corrosion and pitting of metal.96  The potential distribution is obtained approximately by the Laplace equation through associating the anode current density with the interfacial dynamics parameters, the potential distribution is obtained approximately by the Laplace equation. In the case that the anode is not polarizable, the diffusion constraint kinetic is considered as the boundary conditions of the cathode. They also demonstrated the application of the model to simulate the galvanic dual corrosion of hybrid joints and aluminum composites under different environmental conditions, which verified the feasibility of incorporating homogeneous chemical reactions and polarization behavior on the anode into the PF model (as shown in Figure 8).
FIGURE 8.

First application problem: (a) schematic of the problem domain and (b) four snapshots of the PF simulation of the evolving corrosion pit morphology (carbon fiber reinforced plastic [CFRP] plate).96 

FIGURE 8.

First application problem: (a) schematic of the problem domain and (b) four snapshots of the PF simulation of the evolving corrosion pit morphology (carbon fiber reinforced plastic [CFRP] plate).96 

Close modal

Intergranular Corrosion

A multi-phase-field (MPF) model was presented to quantitatively predict the IGC kinetics of metal by Ansari, et al.97  The IGC predicted by the 2D MPF model is in good agreement with the experimental data, as shown in Figure 9. The corrosion process becomes transport controlled even at a low applied potential value due to the saturation of metal ions in the corrosion grain boundary area. Through 3D research, the practicable application of the MPF model in complex 3D geometric problems is examined. The simulation results are in good agreement with the experimental data, which also shows that the corrosion rate along the sensitized grain boundaries (SGBs) is significantly increased, while almost no corrosion is observed within the grain boundary. In fact, in the model, the grains are anodic while most previous modeling studies assume them noncorrosive. But the corrosion rate of grains is much lower than that of SGB due to the difference in the properties of metal in the grains and SGB phase.
FIGURE 9.

IGC of AA2024 thin foil in 0.1 M NaCl solution at −0.6 VSCE. (a) SEM images for corrosion penetration in time; (b) modeled microstructure; and (c) MPF model results for the corrosion evolution.97 

FIGURE 9.

IGC of AA2024 thin foil in 0.1 M NaCl solution at −0.6 VSCE. (a) SEM images for corrosion penetration in time; (b) modeled microstructure; and (c) MPF model results for the corrosion evolution.97 

Close modal

Stress Corrosion Cracking

A PF model was presented to simulate the SCC initiated from surface defects and corrosion pits.98  In order to couple the localized corrosion phenomenon with mechanical stress, the film rupture-dissolution-repassivation mechanism was adopted, and the SCC growth rate is considered as a power function of the stress intensity factor by determining the proportionality constant between the velocity of the interface and the interface kinetics parameter. The proposed PF model could validate the availability of the empirical model in reproducing the growth rate of SCC.

Nguyen, et al.,99  developed a PF framework for simulating the fracture growth caused by stress corrosion incorporating the impact of material anisotropy. The classical phase transition model for dissolution of material is coupled with the mechanics in an acceptable way, which provides an efficient tool for exploring the competition between electrochemical and mechanical contributing to fracture. The model can also be extended to heterogeneous materials or polycrystalline systems by adopting appropriate conditions.

In order to establish a new PF model for investigating mechanical-electrochemical corrosion, Allen-Cahn type equation for governing the phase transformation from metal to ions in a liquid electrolyte, NP equations for diffusion, Poisson’s equation for electrostatic field distribution, and mechanical equilibrium equation for elastic energy density evaluation were integrated.100  The numerical results demonstrated that the effect of stress concentration at the tip of a corrosion pit promotes a higher corrosion rate resulting in an accelerated failure of a metallic structure.

Based on the work of Mai and Soghrati,98  a new theoretical and numerical framework was presented to simulate mechanical-assisted corrosion in elastoplastic materials, which is capable of capturing the process of pitting and SCC, as well as pit-to-crack transition.101  FEM and implicit integration method were adopted to solve the electrochemical-mechanical coupling equation. The transition from pits to cracks is the natural result of the model and it lays a foundation for mechanical evaluation of engineering materials and structures. Future developments may involve extending the model to other cracking mechanisms such as cathodic reaction-driven cracking. Lin and Ruan102  proposed the PF model of mechano-chemical-coupled SCC. To identify quantitatively the critical condition for pit-to-crack transition, the relative-rate parameter difference between SCC and mere corrosion is defined so as to identify quantitatively the critical condition for pit-to-crack transition. When the relative-rate parameter is greater than 1, the pit-crack transition occurs, which characterizes the critical condition where stress-induced degradation occurs faster than electrochemical dissolution. Moreover, the relationship between the stress intensity factor and cracking velocity is exponential, indicating that this is an autocatalytic process caused by stress and corrosion acceleration. The results also show that the SCC was more significant when a metallic component was applied stress. In this case, the effect of the initial pit depth or surface damage was critical in the meanwhile.

Peridynamics Model

The appearance of the PD theory has brought new vigor to the simulation of local corrosion. PD theory was first proposed by Silling103  and applied to solid mechanics, which successfully solved the singularity of classical continuum mechanics when solving discontinuity problems. After more than 20 y of development, the PD method has been successfully used to simulate material crack growth104-107  and heat transfer,108-111  and to study multiscale112  and multiphysical field coupling problems.113  With the indepth study of PD theory by researchers, PD was used to simulate local corrosion for the first time in 2015,114  and relatively satisfactory results were obtained, which laid the foundation for PD’s application in corrosion. In this review, the research progress of PD methods in local corrosion (pitting corrosion, crevices corrosion, IGC, galvanic corrosion, and stress cracking) is reviewed.

Governing Equation

Similar to the PF model, in PD formulations the motion of the pit boundary is autonomous and part of the solution to the governing equations. Similar to PF corrosion models, the autonomous moving interface makes the PD formulation a flexible framework for predictive simulations of various types of localized corrosion. The subsurface partial damage in PD models is similar to a transition region at the corrosion front due to the diffuse interface in PF models. Chen and Bobaru developed a model for corrosion damage based on the concept of mechanical damage in PD solid mechanics.114  This approach considers the dissolution and diffusion processes involved in both liquid and solid phases during the corrosion process. Corrosion can cause material damage and different phases are characterized by their corresponding damage value: points in the solid phase are assigned a damage value of 1 while points with a damage value of 0 are in the solution, and material points where the damage value is between 0 and 1 represent the dissolving/corroding/active region of the solid phase.115  The PD diffusion model over the bi-material domain accounts for the dissolution process and the transport of metal ions by considering damage-dependent diffusivity. The corrosion progression occurs autonomously via a phase change that depends on the molar concentration of metal atoms. The mathematical formulation of the PD corrosion damage model is provided below:115 
formula
formula
formula
Equation (12) is the PD diffusion equation. is the microdiffusivity of diffusion bonds and depends on the damage values of the two points associated with it. C is the metal concentration in the solution. The damage-dependent microdiffusivity k is given in Equation (13), which suggests that if both ends of a bond belong to the liquid phase, the value of microdiffusivity is kL, and it is easily calculated from D, the classical diffusion coefficient of the electrolyte.113  If both ends of a bond are solid points, then k is 0 (no mass transport). If one end is located in the liquid phase and the other end in the solid phase, then k is determined by kdiss, which is called the microdissolvability, determines the dissolution rate and can be calculated from corrosion kinetics. Equation (14) in the PD corrosion damage model describes the autonomous propagation of the corrosion front, where the damage value changes in response to variations in the molar concentration of metal atoms. That is to say, C ≤ Csat is considered as the liquid phase, C = Csolid as an intact solid, and Csat < C < Csolid as the dissolving solid (see Figure 10). This feature of the model enables the prediction of the corrosion progression over time, without the need for external interventions or triggers.
FIGURE 10.

Schematics of different phases and different diffusion bonds at the corrosion front, in the corrosion damage model.116 

FIGURE 10.

Schematics of different phases and different diffusion bonds at the corrosion front, in the corrosion damage model.116 

Close modal

Pitting

Chen and Bobaru114  converted the anodic reaction in the pitting process into equivalent diffusion in PD, and took 304 stainless steel as an example to simulate the pitting damage of the secondary surface layer. Concentration-dependent damage (CDD) model and damage-dependent corrosion (DDC) model are introduced. In the CDD model, when the solid node concentration is reduced due to corrosion, local mechanical damage occurs, and the damage coefficient is directly proportional to the concentration. In the DDC model, when the damage coefficient reaches a critical value, the solid nodes within a certain range will be transformed into liquid nodes, and the equivalent diffusion coefficient will change accordingly. The above model can realize the autonomous movement of the solid/liquid interface. This model is the first exploration of PD theory in corrosion damage, and the obtained results are in good agreement with the experimental data. In addition, Chen and Bobaru114  proposed a corrosion model for heterogeneous materials, in which the phase with good corrosion resistance was taken as the parent phase and the phase with poor corrosion resistance was taken as the second phase, as shown in Figure 11. Pitting damage that occurred in the vertical direction was larger. This model plays an important role in simulating or corrosion near inclusions.
FIGURE 11.

Illustrations of two cases of two-phase heterogeneous materials used to study pitting corrosion damage profiles after 20 s corrosion (DAM—corrosion damage index).114 

FIGURE 11.

Illustrations of two cases of two-phase heterogeneous materials used to study pitting corrosion damage profiles after 20 s corrosion (DAM—corrosion damage index).114 

Close modal

The effect of PD simulated repassivation and recrystallization on pitting corrosion of 304 stainless steel in 0.1 mol/L NaCl solution did not fully consider the electrochemical reaction on the electrode surface covered by the salt film, resulting in smaller pitting and lace-cover in 3D than the experimental results.117-118  Jafarzadeh, et al.,119  improved the model by using the classical pitting theory,120  and simulated the pitting growth under coverage of the salt film. Compared with the original model,114  the results of this model are closer to the real condition.

Crevice Corrosion

Jafarzadeh, et al.,121  introduced a new crevice corrosion damage PD model and verified it with experiments in the literature. The electrochemical mechanism is considered as a simple pattern that ion concentration is dependent on corrosion rate. This simplification makes it feasible to define local dissolved microfluxes at the solid/liquid interface. By modifying the PD formula to accommodate the event horizon of any shape, calculations can be made efficiently for areas with extreme aspect ratios, such as those observed in cracks. The model was verified by the experimental data of bolt washers. The simulation predicted the site and depth of a deep corrosion trench formed at certain distances from the crevice opening. The experimental crevice corrosion process is reproduced by far-field boundary conditions. The results also show that the two factors considered (diffusion-driven transport of dissolved metal ions and dissolution dependent on their concentration) are effective in simulating the evolution of crevice corrosion.

Intergranular Corrosion

IGC is a kind of local corrosion occurring along the grain boundaries of materials. As the anodic reaction rate of grain boundaries is larger than that of grains, grain boundaries are always corroded preferentially than grains. According to the heterogeneous material corrosion model,114  grains can be regarded as the parent phase and grain boundaries as the second phase when studying IGC.122  The model can also simulate the process from single IGC to the whole grain dissolution, that is, the corrosion of polycrystalline materials. The microparticles or precipitated phases of IMC on grain boundaries may be the second phase. Both grains and grain boundaries dissolve, and the crystal orientation can affect the corrosion of polycrystalline materials. It is worth noting that there are differences between IGC and galvanic corrosion here. IGC does not consider galvanic coupling and mainly focuses on the influence of grain size, orientation, and other factors on localized corrosion.

Galvanic Corrosion

Coupled with a peridynamics-corrosion fracture (PD-CF) model,123  the corrosion rate can be determined by solving the corresponding electrostatic problems. The experimental profile and that obtained by the PD model for the AE44-mild steel galvanic couple after 3 d of corrosion are shown in Figure 12. Although the magnitude of corrosion depth obtained by PD simulation is slightly smaller than that from the experimental data, their patterns agree very well.
FIGURE 12.

Corrosion profiles for the AE44-mild steel galvanic couple obtained from (a) experiments and (b) PD corrosion model (δ = 0.4 mm).123 

FIGURE 12.

Corrosion profiles for the AE44-mild steel galvanic couple obtained from (a) experiments and (b) PD corrosion model (δ = 0.4 mm).123 

Close modal

Research has shown that FEM-based or PF models of corrosion in galvanic couples often require an artificial “step-down” in the geometry at the material interface of the couple in order to produce reasonable results.123  This step is necessary for models based on partial differential equations (PDEs) due to difficulties in assigning appropriate boundary conditions at the galvanic couple interface and properly initializing the motion of the corrosion front. However, the introduction of such artificial geometrical modifications reduces the generality of the model, and when mechanical loadings are applied to a galvanic couple system, the stress profile near the interface may differ from the actual one, particularly in the early stages of the corrosion process. Additionally, if there is a strong stress dependency on the corrosion rate, the probability of obtaining incorrect results for coupled problems may be further increased. In contrast, the PD model presented here does not require any artificial changes in geometry at the interface to accurately predict corrosion progression as observed in experiments.

Stress Corrosion Cracking

PD model has been used for the first time to study hydrogen-induced SCC.124  The PD micromechanical model and PD hydrogen grain boundary-diffusion model have been established. It lays a foundation for further study of SCC with complex shapes, loading conditions, multiple fractures, and three dimensions. Crack initiation and propagation at pitting can induce SCC, and such cracks can propagate along the grain/transgranular. Relevant studies show that the microstructure of pits has a significant influence on the nucleation and propagation of simulated cracks.125  Research shows that the crack proceeding from the damaged layer on the surface of metal materials will lead to unpredictable brittle fracture after propagation through the corrosion product layer.126  SCC was also simulated by applying some fracture conditions in the mechano-chemical coupled PD model.116  Shi, et al.,127  used PD to study the crack growth of a carbon steel tube with a bit of pitting on its surface at the pitting under two different loading conditions, compressive stress, and bending stress. As shown in Figure 13, simulation results show that under compressive stress, cracks nucleate longitudinally and propagate along the direction of carbon steel pipe, while under bending stress, the crack nucleated in the circumferential direction and spread along the direction.
FIGURE 13.

Snapshots of crack propagation for corrosion pits at 1,500 time steps under (a) compressive stress and (b) bending stress (DMG—damage index).127 

FIGURE 13.

Snapshots of crack propagation for corrosion pits at 1,500 time steps under (a) compressive stress and (b) bending stress (DMG—damage index).127 

Close modal
In the subsequent study, Chen, et al.,128  improved the model proposed by Jafarzadeh, et al.,116  which could simulate any corrosion damage caused by the force-chemical coupling and mechanical strain. The transition time from pitting corrosion to crack nucleation, the size and shape of pitting at the moment of fracture, and the crack nucleation and growth in the pitting were obtained by taking the 2D SCC of a metal beam under the condition of three-point bending as an example. As shown in Figure 14, the model was verified by comparing it with the SCC experimental results of steam turbine steel.
FIGURE 14.

Pits and cracks: (a, c) experimental results (after 7,173 h exposure); and (b, d) PD-computed results (at 6,688 h exposure). (a) SEM image of the sample surface; (b) top view for PD-computed damage map; (c) back-side tomogram of experimental pits and cracks; and (d) back-side view of PD-computed damage (part of the crack deep surface is cut for a better view of the pits’ back side morphology and the crack linking them).128 

FIGURE 14.

Pits and cracks: (a, c) experimental results (after 7,173 h exposure); and (b, d) PD-computed results (at 6,688 h exposure). (a) SEM image of the sample surface; (b) top view for PD-computed damage map; (c) back-side tomogram of experimental pits and cracks; and (d) back-side view of PD-computed damage (part of the crack deep surface is cut for a better view of the pits’ back side morphology and the crack linking them).128 

Close modal

The CA method has been successfully applied in the simulation of many electrochemical problems. It is a powerful tool to study the complex physical and chemical processes such as pitting corrosion, formation and dissolution of passivated film, material diffusion, and formation of corrosion products. It is capable of simulating the mesoscopic corrosion damage evolution process. It should be pointed out that the simulation of the corrosion process by using CA also has its limitations, which are mainly manifested as the limitations of cellular state and cellular transformation rules. Although some basic rules of electrochemical reactions are introduced into the CA model, they are only expressed qualitatively, and the model cannot be calibrated by experimental observation. The results obtained by the model, especially the dynamic evolution process, have no specific time scale. Therefore, the CA model can be called a semideterministic model.

The principle of FEM is to solve the governing equation of a corrosion electric field. Its calculation is relatively high in spatial scale and time scale but accompanied by relatively weak characterization capability of microscopic characteristics. FEM is suitable for solving nonlinear and nonuniform corrosive medium problems, so it has certain advantages for the simulation of corrosion development and the dynamics process of evolution. However, the cost is expensive and prohibitive in 3D, and it is difficult to deal with infinite and semi-infinite domain problems.

PF is a flexible framework for the prediction and simulation of corrosion due to the autonomous movable interface. In the PF model, the phase transition and mass transport process are characterized by two distinguished but coupled PDEs, which significantly increase the calculation burden and algorithm complexity.129  Although the PF model introduces a length scale into the model through the thickness of the diffusion layer (the PF transition region between phases is the input of the problem), they are still local models and impose classical boundary conditions on the set of PDEs. The diffusion interface in the PF model leads to the transition area of the corrosion front. Moreover, the damage evolution in the PF model is strongly dependent on the specific choice of the energy functional.130-132 

As a nonlocal theory, PD has a mature application in the aspects of material fracture and heat transfer and can simulate the corrosion damage evolution of metal materials. The model can successfully simulate the local corrosion process through surface diffusion, ion transport, electrochemical reaction at the metal/solution interface, and free movement of the solid/liquid interface. The PD corrosion damage model can simulate the damage of the secondary surface layer and the formation of lacy cover in pitting corrosion, the growth of pitting corrosion under the salt film, simplifying the autocatalytic effect in crevice corrosion, IGC, potential distribution at the interface of corroded metal/electrolyte solution, and crack growth at pitting under stress. PD corrosion damage model can be calibrated by other numerical simulation results and experimental results, but it is still in the developing stage.

Table 1 compares the models reviewed in this study in terms of governing equation, numerical method, main applications, advantages, and disadvantages.

Table 1.

Comparison of Cmputational Models for Localized Corrosion115 

Comparison of Cmputational Models for Localized Corrosion115
Comparison of Cmputational Models for Localized Corrosion115

Explicit and Implicit Interface Tracking

The surface potential and current density distribution of the mobile interface can be influenced by metal corrosion particles, which can change the corrosion rate. The dynamic tracking of the corrosion interface is therefore crucial to build an accurate simulation model and is currently a major research challenge.133  In recent years, a dynamic interface model based on FEM has been proposed and used to simulate the evolution process of the active metal dissolution interface134-135  but the SIM, which needs to track the metal interface at every step, requires a lot of remapping in the whole simulation process. To ensure that each step of geometric evolution can meet the computational requirements, several methods have been introduced to arbitrary modeling, including the level set method (LSM)136-137  and the ALE138 . However, maintaining a high-quality grid throughout the simulation process remains a challenge, especially for complex corrosion modeling problems

With the continuous development of computer technology, some complex dynamic interface models have also been successfully implemented in the research of tracking corrosion interfaces, such as (i) FEM, some researchers have successfully used the finite volume method to simulate metal pitting.139  This method is especially suitable for dealing with mathematical problems with spatial discontinuities. The location of the pitting interface is determined according to the ion concentration in each volume element, thus eliminating the difficulty of creating a new grid.140  (ii) Meshfree method, which avoids the process of grid generation in numerical calculation and uses some arbitrarily distributed coordinate points to construct interpolation function discrete governing equation to track the expansion of complex interface, is a feasible method to simulate the problem of corrosion moving interface.141  (iii) Extended/generalized FEM (E/GFEM), which is a major improvement of the traditional FEM, uses discontinuous form function to characterize the discontinuity in the domain of domain, and is also a mesh-independent FEM.142-143  There have been relevant studies on the simulation of pitting corrosion by using it in combination with LSM.144 

Although there are more applications of the method for tracking moving interfaces, most of the methods are still in the theoretical stage. In terms of engineering applications and on a larger scale, popular models like PF and PD are not supported by definite commercial software and most researchers still need to write programs line by line. Implementation of these methods requires a certain reserve of professional knowledge, relevant mathematical basis, and strong computer programming ability, making it difficult for quite a few researchers and restricting their application to a large extent. Furthermore, the computational cost of implementing both models is prohibitively expensive, especially in 3D. However, recent advancements have considerably improved this situation.145-147  A fast convolution-based method has been proposed for efficient discretization of PD/nonlocal models, reducing memory allocation by several orders of magnitude. The result shows that PD problems that would have required years of computations with existing discretization methods can now be solved in a matter of days with the proposed method, making fast computation of fracture and damage with high accuracy possible.

Multiscale and Multiphysics Field

In the first place, computational modeling of localized corrosion has advanced to the point that major uncertainties in predictions are due to the lack of experimental measures of input parameters. Therefore, it is highly desirable to adapt emerging multiphysics programs for modeling corrosion processes, allowing corrosion scientists and engineers to access guidelines for material design more easily. Several mechanisms contribute to structural corrosion, including processes at the megascale, material deformation, electrochemical reactions, mass transport, and fluid dynamics.148-149 

Due to the interdependency between these mechanisms, it is highly desirable to use a multiphysics approach to model and simulate the corrosion process within a feasible timeframe. The problem is multidisciplinary, with a large number of variables and controlling factors involved such as material composition, strengthening mechanisms, microstructure, loading modes and frequency, machining and residual stresses, environmental composition, cyclic effects of environmental variables, etc. Expertise in many scientific fields is required for the analysis and control of corrosion phenomena in structural alloys.

In the next place, corrosion degradation involves processes that are active at different time and length scales, ranging from electrochemically driven reactions occurring at the interface between material and environment at the atomistic level to intergranular crack propagation driven by grain boundary misorientation and selective attack at the mesoscale level. At the system and component scales, computational tools are mainly used to predict the coupling between dissimilar materials. Therefore, studying multiscale corrosion modeling can help corrosion researchers systematically understand the occurrence and development of corrosion, and take beneficial measures to reduce corrosion hazards. However, there are currently few relevant reports on multiscale corrosion modeling. In 2017, a review article on multiscale and multifield corrosion modeling was published.150  Figure 15 outlines the research objectives and the implementation process of multiscale corrosion modeling. Nickerson, et al.,150  conducted macroscopic and microscopic scale analyses on galvanic corrosion. The macroscopic scale calculation model was used to predict the galvanic corrosion rate, assist in material selection for structures, and complete the relevant modeling at the system and component scales. The microscale calculation model was used to optimize material processing technology, provide guidance for material design, obtain the ideal material microstructure, increase material corrosion resistance, and achieve comprehensive protection from the material to the structure and targeted optimization from the structure to the material.
FIGURE 15.

Two-fold objective of a multiscale corrosion modeling framework.150 

FIGURE 15.

Two-fold objective of a multiscale corrosion modeling framework.150 

Close modal

Moreover, corrosion of materials is often not solely caused by a single physical driver. It is often accompanied by various influences such as stress, flow, thermal, and magnetic fields. The existence of multiple physical influences makes it difficult to model accurately.151  Taking aluminum alloy materials used in seaplanes, ships, and offshore photovoltaic supports as examples, their service environment is mainly the seawater/atmosphere interface, which is a typical environment with multifield and multiphase couplings.152-154  The corrosion of these structural materials under the influence of temperature, pH, dissolved oxygen, sea/atmosphere alternations, and waves is a key factor affecting the long-term stable use of marine engineering and weapons equipment. However, marine tests are complicated, time-consuming, and costly. Therefore, it is particularly important to use computational models to simulate the local corrosion damage of metal materials under multifield and multiphase couplings.

  • In this paper, the characteristics and the application of four numerical simulation methods in localized corrosion, including CA, FEM, PF model, and PD model, are reviewed. The review shows that the CA is a powerful tool for studying pitting corrosion and passivation film formation and complex microprocesses, such as dissolution and substance diffusion, widely used in the microscopic mechanism research of metal corrosion. FEM is appropriate for solving nonlinear and nonuniform corrosive medium problems, so it has certain advantages for the simulation of corrosion development and the dynamics process of evolution. As another important method to simulate the problem of the variable interface, the PF method does not require the application of boundary conditions at the interface, which promotes the application in the problem of corrosion morphology evolution. At present, PD, as a nonlocal theory, can successfully simulate the corrosion damage evolution of metal, which can be calibrated by other numerical simulation results and experimental results. However, the corrosion of metal materials is often affected by multiple influencing factors and multiple physics fields, so the research on computational models on the evolution of local corrosion damage is still in the development stage.

  • In summary, numerical simulation has become one of the important means to study metal corrosion and protection, and with the rapid development of software science and computer technology, the operation speed and accuracy of numerical simulation will be greatly improved, numerical simulation in the field of corrosion application will be more prominent advantages. According to the different corrosion systems and the problem to be solved of selecting the appropriate simulation method, one can make full use of its strengths and circumvent its weaknesses.

Trade name.

(1)

UNS numbers are listed in Metals & Alloys in the Unified Numbering System, published by the Society of Automotive Engineers (SAE International) and cosponsored by ASTM International.

This research effort was financially supported by Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai) [Fund, D5110200598]. Generous thanks are extended to the Fund Monitors for their encouragement and support of this exhaustive study.

1.
Hou
B.
,
Li
X.
,
Ma
X.
,
npj Mater. Degrad.
1
(
2017
):
p
.
1
10
.
2.
Koch
G.
,
Varney
J.
,
Thompson
N.
,
Moghissi
O.
,
Gould
M.
,
Payer
J.
,
International Measures of Prevention, Application, and Economics of Corrosion Technologies Study
,
ed.
Jacobson
G.
(
Houston, TX
:
NACE
,
2016
), https://www.doc88.com/p-99559428192496.html.
3.
Doane
S.
,
Fukushima Cleanup Could Drag on for Decades. CBSNews
,
2014
.
4.
Jiuzi
Q.
,
China Emerg. Rescue
(
2014
):
p
.
43
45
.
5.
Yan
T.
,
The Hidden Crisis of “More Dragon Water Control” Urgently Needs to End. China Land and Resources News, 2014-08-19 (003)
.
6.
Chen
Z.
,
Chin. J. Solid Mech.
40
(
2019
):
p
.
99
116
.
7.
Made in China 2025. State Council. 2015
.
8.
Council N R
,
Research Opportunities in Corrosion Science and Engineering
(
Washington, DC
:
The National Academies Press
,
2011
),
p
.
120
.
9.
Xiaodong
D.
,
Cellular Automata Theory Research and Simulation Applications
(
Beijing, China
:
Science Press
,
2012
).
10.
Chowdhury
D.
,
Santen
L.
,
Schadschneider
A.
,
Phys. Rep.-Rev. Sec. Phys. Lett.
329
(
2000
):
p
.
199
329
.
11.
Hernández Encinas
A.
,
Hernández Encinas
L.
,
Hoya White
S.
,
Martín Del Rey
A.
,
Rodríguez Sánchez
G.
,
Adv. Eng. Softw.
38
(
2007
):
p
.
372
378
.
12.
Lee
Y.
,
Kouvroukoglou
S.
,
McIntire
L.V.
,
Zygourakis
K.
,
Biophys. J.
69
(
1995
):
p
.
1284
1298
.
13.
Chopard
B.
,
Droz
M.
,
Cellular Automata Modeling of
Physical Systems
(
Cambridge, United Kingdom
:
Cambridge University Press
,
1998
).
14.
Meakin
P.
,
Jøssang
T.
,
Feder
J.
,
Phys. Rev. E
48
(
1993
):
p
.
2906
2916
.
15.
Taleb
A.
,
Stafiej
J.
,
Chaussé
A.
,
Messina
R.
,
Badiali
J.P.
,
J. Electroanal. Chem.
500
(
2001
):
p
.
554
561
.
16.
Taleb
A.
,
Chaussé
A.
,
Dymitrowska
M.
,
Stafiej
J.
,
Badiali
J.P.
,
J. Phys. Chem. B
108
(
2004
):
p
.
952
958
.
17.
Li
L.
,
Li
X.G.
,
Dong
C.F.
,
Cheng
Y.F.
,
Corros. Eng. Sci. Technol.
46
(
2011
):
p
.
340
345
.
18.
Lei
L.
,
Xiaogang
L.
,
Chaofang
D.
,
Yizhong
H.
,
Electrochim. Acta
54
(
2009
):
p
.
6389
6395
.
19.
Ren
K.
,
Gao
S.
,
Liu
P.
,
Dong
A.Y.
,
Theor. Appl. Mech. Lett.
1
(
2011
):
p
.
61004
61007
.
20.
Lishchuk
S.V.
,
Akid
R.
,
Worden
K.
,
Michalski
J.
,
Corros. Sci.
53
(
2011
):
p
.
2518
2526
.
21.
Taleb
A.
,
Vautrin-Ul
C.
,
Mendy
H.
,
Stafiej
J.
,
Chausse
A.
, “
Mesoscopic Modeling of Corrosion Processes: Pitting Morphology Evolution
,”
in
2nd International Conference on Simulation of Electrochemical Processes
,
Myrtle Beach, SC
(
2007
).
22.
Stafiej
J.
,
Di Caprio
D.
,
Bartosik
Ł.
,
J. Supercomput.
65
(
2013
):
p
.
697
709
.
23.
Aarão Reis
F.D.A.
,
Stafiej
J.
,
Badiali
J.P.
,
J. Phys. Chem. B
110
(
2006
):
p
.
17554
17562
.
24.
Pidaparti
R.M.
,
Fang
L.
,
Palakal
M.J.
,
Comput. Mater. Sci.
41
(
2009
):
p
.
255
265
.
25.
Pidaparti
R.M.
,
Palakal
M.J.
,
Fang
L.
,
AIAA J.
42
(
2004
):
p
.
2562
2569
.
26.
Pidaparti
R.M.
,
Puri
P.A.
,
Palakal
M.J.
,
Kashyap
A.
,
AIAA J.
45
(
2008
):
p
.
2575
2579
.
27.
Taleb
A.
,
Chaussé
A.
,
Dymitrowska
M.
,
Stafiej
J.
,
Badiali
J.P.
,
J. Phys. Chem. B
108
(
2004
):
p
.
952
958
.
28.
Vautrin-Ul
C.
,
Chaussé
A.
,
Stafiej
J.
,
Badiali
J.P.
,
Condens. Matter Phys.
7
(
2004
):
p
.
813
828
.
29.
Vautrin-Ul
C.
,
Taleb
A.
,
Stafiej
J.
,
Chaussé
A.
,
Badiali
J.P.
,
Electrochim. Acta
52
(
2008
):
p
.
7802
7810
.
30.
Saunier
J.
,
Chaussé
A.
,
Stafiej
J.
,
Badiali
J.P.
,
J. Electroanal. Chem.
563
(
2004
):
p
.
239
.
31.
Saunier
J.
,
Dymitrowska
M.
,
Chaussé
A.
,
Stafiej
J.
,
Badiali
J.P.J.S.
,
J. Electroanal. Chem.
582
(
2005
):
p
.
267
273
.
32.
Fatoba
O.O.
,
Leiva-Garcia
R.
,
Lishchuk
S.V.
,
Larrosa
N.O.
,
Akid
R.
,
Corros. Sci.
137
(
2018
):
p
.
83
97
.
33.
Van der Weeën
P. N
,
Zimer
A.M.
,
Pereira
E.C.
,
Mascaro
L.H.
,
Bruno
O.M.
,
De Baets
B.
,
Corros. Sci.
82
(
2014
):
p
.
133
144
.
34.
Pérez-Brokate
C.F.
,
di Caprio
D.
,
Féron
D.
,
de Lamare
J.
,
Chaussé
A.
,
Corros. Eng. Sci. Technol.
52
(
2017
):
p
.
186
193
.
35.
Pérez-Brokate
C.F.
,
di Caprio
D.
,
Féron
D.
,
de Lamare
J.
,
Chaussé
A.
,
Corros. Eng. Sci. Technol.
52
(
2017
):
p
.
605
610
.
36.
Zenkri
M.
,
di Caprio
D.
,
Pe Rez-Brokate
C.
,
Fe Ron
D.
,
de Lamare
J.
,
Chausse
A.
,
Ben Cheikh Larbi
F.
,
Raouafi
F.
,
Condens. Matter Phys.
20
(
2017
):
p
.
33802
.
37.
Xuefeng
L.
,
Lei
H.
,
Dongmei
C.
,
Model. Simul. Mater. Sci. Eng.
29
(
2021
):
p
.
65018
.
38.
Sun
W.
,
Wang
L.
,
Wu
T.
,
Liu
G.
,
Corros. Sci.
78
(
2014
):
p
.
233
243
.
39.
Deshpande
K.B.
,
Corros. Sci.
52
(
2010
):
p
.
3514
3522
.
40.
Pan
Z.
,
Finite Element Analysis and Application
(
Beijing, China
:
Tsinghua University
,
2004
).
41.
Sharland
S.M.
,
Corros. Sci.
28
(
1988
):
p
.
621
630
.
42.
Laycock
N.J.
,
White
S.P.
,
J. Electrochem. Soc.
148
(
2001
):
p
.
B264
B275
.
43.
Sharland
S.M.
,
Tasker
P.W.
,
Corros. Sci.
28
(
1988
):
p
.
603
620
.
44.
Sharland
S.M.
,
Jackson
C.P.
,
Diver
A.J.
,
Corros. Sci.
29
(
1989
):
p
.
1149
1166
.
45.
Engelhardt
G.
,
Urquidi-Macdonald
M.
,
Macdonald
D.D.
,
Corros. Sci.
38
(
1997
):
p
.
419
441
.
46.
Laycock
N.J.
,
White
S.P.
,
Noh
J.S.
,
Wilson
P.T.
,
Newman
R.C.
,
J. Electrochem. Soc.
145
(
1998
):
p
.
1101
1108
.
47.
Ghahari
S.M.
,
Krouse
D.P.
,
Laycock
N.J.
,
Rayment
T.
,
Padovani
C.
,
Suter
T.
,
Mokso
R.
,
Marone
F.
,
Stampanoni
M.
,
Monir
M.
,
Corros. Eng. Sci. Technol.
46
(
2011
):
p
.
205
211
.
48.
Galvele
J.R.
,
J. Electrochem. Soc.
123
(
1976
):
p
.
464
474
.
49.
Gravano
S.M.
,
Galvele
J.R.
,
Corros. Sci.
24
(
1984
):
p
.
517
534
.
50.
Malki
B.
,
Souier
T.
,
Baroux
B.
,
J. Electrochem. Soc.
155
(
2008
):
p
.
583
587
.
51.
Amri
J.
,
Gulbrandsen
E.
,
Nogueira
R.P.
,
Corros. Sci.
52
(
2010
):
p
.
1728
1737
.
52.
Alkire
R.C.
,
Deligianni
H.
,
Ju
J.B.
,
J. Electrochem. Soc.
137
(
1990
):
p
.
818
824
.
53.
Harb
J.N.
,
Alkire
R.C.
,
J. Electrochem. Soc.
138
(
1991
):
p
.
2594
2600
.
54.
Srinivasan
J.
,
Liu
C.
,
Kelly
R.G.
,
J. Electrochem. Soc.
163
(
2016
):
p
.
C694
C703
.
55.
Srinivasan
J.
,
Kelly
R.G.
,
Corrosion
76
(
2017
):
p
.
613
633
.
56.
Marshall
R.S.
,
Katona
R.M.
,
Melia
M.A.
,
Kelly
R.G.
,
J. Electrochem. Soc.
169
(
2022
):
p
.
021506
.
57.
Xu
Y.
,
Pickering
H.W.
,
J. Electrochem. Soc.
140
(
1993
):
p
.
658
668
.
58.
Cho
K.
,
Pickering
H.W.
, “
Critical Factors in Localized Corrosion
,”
in
The Electrochemical Society Softbound Proceedings Series
(
Pennington, NJ
:
The Electrochemical Society
,
1992
).
59.
Vankeerberghen
M.
,
Corrosion
60
(
2004
):
p
.
707
717
.
60.
Lee
J.S.
,
Reed
M.L.
,
Kelly
R.G.
,
J. Electrochem. Soc.
151
(
2004
):
p
.
B423
B433
.
61.
Percheron
A.
,
Vuillemin
B.
,
Oltra
R.
,
Markey
L.
,
J. Appl. Electrochem.
41
(
2011
):
p
.
355
361
.
62.
Malki
B.
,
Berthomé
G.
,
Souier
T.
,
Boissy
C.
,
Guillotte
I.
,
Baroux
B.B.
,
J. Electrochem. Soc.
168
(
2021
):
p
.
101504
.
63.
Ding
J.
,
He
W.
,
Liu
Y.
,
Zhang
C.
,
Wang
H.
,
Han
E.
,
Coatings
12
(
2022
):
p
.
592
.
64.
Deshpande
K.B.
,
Electrochim. Acta
56
(
2011
):
p
.
1737
1745
.
65.
Xiao
J.
,
Chaudhuri
S.
,
Electrochim. Acta
56
(
2011
):
p
.
5630
5641
.
66.
Yin
L.
,
Jin
Y.
,
Leygraf
C.
,
Birbilis
N.
,
Pan
J.
,
J. Electrochem. Soc.
56
(
2017
):
p
.
C75
C84
.
67.
Yin
L.
,
Jin
Y.
,
Leygraf
C.
,
Pan
J.
,
J. Electrochem. Soc.
164
(
2017
):
p
.
C768
C778
.
68.
Wang
Y.
,
Yin
L.
,
Jin
Y.
,
Pan
J.
,
Leygraf
C.
,
J. Electrochem. Soc.
164
(
2017
):
p
.
C1035
C1043
.
69.
Oltra
R.
,
Zimmer
A.
,
Sorriano
C.
,
Rechou
F.
,
Borkowski
C.
,
Néel
O.
,
Electrochim. Acta
56
(
2011
):
p
.
7038
7044
.
70.
Deshpande
K.B.
,
Corros. Sci.
52
(
2010
):
p
.
2819
2826
.
71.
King
A.D.
,
Lee
J.S.
,
Scully
J.R.
,
J. Electrochem. Soc.
62
(
2016
):
p
.
C342
C356
.
72.
Cui
F.
,
Presuel-Moreno
F.J.
,
Ke
R.G.
,
Corrosion
53
(
2006
):
p
.
251
263
.
73.
Thébault
F.
,
Vuillemin
B.
,
Oltra
R.
,
Ogle
K.
,
Allely
C.
,
Electrochim. Acta
53
(
2008
):
p
.
5226
5234
.
74.
Presuel-Moreno
F.J.
,
Goldman
M.E.
,
Kelly
R.G.
,
Scully
J.R.
,
Proc. Electrochem. Soc.
152
(
2004
):
p
.
255
269
.
75.
Cross
S.R.
,
Woollam
R.
,
Shademan
S.
,
Schuh
C.A.
,
Corros. Sci.
77
(
2013
):
p
.
297
307
.
76.
Cross
S.R.
,
Schuh
C.A.
,
Corros. Sci.
116
(
2017
):
p
.
53
65
.
77.
Wenman
M.R.
,
Trethewey
K.R.
,
Jarman
S.E.
,
Chard-Tuckey
P.R.
,
Acta Mater.
56
(
2008
):
p
.
4125
4136
.
78.
Badics Nuclear Fuel Industries Ltd.
Z.
,
Nondestruct. Test. Mater.
(
1995
):
p
.
21
29
.
79.
Cailletaud
G.
,
Musienko
A.
,
Acta Mater.
57
(
2009
):
p
.
3840
3855
.
80.
Turnbull
A.
,
Wright
L.
,
Crocker
L.
,
Corros. Sci.
52
(
2010
):
p
.
1492
1498
.
81.
Zhu
L.K.
,
Yan
Y.
,
Qiao
L.J.
,
Volinsky
A.A.
,
Corros. Sci.
77
(
2013
):
p
.
360
368
.
82.
Jasra
Y.
,
Singhal
S.
,
Upman
R.
,
Saxena
R.K.
,
Mater. Today Proc.
26
(
2020
):
p
.
2314
2322
.
83.
Katona
R.M.
,
Burns
J.T.
,
Schaller
R.F.
,
Kelly
R.G.
,
Corros. Sci.
209
(
2022
):
p
.
110756
.
84.
Karma
A.
,
Rappel
W.-J.
,
Phys. Rev. E
53
(
1996
):
p
.
R3017
R3020
.
85.
Chen
L.
,
Ann. Rev. Mater. Res.
32
(
2002
):
p
.
113
140
.
86.
Chen
L.
,
J. Am. Ceram. Soc.
91
(
2008
):
p
.
1835
1844
.
87.
Wang
J.
,
Zhang
J.
,
Int. J. Solids Struct.
50
(
2013
):
p
.
3597
3609
.
88.
Mai
W.
,
Soghrati
S.
,
Buchheit
R.G.
,
Corros. Sci.
110
(
2016
):
p
.
157
166
.
89.
Ansari
T.Q.
,
Xiao
Z.
,
Hu
S.
,
Li
Y.
,
Luo
J.
,
Shi
S.
,
npj Comput. Mater.
4
(
2018
):
p
.
330
338
.
90.
Chadwick
A.F.
,
Enrique
R.A.
,
Du
S.
,
Stewart
J.A.
,
Thornton
K.
,
J. Electrochem. Soc.
165
(
2018
):
p
.
C633
C646
.
91.
Tsuyuki
C.
,
Yamanaka
A.
,
Ogimoto
Y.
,
Sci. Rep.
8
(
2018
):
p
.
1
14
.
92.
Brewick
P.T.
,
J. Electrochem. Soc.
169
(
2022
):
p
.
11503
.
93.
Sahu
S.
,
Frankel
G.S.
,
J. Electrochem. Soc.
169
(
2022
):
p
.
20557
.
94.
Gao
H.
,
Ju
L.
,
Li
X.
,
Duddu
R.
,
J. Comput. Phys.
406
(
2020
):
p
.
109191
.
95.
Xiao
Z.H.
,
Hu
S.H.
,
Luo
J.L.
,
Shi
S.Q.
Jr.
,
Henager
C.H.
,
Comput. Mater. Sci.
149
(
2018
):
p
.
37
48
.
96.
Mai
W.
,
Soghrati
S.
,
Electrochim. Acta
260
(
2018
):
p
.
290
304
.
97.
Ansari
T.Q.
,
Luo
J.L.
,
Shi
S.Q.
,
J. Electrochem. Soc.
167
(
2020
):
p
.
61508
.
98.
Mai
W.
,
Soghrati
S.
,
Corros. Sci.
125
(
2017
):
p
.
87
98
.
99.
Nguyen
T.
,
Bolivar
J.
,
Shi
Y.
,
Eacute
R.
,
Eacute
T.
,
King
J.A.
,
Fregonese
M.
,
Adrien
J.
,
Buffiere
J.
,
Baietto
M.
,
Corros. Sci.
132
(
2018
):
p
.
146
160
.
100.
Lin
C.
,
Ruan
H.
,
Shi
S.
,
Electrochim. Acta
310
(
2019
):
p
.
240
255
.
101.
Cui
C.
,
Ma
R.
,
Martínez-Pañeda
E.
,
J. Mech. Phys. Solids
147
(
2021
):
p
.
104254
.
102.
Lin
C.
,
Ruan
H.
,
Electrochim. Acta
395
(
2021
):
p
.
139196
.
103.
Silling
S.A.
,
J. Mech. Phys. Solids
48
(
2000
):
p
.
175
209
.
104.
Rabczuk
T.
,
Ren
H.
,
Eng. Geol.
225
(
2017
):
p
.
42
48
.
105.
Sanchez
G.
,
Aperador
W.
,
Cerón
A.
,
Indian Chem. Eng.
62
(
2020
):
p
.
277
286
.
106.
Hu
W.
,
Ha
Y.D.
,
Bobaru
F.
,
Comput. Methods Appl. Mech. Eng.
271-220
(
2012
):
p
.
247
261
.
107.
Karpenko
O.
,
Oterkus
S.
,
Oterkus
E.
,
Eng. Fract. Mech.
261
(
2022
):
p
.
108212
.
108.
Bobaru
F.
,
Duangpanya
M.
,
Int. J. Heat Mass Transfer
53
(
2010
):
p
.
4047
4059
.
109.
Bobaru
F.
,
Duangpanya
M.
,
J. Comput. Phys.
231
(
2012
):
p
.
2764
2785
.
110.
Wang
L.
,
Xu
J.
,
Wang
J.
,
Int. J. Heat Mass Transfer
118
(
2018
):
p
.
1284
1292
.
111.
Nikolaev
P.
,
Sedighi
M.
,
Jivkov
A.P.
,
Margetts
L.
,
Int. J. Heat Mass Transfer
185
(
2022
):
p
.
122327
.
112.
Askari
E.
,
Bobaru
F.
,
Lehoucq
R.B.
,
Parks
M.L.
,
Silling
S.A.
,
Weckner
O.
,
J. Phys. Conf. Ser.
125
(
2008
):
p
.
12078
.
113.
Gerstle
W.
,
Silling
S.
,
Read
D.
,
Tewary
V.
,
Lehoucq
R.
,
Comput. Mater. Contin.
8
(
2008
):
p
.
75
92
.
114.
Chen
Z.
,
Bobaru
F.
,
J. Mech. Phys. Solids
78
(
2015
):
p
.
352
381
.
115.
Jafarzadeh
S.
,
Chen
Z.
,
Bobaru
F.
,
Corros. Rev.
37
(
2019
):
p
.
419
439
.
116.
Jafarzadeh
S.
,
Chen
Z.
,
Li
S.
,
Bobaru
F.
,
Electrochim. Acta
323
(
2019
):
p
.
134795
.
117.
Jafarzadeh
S.
,
Chen
Z.
,
Zhao
J.
,
Bobaru
F.
,
Corros. Sci.
150
(
2019
):
p
.
17
31
.
118.
Almuaili
F.A.
,
Characterisation of 3D Pitting Corrosion Kinetics of Stainless Steel in Chloride Containing Environments
(
Manchester, United Kingdom
:
University of Manchester
,
2017
).
119.
Jafarzadeh
S.
,
Bobaru
F.
,
Chen
Z.
,
Corrosion
74
(
2018
):
p
.
393
414
.
120.
Gaudet
G.T.
,
Mo
W.T.
,
Hatton
T.A.
,
Tester
J.W.
,
Tilly
J.
,
Isaacs
H.S.
,
Newman
R.C.
,
AIChE J.
32
(
1986
):
p
.
949
958
.
121.
Jafarzadeh
S.
,
Zhao
J.
,
Shakouri
M.
,
Bobaru
F.
,
Electrochim. Acta
401
(
2022
):
p
.
139512
.
122.
Jafarzadeh
S.
,
Chen
Z.
,
Bobaru
F.
,
J. Electrochem. Soc.
165
(
2018
):
p
.
C362
C374
.
123.
Zhao
J.
,
Jafarzadeh
S.
,
Rahmani
M.
,
Chen
Z.
,
Kim
Y.
,
Bobaru
F.
,
Electrochim. Acta
391
(
2021
):
p
.
138968
.
124.
De Meo
D.
,
Diyaroglu
C.
,
Zhu
N.
,
Oterkus
E.
,
Siddiq
M.A.
,
Int. J. Hydrogen Energy
41
(
2016
):
p
.
6593
6609
.
125.
De Meo
D.
,
Russo
L.
,
Oterkus
E.
, “
Peridynamics for Predicting Pit-to-Crack Transition
,”
in
58th AIAA/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference (
Grapevine, TX
:
AIAA
,
2017
).
126.
Li
S.
,
Chen
Z.
,
Wang
F.
,
Cui
B.
,
Tan
L.
,
Bobaru
F.
,
J. Electrochem. Soc.
323
(
2016
):
p
.
C784
C790
.
127.
Shi
C.
,
Gong
Y.
,
Yang
Z.G.
,
Tong
Q.
,
Eng. Fract. Mech.
219
(
2019
):
p
.
106604
.
128.
Chen
Z.
,
Jafarzadeh
S.
,
Zhao
J.
,
Bobaru
F.
,
J. Mech. Phys. Solids
146
(
2021
):
p
.
104203
.
129.
Soghrati
S.S.O.E.
,
Chen
Y.
,
Mai
W.
,
Finite Elem. Anal. Des.
140
(
2018
):
p
.
23
37
.
130.
Borden
M.J.
,
Verhoosel
C.V.
,
Scott
M.A.
,
Hughes
T.J.R.
,
Landis
C.M.
,
Comput. Methods Appl. Mech. Eng.
217-220
(
2012
):
p
.
77
95
.
131.
Borden
M.J.
,
Hughes
T.J.R.
,
Landis
C.M.
,
Anvari
A.
,
Lee
I.J.
,
Comput. Methods Appl. Mech. Eng.
312
(
2016
):
p
.
130
166
.
132.
Geelen
R.J.M.
,
Liu
Y.
,
Hu
T.
,
Tupek
M.R.
,
Dolbow
J.E.
,
Comput. Methods Appl. Mech. Eng.
348
(
2019
):
p
.
680
711
.
133.
Sarkar
S.
,
Warner
J.E.
,
Aquino
W.
,
Corros. Sci.
65
(
2012
):
p
.
502
511
.
134.
Deshpande
K.B.
,
Electrochim. Acta
56
(
2011
):
p
.
1737
1745
.
135.
Brewick
P.T.
,
Kota
N.
,
Lewis
A.C.
,
DeGiorgi
V.G.
,
Geltmacher
A.B.
,
Qidwai
S.M.
,
Corros. Sci.
129
(
2017
):
p
.
54
69
.
136.
Wang
Y.
,
Wharton
J.A.
,
Shenoi
R.A.
,
Eng. Struct.
128
(
2016
):
p
.
1
14
.
137.
Sethian
J.A.
,
Proc. Natl. Acad. Sci. USA
93
(
1996
):
p
.
1591
1595
.
138.
Sun
W.
,
Wang
L.
,
Wu
T.
,
Liu
G.
,
Corros. Sci.
78
(
2014
):
p
.
233
243
.
139.
Scheiner
S.
,
Hellmich
C.
,
Comput. Methods Appl. Mech. Eng.
198
(
2009
):
p
.
2898
2910
.
140.
Onishi
Y.
,
Takiyasu
J.
,
Amaya
K.
,
Yakuwa
H.
,
Hayabusa
K.
,
Corros. Sci.
63
(
2012
):
p
.
210
224
.
141.
Soghrati
S.
,
Mai
W.
,
Liang
B.
,
Buchheit
R.G.
,
J. Comput. Phys.
281
(
2015
):
p
.
225
236
.
142.
Belytschko
T.
,
Gracie
R.
,
Ventura
G.
,
Model. Simul. Mater. Sci. Eng.
17
(
2009
):
p
.
43001
.
143.
Chessa
J.
,
Smolinski
P.
,
Belytschko
T.
,
Int. J. Numer. Methods Eng.
53
(
2002
):
p
.
1959
1977
.
144.
Duddu
R.
,
Comput. Mech.
54
(
2014
):
p
.
613
627
.
145.
Jafarzadeh
S.
,
Larios
A.
,
Bobaru
F.
,
J. Peridyn. Nonlocal Model.
2
(
2020
):
p
.
85
110
.
146.
Jafarzadeh
S.
,
Wang
L.
,
Larios
A.
,
Bobaru
F.
,
Comput. Methods Appl. Mech. Eng.
375
(
2021
):
p
.
113633
.
147.
Wang
L.
,
Jafarzadeh
S.
,
Bobaru
F.
,
J. Peridyn. Nonlocal Model.
, https://doi.org/10.21203/rs.3.rs-2046856/v1.
148.
Taylor
C.D.
,
Electrochem. Soc. Interface
23
(
2014
):
p
.
59
64
.
149.
Taylor
C.D.
,
Corros. Eng. Sci. Technol.
50
(
2015
):
p
.
490
508
.
150.
Nickerson
W.C.
,
Iyyer
N.
,
Legg
K.
,
Amiri
M.
,
Corros. Rev.
35
(
2017
):
p
.
205
223
.
151.
de Borst
R.R.D.B.
,
Comput. Mater. Sci.
43
(
2008
):
p
.
1
15
.
152.
Xia
D.H.
,
Ji
Y.Y.
,
Mao
Y.C.
,
Deng
C.M.
,
Zhu
Y.
,
Hu
W.B.
,
Acta Metall. Sin.
59
(
2023
):
p
.
297
308
.
153.
Mao
Y.C.
,
Zhu
Y.
,
Sun
S.K.
,
Qin
Z.B.
,
Xia
D.H.
,
Hu
W.B.
,
J. Chin. Soc. Corros. Prot.
43
(
2023
):
p
.
47
54
.
154.
Liu
W.
,
Wang
J.
,
J. Chin. Soc. Corros. Prot.
30
(
2010
):
p
.
504
512
.