Next Article in Journal
A New Absorption Configuration of Partial Lean Solution Vaporization–Compression for CO2 Capture
Previous Article in Journal
Mine Exogenous Fire Detection Algorithm Based on Improved YOLOv9
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of CO2 Corrosion Modeling for Carbon Capture, Utilization and Storage (CCUS) Infrastructure

Department of Energy, Aalborg University, Niels Bohrs Vej 8, 6700 Esbjerg, Denmark
*
Author to whom correspondence should be addressed.
Processes 2026, 14(1), 170; https://doi.org/10.3390/pr14010170
Submission received: 8 December 2025 / Revised: 22 December 2025 / Accepted: 31 December 2025 / Published: 4 January 2026
(This article belongs to the Section Energy Systems)

Abstract

CO2 corrosion remains a critical challenge for the safe and reliable operation of Carbon Capture, Utilization, and Storage (CCUS) infrastructure. This review summarizes CO2 corrosion implications from material selection, exposure time, CO2 phase behavior, flow conditions, and impurities such as H2O, O2, SOx, NOx, and H2S. CO2 corrosion modeling has, since early works by de Waard in 1975, expanded to a wide range of models and software tools, many of which have already been reviewed and compared. This work provides a historical timeline and a comparative summary of models and software tools to assist in selecting models for CCUS applications. Modeling approaches are classified into empirical, semi-empirical, and mechanistic categories, with their assumptions, strengths, and limitations. CO2 corrosion modeling has persistent challenges relating to data quality, data quantity, and parameter interactions, which reduce model accuracy, especially for machine learning approaches. The provided perspective emphasizes that machine learning and hybrid modeling approaches for CO2 corrosion prediction are gaining popularity, and their effectiveness is currently limited by the quality and quantity of available corrosion data. The provided opportunities include recommendations for standardized experimental procedures and hybrid modeling strategies that combine physics-based insights from mechanistic modeling approaches with data-driven machine learning approaches.

1. Introduction

Carbon dioxide (CO2) is a major greenhouse gas, and large-scale carbon capture, utilization, and storage (CCUS) is essential to mitigate its impact on climate change [1,2,3]. In CCUS, CO2 is captured from industrial sources or from the atmosphere and then generally transported by ship or pipeline for storage or utilization [4,5]. However, the captured CO2 often contains impurities such as H2O, NOx, SOx, H2S, and O2, where the concentration and type of impurity depend on the capture technology and the source from which the CO2 is captured [6,7]. These impurities can compromise the infrastructure’s integrity by creating a corrosive environment [8,9]. Enhanced oil recovery (EOR) has been used since the 1970s by injecting CO2 into oil wells to extract more oil from reservoirs [10]. The pipelines used for these have been constructed of carbon steel, which makes it crucial to understand the risks of impurity-induced corrosion [7,10]. Studies have examined CO2 corrosion and its mechanisms in depth [11,12]. Studies have also analyzed corrosion from impure CO2, as demonstrated by Simonsen et al. in their review study [13]. Additionally, studies have shown that various factors, such as temperature, pressure, impurity type, impurity concentration, flow rate, and exposure time, can significantly influence corrosion behavior [12,14,15,16,17,18,19,20]. Other studies show synergistic effects between impurities such as NO2 and SO2, where NO2 catalyzes the reaction with SO2 and H2O, posing further challenges for the occurrence of corrosion [21].
The primary challenges for CO2 transportation are to balance the cost of pipeline materials, including maintenance, and potential corrosion. Failure to strike the right balance relates to unnecessary expenses and safety risks. The benefits of corrosion modeling are a common initiative to make informed decisions on this balance, which, in the case of CO2 corrosion, includes the following:
  • Risk assessment: Modeling enables the assessment of corrosion risks under various operating conditions, allowing engineers to identify potential trouble spots and take preventive measures [13,22].
  • Material selection: Engineers can use models to evaluate the performance of different materials under specific conditions and choose materials that are more resistant to CO2 corrosion. Material challenges for CO2 transport related to CCUS are investigated in [22,23].
  • Corrosion control strategies: Modeling helps in the development and optimization of corrosion control strategies, such as the use of inhibitors or coatings, to mitigate the impact of CO2 on equipment and pipelines [22,24].
  • Cost reduction: By predicting corrosion rates and understanding the factors influencing corrosion, engineers can implement more targeted and cost-effective corrosion management practices, reducing maintenance and replacement costs [11,13,25].
  • Transport type: CAPEX and OPEX costs related to CO2 varies on the transportation type. Fluid regime, pressure, and temperature are very different between CO2 transport by train, truck, or pipeline. Here, corrosion modeling helps engineers select the appropriate type of transportation [25,26,27].
The area of corrosion modeling has been investigated since 1975 [28], while corrosion modeling using neural networks has been studied as early as 2001 [29], although recent focus on machine learning has substantially increased the research focus [30,31,32,33,34].
Different review studies on CO2 corrosion modeling have been conducted, presenting various corrosion models, including de Waard, Norsok M-506, Hydrocor, Corplus, Cassandra, KSC, Multicorp, ECE, Tulsa, ULL, OLI, and SweetCor [35,36,37,38]. They discuss the effect of impurities such as H2O, O2, SO2, H2S, NO2, and the effects of acids, alkali, and salts. Furthermore, the effects of pressure and temperature, electrolyte composition, electrochemical reactions, and the influence of corrosion products [13]. However, these studies do not consider the broader trends in CO2 corrosion modeling, which may lead to solution overlap or to the modeling approaches encountering the same limitations or challenges. In addition, although there is an increasing tendency to use machine learning for CO2 corrosion models, challenges related to the quality and quantity of data from corrosion experiments remain. Thus, the motivation and the contributions of this work are summarized as follows:
  • A historical overview of CO2 corrosion models and software tools, illustrating the major trends, with the recent popularity of machine learning approaches.
  • An overview of CO2 corrosion models with their respective strengths, assumptions, and limitations.
  • Identification of challenges in CO2 corrosion modeling, with emphasis on data-related problems when applying machine learning approaches to predict CO2 corrosion rate.
  • Proposed opportunities given these challenges, through gathering data and increasing the quality of new data by more unified experimental approaches.
This study does not present the mathematics or theoretical approach of the different models, as these have already been presented in various studies [39,40,41,42,43]. Instead, this study focuses on CCUS applications and the current trends in corrosion modeling.
The remainder of the study consists of Section 2, which discusses the factors that influence CO2 corrosion, including the presence of impurities and flow conditions. Section 3 then presents the various modeling approaches, including empirical, semi-empirical, and mechanistic models, along with a historical overview and the direction of corrosion modeling. Then, Section 3.4 presents the challenges and opportunities in CO2 corrosion modeling, before concluding in Section 4.

2. Properties of CO2 Corrosion

This section summarizes the properties of CO2 that can affect corrosion, which have been investigated in several studies previously [44,45,46,47,48] and reviewed by [7,11,13,20].
When discussing CO2 corrosion, it refers to the electrochemical process that occurs between CO2 and iron (Fe) in an aqueous environment. The CO2 dissolves in water (H2O) to form carbonic acid (H2CO3), which facilitates the oxidation of iron to its ferrous state (Fe2+). The process can be expressed as redox reactions:
C O 2 + H 2 O H 2 C O 3
F e + 2 H 2 C O 3 F e 2 + + H C O 3 + H 2
F e 2 + + C O 3 2 F e C O 3
In the following subsections, the properties that influence corrosion, including the phase of CO2, the presence of impurities, flow, material, and time after exposure to a corrosive environment, are summarized, along with their relevance to incorporating them into corrosion prediction models.

2.1. CO2 Phase Conditions

CO2 can exist in different phases, depending on the temperature and pressure conditions, including solid, gaseous, liquid, and supercritical phases [8,49]. The supercritical phase occurs when CO2 surpasses its critical temperature and pressure, which is 30.98 °C and 73.77 bar [49]. However, the critical condition of CO2 can be impacted by impurities, such as the non-condensable gases (CO, H2, O2, and N2) [50,51]. For corrosion, different phases can have significant differences in corrosion rate, which are important when modeling corrosion.
The phase of matter of CO2 is typically chosen based on the transportation mode, which generally is within its gaseous, liquid, or supercritical phase [52,53]. CO2 can be transported by pipeline in the gaseous, dense, or supercritical phase, while stationary transportation modes such as truck, ship, and train generally are in the liquid phase [13,53,54,55]. For pipeline transportation, CO2 is first compressed, raising its pressure and temperature, pushing it into the supercritical state. Along the pipeline, where ambient temperatures are generally 5–20 °C, the CO2 transitions from supercritical to dense phase as it cools [13,55]. The exact phase depends on both the operating pressure and whether the system is designed for dense phase or gaseous transportation [55]. For stationary transportation, CO2 is maintained in the liquid phase at approximately −28–−50 °C and 7–18 bar, with no continuous flow during storage or transfer [56].
The amount of CO2 that can be packed into a given space depends heavily on its phase. As a gas at standard pressure and room temperature, its density is relatively low at around 1.8 kg/m3; however, when cooled and pressurized into a liquid, its density can reach about 1150 kg/m3 [57,58]. Transporting CO2 in liquid or supercritical form is usually more efficient because it allows much more mass to be transported in the same volume. This has a significant impact on transport costs and system design, especially since changes in phase affect energy use and can lead to operational challenges, such as phase separation [59,60,61]. In addition, supercritical CO2 has a higher solubility for H2O, which means that the CO2 can contain a greater amount of H2O before water droplets form, which would lead to an accelerated corrosion rate [62]. Phase behavior of impure CO2 has been studied by Sun et al. [63], where H2S was found to increase water precipitation by reducing the H2O solubility in CO2.
When modeling the corrosion rate, it is essential to consider the phase of CO2, as a change in phase can alter the solubility of impurities such as H2O, H2SO4, and HNO3 [11,50]. Pressure and temperature influence the corrosion rate but also affect the CO2 phase and the solubility of H2O and corrosive species, such as HNO3 and H2SO4. Corrosion tends to increase between 80 and 100 °C, but at temperatures above 120–150 °C, it may decrease due to protective surface layers forming [64]. As Brown et al. [65] showed, an increase in temperature from 4 to 50 °C at 10 MPa resulted in a corresponding increase in the corrosion rate, from 0.017 to 0.088 mm/y. Additionally, higher pressure accelerates corrosion by enhancing internal stress and increasing chemical reactivity [19,64].

2.2. Presence of Impurities

This section provides a brief overview of the impact of impurities present in CO2, as other studies have already conducted in-depth research on this, as well as the chemical reactions and reaction products that the impurities can form [11,50,66]. The types of impurities present in CO2 depend not only on the source of the CO2 but also on the capture process [6,62]. These can include H2O, O2, SO2, NO2, and H2S. The impurities can be roughly divided into two groups: corrosion-inducing impurities and non-condensable gases [50]. The non-condensable gases, such as Ar, N2, H2, and CH4, reduce the density while also increasing the pressure loss, resulting in greater energy requirements for compression [59]. In particular, H2 increases the vapor pressure of CO2, which is the pressure where a liquid tends to evaporate at a given temperature. Even at concentrations as low as 500 ppm H2, can significantly increase the vapor pressure, thereby reducing the efficiency of liquefaction and compression [50,58]. The solubility of H2O in CO2 is influenced by the presence of non-condensable gases, where the presence of 2.5 mol% O2, N2, or CH4 decreases the H2O solubility in dense and supercritical CO2 by 75 to 200 ppm [50,67].
The corrosion-inducing impurities, such as H2S, SO2, NO2, O2, and H2O, significantly increase the material costs of the pipeline, due to a need for a higher material grade to withstand corrosion [16]. The presence of water is a prerequisite for electrochemical corrosion processes [11]. Even trace amounts of water increase the possibility of corrosion [11,50]. Studies have examined the impact of water content in CO2, spanning a wide range from 50 ppm to saturated levels of around 50,000 ppm, as reviewed by Simonsen et al. [13]. Impurity type and concentration, particularly H2O, NO2, H2S, and SO2, have a strong impact on the corrosion severity. These compounds can interact to form strong acids like H2SO4 and HNO3, which aggressively oxidize exposed materials [12,21]. Among these, water is especially problematic due to its role as a reactant in most impurity-driven reactions. Corrosion risk remains relatively low when water is kept below its solubility limit in CO2 [68].
In addition, amines such as monoethanolamine (MEA) or methyldiethanolamine (MDEA) can be present in the CO2 after flue gas treatment [50]. When these amines are combined with CO2 and H2O, they can create a corrosive environment. However, at low amine concentrations, they have been reported to inhibit corrosion, suggesting that their influence on corrosion depends on concentration [19,50,69].
The presence of multiple impurities has been shown to amplify corrosion through synergistic interactions. For example, NO2 accelerates reactions between other impurities, such as the conversion of SO2 and H2O into sulfuric acid [11,45,50,70]. Even a few ppm of NO2 can initiate reactions that can lead to the formation of acid [50]. Similarly, H2S can react with oxidants like O2, SO2, or NO2 to generate water, which in turn increases the risk of condensation and enhances corrosive effects [65]. At concentrations exceeding 5 ppm, H2S can also promote acid drop out depending on the impurity composition [50]. Likewise, the presence of both O2 and H2S will lead to H2 oxidation, producing SO2 and elemental sulfur, and causing additional pitting corrosion [11,14,20,50].

2.3. Flow Conditions

Flow conditions also affect CO2 corrosion of pipelines. Typically, pipeline flow velocities range from 0.9 to 2.6 m/s [53]. Flow velocity affects the exchange of reaction products at the metal surface. Higher flow velocities increase corrosion because they remove protective layers of reaction products and bring fresh corrosive species to the steel surface [71]. This keeps the metal exposed and active, leading to higher corrosion. Once a stable protective layer forms, however, the effect of flow velocity decreases because the layer protects the steel and reduces the impact of fluid movement [20]. Flow conditions can significantly promote erosion, thereby increasing corrosion rates, as observed by Dugstad et al. [17] and Xiang et al. [72], where a higher flow rate corresponded to a higher corrosion rate. In addition, the presence of solid particles introduces further risks related to erosion, equipment damage, and reduced injectivity [50]. Moreover, CO2 multiphase flow can increase corrosion and erosion mechanisms due to phase interactions and localized shear stresses [30,73].

2.4. Materials

Material selection is important in ensuring durability and performance, as different materials exhibit varying levels of resistance to the environment. For example, carbon steel is a cost-effective option but offers limited corrosion resistance [16,23]. Enhancing the material by alloying with chromium (Cr) and nickel (Ni) significantly improves its resistance to corrosion but comes with an increased cost [20,23]. In addition to material selection, certain corrosion products can form a stable, protective layer on the surface, which helps decrease or prevent corrosion of the underlying material [11,20]. The corrosion product formed on carbon steel material during CO2 exposure with impurities is typically FeCO3; however, it depends on the other impurities. For instance, if SO2, O2, and H2O are present, the main corrosion product formed is FeSO3 [16,19,74,75,76]. On the other hand, at water saturated conditions and O2 concentration of <1000 ppm, FeSO4 was the main corrosion product observed [16]. Materials with higher content of mangan (Mn) and silicium (Si) are known to enhance the corrosion resistance, by forming more protective and stable corrosion films, where the formation of FeCO3 is protective for the material, causing a decrease in corrosion rate [16,77]. In addition, the application of corrosion-resistant coatings provides an effective strategy to further improve the material against corrosion [78]. While corrosion-resistant alloys offer improved tolerance to impurities in CO2, their higher material and fabrication costs limit large-scale deployment in CCUS infrastructure. Therefore, carbon steel remains the preferred material for CO2 transportation due to its lower cost [25,78].

2.5. Time

The exposure time of the experimental conditions affects the calculated corrosion rate, based on the material loss, as shown in Equation (4) [79,80].
C R = K · W A · T · D ,
where K is a constant, W is the mass loss in grams, A, is the exposed area, T is the time of exposure in hours, and D is the density of the material exposed in g/cm3. A corrosion experiment based on mass loss can give a misleading impression of the corrosion rate; the accumulation of corrosion products on the metal surface can lead to a decline in the corrosion rate, due to acting as a passivating layer that limits interaction between the metal and the corrosive environment [25]. This was observed by Elgaddafi et al [81], who investigated the corrosion rate using linear polarization resistance (LPR), rather than the mass loss method. LPR offers the advantage of real-time observation of the corrosion behavior throughout the experiment [82]. In their study, the corrosion rate reached its maximum after ∼25 h of exposure to the corrosive environment, and then the corrosion rate decreased until the end of the experiment. The authors attributed this reduction in corrosion rate over time to the formation of a passivating layer on the metal’s surface, which inhibited further corrosion. Another study by Xiang et al. [72] conducted supercritical CO2 corrosion testing using the mass loss method over exposure times of 1, 3, 5, 10, and 20 days under stagnant conditions in the presence of H2O, SO2, NO2, and O2. The results showed a very high initial corrosion rate of ∼17 mm/y after the first day of exposure, followed by a sharp decrease between 1 and 3 days. Beyond this period, the corrosion rate continued to decline more gradually with increasing exposure time, reaching a considerably lower value after 20 days. This behavior was attributed to the formation of a protective surface scale on the material.

3. Modeling Approaches

Modeling methodologies can be categorized into three main types: mechanistic, semi-empirical, and empirical [83]. The three model types are complementary, such that empirical models enable fast input/output correlation; semi-empirical models add knowledge of behavior to reduce the number of tuning parameters to provide better predictions outside of the training data range; and mechanistic models provide explanatory power and improved extrapolation when chemistry, flow, and film formation are coupled and time-dependent.
Several studies have provided a review of CO2 corrosion modeling, including historical progression and modeling types [39,83,84,85,86]. This paper extends these review studies by summarizing the empirical, semi-empirical, and mechanistic CO2 model types, while focusing on the recent machine learning (ML) models.
In this work, the model types are defined as follows:
  • Empirical model type is defined by arbitrary mathematical expressions, where parameters are found by fitting the expression to corrosion data.
  • Semi-empirical model type is defined by a mathematical expression informed by at least some electrochemical or physical principles, where parameters are found by fitting the expression to corrosion data.
  • Mechanistic model type is defined by electrochemical and physics-based equations that generally require more inputs but little to no experimental data.
The literature generally agrees on the broad definition; however, semi-empirical is sometimes sub-grouped into empirical. An example of this is that other works refer to de Waard’s early works and Norsok M506 as either empirical [87,88] or semi-empirical [37,40,84,89] or mechanistic [84].
This difference in model type classifications results from different definitions of what is meant by semi-empirical and mechanistic. In some works, a model is mechanistic if some of its structure is merely inspired by physics, whereas in other works, a model has to have elaborate electrochemical and physics-based equations.
To visualize the progression of CO2 corrosion models and software tools, a historical timeline is constructed in Figure 1, summarizing the development of key models since 1975. This timeline captures the transition from early empirical models based on correlations from experimental data to more sophisticated semi-empirical approaches and, ultimately, to fully mechanistic and hybrid models. Each model is positioned chronologically to illustrate how earlier work, such as the foundational de Waard–Milliams model, provided the basis for subsequent models. In addition to Figure 1, Table 1 compares selected corrosion models, including Freecorp, OLI, Predict, NORSOK, and de Waard.

3.1. Empirical Models

Empirical models correlate measured corrosion rates to operating variables, such as temperature, CO2 partial pressure, flow, pH, impurity concentrations, with mathematical functions [37,89]. These mathematical functions are arbitrary and do not correspond to any real physical meaning [37,39]. Additionally, the models are also limited to the experimental conditions used in the calibration, which reduces the accuracy when extrapolating, due to the absence of theoretical principles [37,39]. Examples of empirical models include the Copra, Dugstad’s earlier work, and Sweetcor, as illustrated in Figure 1 and Table 1. Sweetcor by Shell (1998) uses a large laboratory and field database of steel corrosion in water and CO2, grouping data by ranges of CO2 partial pressure, temperature, and flow conditions [20,90,101].
Recent approaches to empirical modeling have expanded to use machine learning techniques. The premise of these methods is to utilize large datasets to identify complex, non-linear relationships between variables without requiring explicit mechanistic assumptions [31,32,33,34]. The following subsection explores machine learning models in greater detail.

Machine Learning

The complexity of CO2 corrosion, especially under CCUS conditions, has motivated the use of ML as a complementary approach to traditional models [31,32,33,34]. ML models learn patterns directly from data, unlike empirical or mechanistic models, making them well-suited for systems influenced by multiple interacting variables such as temperature, pressure, impurities, flow conditions, and inhibitor performance. Table 2 shows a comparison between traditional model types and ML models for CO2 corrosion prediction.
The idea of using ML for corrosion prediction is not new, as one of the earliest applications was by Nesic et al., in 2001, who employed neural networks to model CO2 corrosion laboratory data [29]. They demonstrated that data-driven approaches could capture complex relationships in corrosion processes, even when mechanistic understanding was incomplete.
Traditional models often struggle with non-linear interactions and data variability, particularly when impurities and phase changes are involved [11,30,37,64]. ML has the potential to capture these relationships without requiring explicit assumptions about reaction mechanisms. For predicting corrosion using ML, several different methods have been used, such as random forest (RF), AdaBoost, CatBoost, XGBoost, decision tree (DT), multilayer perceptron (MLP), support vector regression (SVR), and K-nearest neighbors (KNN) [31,32,33].
Modern ML applications in CO2 corrosion modeling have expanded significantly. Aghaaminiha et al. [32] applied supervised ML algorithms, including RF, Artificial Neural Networks (ANNs), SVR, and KNN, to model time-dependent corrosion rates of carbon steel in the presence of inhibitors. RF achieved the best performance, predicting the entire corrosion-rate profiles with mean squared errors as low as 0.002, and provided useful sensitivity analyses for variables such as CO2 partial pressure and temperature [32]. However, the authors emphasized the risk of extrapolation beyond the training domain and the need for expert oversight when interpreting ML predictions.
Recent studies have extended ML applications to CO2 corrosion in multiphase pipelines and CCUS environments. Dong et al. [31] compared RF, XGBoost, and LightGBM for corrosion rate prediction and reported R2 values above 0.9, outperforming traditional semi-empirical models. Beyond purely data-driven regressors, some hybrid approaches now integrate physics explicitly by using computational fluid dynamics (CFD) or transport models or first-principles electrochemical simulators to generate training data and features, or by embedding physics constraints within the learning algorithm, achieving higher computational speeds over high-fidelity solvers while maintaining accuracy [34].
Qu et al. [108] employed principal component analysis (PCA), particle swarm optimization (PSO), and SVR. The authors concluded that the PCA identified the key parameters influencing the corrosion rate while reducing redundant information. The PCA-PSO-SVR model demonstrated higher accuracy and robustness, indicating its potential application in corrosion protection of offshore pipelines.
ML models require large, high-quality datasets, which are often inconsistent in corrosion research [13,20,68]. They also lack inherent physical interpretability and can fail when applied outside their training range [32]. Future work should explore the potential of physics-informed ML and use more standardized experimental protocols to improve data quality. Machine learning is not a replacement for mechanistic understanding but a complementary tool that can accelerate model development and decision-making in CCUS applications.
An issue with using methods such as deep learning is that the available data is not comprehensive enough [33]. However, if more standardized data become available, the applicability of ML-based models is expected to increase.

3.2. Semi-Empirical Models

In contrast to empirical models, semi-empirical models include physics-informed equation structures, from either experience with CO2 corrosion or physics-based equations with unknown parameters [37]. Additionally, like the empirical models, the semi-empirical models are limited to their calibration range, and extrapolation may lead to physically unrealistic predictions [37,39].
The early de Waard Milliams model, which is also the first semi-empirical CO2 corrosion model, was based on the corrosion current on a linear correlation of pH [20,28,37,40]. They subsequently adapted to relate the corrosion rate to CO2 partial pressure and temperature by implementing the cathodic reaction of hydrogen evolution, a relationship that continues to be applied in current practice [37,84]. Later, versions of the model by de Waard et al. accounted for effects such as flow rate, protective scales, glycol, pH, and the microstructure of the steel [28,37,99,104,105]. These additional effects were introduced by implementing empirical correction factors to the original de Waard–Milliams model.
Several companies have developed software tools based on semi-empirical models to predict CO2 corrosion rates. These semi-empirical software tools include Cormed, Corpos, Predict, Corplus, and early Freecorp. Cormed by Elf (1989) started as a semi-empirical model, but was later merged with Total’s Lipucor, which resulted in Corplus [20,36].
Freecorp 1.0 (2008), from Ohio University, was exclusively based on publicly available data, with all its code, including equations, assumptions, and limitations, available to its users. The first version of Freecorp used a semi-empirical structure to predict the uniform corrosion rate of carbon steel [91]. Later versions of Freecorp shifted into mechanistic modeling.

3.3. Mechanistic Models

Mechanistic models describe CO2 corrosion by explicitly formulating the related electrochemical and transport processes rather than relying on empirical correlations [37,84,109,110,111]. The mechanistic modeling approach offers interpretability and allows better extrapolation outside the range of calibration [83,84].
Kahyarian et al. [83] divided mechanistic models into two main categories: elementary mechanistic models, focusing on basic electrochemical reactions, and comprehensive mechanistic models, which include multiple processes such as mass transport, surface film formation, and flow effects. The comprehensive mechanistic models are considerably more complex to develop and apply, which is why simpler elementary mechanistic models remain the preferred choice among corrosion engineers and academic researchers [83]. The theoretical background for the inputs for mechanistic models has been extensively examined in prior research [37,39,84,109,110,111].
Examples of mechanistic models include the Freecorp 2.0 model developed by Nesic et al. [97], which accounts for multiple species, including H2S and organic acids. It allows users to include or exclude specific reactions, providing insight into the contribution of each species to the overall corrosion rate. Freecorp 2.0 improves upon version 1.0 by replacing empirical predictions of formation and protection of corrosion layers with a mechanistic approach [97].
Another example is the Hydrocor model developed by Shell, which estimates CO2 corrosion in aqueous systems by using electrochemical kinetics, mass transport, and protective layer formation under varying pressure and temperature conditions [20,101].
Ref. [95] propose a modeling approach for predicting corrosion rates in supercritical CO2 pipelines, under thin water film conditions. The model divides the process into six layers: the supercritical CO2 mixture, the interface region, the water film region, the deposition region, the electrodic region, and finally the pipeline steel itself. Mathematically, it is prepared by setting initial and boundary conditions, and then solving several partial differential equations. The authors refer to this model as mechanistic, as it uses physics-based equations, even though it requires empirical coefficients to be known.

3.4. Challenges and Opportunities

Since 1975, multiple companies and authors have advanced corrosion modeling by extending or modifying previous models, as seen in Figure 1, or contributing datasets for training. The result is a broad range of models, some of which are created to suit specific ranges for parameters such as temperature and pressure. The modeling approach needs to be more advanced than linear regression, as there are many non-linear behaviors between the input parameters, such as the combination effects of the concentration of various impurities, temperature, pressure, pH, and time [11,16,37,97,112].
An increased number of recent studies have used machine learning for corrosion prediction [30,31,32,33,34,108]. While machine learning approaches have great potential for CO2 corrosion modeling, they generally require a large amount of data to produce a reasonably accurate model for CO2 corrosion, but the needed data are normally not available [31,32].
For corrosion prediction software tools, it is also common for the underlying data to be restricted due to competitive or legal reasons or a strategic choice of the owners [31,32]. However, for training models, it is a barrier, as it limits the quantity of publicly available data for training. The study by Qu et al. [108] presented a hybrid model, along with the 178 data points used to train and validate the model, which can be used by others. The data points had 14 different input parameters, such as CO2 partial pressure, temperature, pH, flow rate, and ion concentration. Data contributions are valuable for future machine learning approaches to reach the data quantities where the methods are effective [113].
Data already publicly available from various studies investigating corrosion in CO2 with impurities presents significant challenges due to inconsistencies in experimental procedures across studies [13]. The result of this problem is varying data quality, which is known to cause poorly performing data-driven models [14]. Initiatives to reduce the inconsistencies are described in [13]. Here, suggestions are presented for the experimental procedures, which include the following:
  • Follow the standard cleaning process of corrosion coupons.
  • Verify concentration of impurities.
  • State flow conditions, rather than flow versus no flow.
  • Consider obtaining samples with different exposure times.
In addition to addressing data quantity and quality challenges, a promising possibility is to use a hybrid ML approach, as demonstrated in [34], where the ML model structure is informed by physics to reduce the data requirements. Generally, the mechanistic knowledge can help provide boundaries for complexity to make sure the model has an advanced enough structure to capture the corrosion behavior, but not too advanced a structure that it is prone to overfitting, which is especially a challenge in the field of CO2 corrosion, where public data are limited.
From an engineering perspective, current approaches benefit from increasing methodological maturity. However, their practical deployment is still constrained by limited quality field data, challenges in transferability across operating conditions, and uncertainty under different CO2 compositions and flow regimes. These limitations highlight clear research opportunities in hybrid physics-informed models, standardized datasets, and validation under realistic pipeline and storage conditions.

4. Conclusions

This study examined the current state of CO2 corrosion modeling with a focus on its relevance to CCUS applications. The motivation for this work originates from the growing global emphasis on CO2 sequestration and the need to ensure the integrity of transport and storage infrastructure. For this infrastructure, impure CO2-induced corrosion represents a challenge.
Key factors influencing CO2 corrosion have been summarized, including the presence of impurities such as H2O, O2, SOx, NOx, and H2S. This work also summarizes the effect of transportation mode, such as transportation by pipeline or truck, material selection, exposure time, CO2 phase, and flow regime, to provide an overview. Empirical, semi-empirical, and mechanistic models were compared, with an emphasis on their respective advantages and limitations.
This study illustrates the progression of CO2 corrosion models and software tools. The historical timeline provided in Figure 1 illustrates how models, such as de Waard-Milliams, inspired subsequent developments, including ECE, Predict, Norsok, and Corpos. This chronological perspective highlights the gradual shift from semi-empirical correlations to physics-based mechanistic models and, more recently, to data-driven machine learning models. One of the key contributions is the comparative table (Table 1), which summarizes CO2 corrosion models and software tools, supports engineers and researchers in making informed selections and applications of models under specific operating conditions. This work also highlights challenges, including inconsistencies in experimental data, and the need to account for multiple interacting factors, such as impurities, flow regime, and temperature. For data-driven CO2 corrosion models, this challenge involves selecting which inputs to include and preparing the training and validation datasets. As the typical datasets originate from a broad range of researchers, the conditions stated in the experiments vary. Therefore, future opportunities include the development of standardized experimental protocols to reduce data variability. Additionally, integrating mechanistic principles into ML models to improve interpretability and predictive accuracy, enabling robust corrosion management strategies for CCUS pipelines and storage systems.
CO2 corrosion modeling remains a dynamic research field at the intersection of materials science, electrochemistry, and computational modeling. The progress summarized in this work provides an overview of current research trends that aim to improve corrosion predictions through ML approaches supported or informed by physical and chemical behavior. The continued advance in CO2 corrosion modeling ultimately supports the CCUS industry by enabling more informed material selection, defining purification demands, and predicting maintenance needs. The recent machine learning approaches further call for more open CO2 corrosion data availability with shared interests instead of competitive and restrictive commercial interests.

Author Contributions

Conceptualization, K.R.S., M.O., M.Z., S.P. and M.V.B.; methodology, K.R.S., M.O. and M.V.B.; investigation, K.R.S. and M.V.B.; data curation, K.R.S. and M.V.B.; writing—original draft preparation, K.R.S., M.O. and M.V.B.; writing—review and editing, K.R.S., M.O., S.P., and M.V.B.; visualization, M.V.B.; supervision, S.P.; project administration, S.P.; funding acquisition, S.P. All authors have read and agreed to the published version of the manuscript.

Funding

Innovation Fund Denmark supported the research through the 3-P6 CarbonAdapt project.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ANNArtificial neural networks
CCUSCarbon capture, utilization, and storage
CFDComputational fluid dynamics
COCarbon monoxide
CO2Carbon dioxide
CrChromium
DTDecision tree
EOREnhanced oil recovery
FeIron
Fe2+Ferrous ion (iron in +2 oxidation state)
FeCO3Iron carbonate
FeSO3Iron sulfite
FeSO4Iron sulfate
H2Hydrogen
H2CO3Carbonic acid
H2OWater
H2SHydrogen sulfide
KNNK-nearest neighbors
LPRLinear polarization resistance
MEAMonoethanolamine
MDEAMethyldiethanolamine
MLMachine learning
MLP      Multilayer perceptron
MnMangan
N2Nitrogen
NiNickel
NOxNitrogen oxides (e.g., NO, NO2)
O2Oxygen
PCAPrincipal component analysis
PSOParticle swarm optimization
RFRandom forest
SiSilicon
SOxSulfur oxides (e.g., SO2, SO3)
SVRSupport vector regression

References

  1. Masson-Delmotte, V.; Zhai, P.; Pörtner, H.O.; Roberts, D.; Skea, J.; Shukla, P.R.; Pirani, A.; Moufouma-Okia, W.; Péan, C.; Pidcock, R.; et al. Global Warming of 1.5 °C. An IPCC Special Report on the Impacts of Global Warming of 1.5 °C above Pre-Industrial Levels and Related Global Greenhouse Gas Emission Pathways, in the Context of Strengthening the Global Response to the Threat of Climate Change, Sustainable Development, and Efforts to Eradicate Poverty; Technical Report; IPCC: Geneva, Switzerland, 2019. [Google Scholar]
  2. IPCC. Carbon Dioxide Capture and Storage; Cambridge University Press: Cambridge, UK, 2005. [Google Scholar]
  3. Wu, J.; Ansari, U. From CO2 Sequestration to Hydrogen Storage: Further Utilization of Depleted Gas Reservoirs. Reserv. Sci. 2025, 1, 19–35. [Google Scholar] [CrossRef]
  4. Osman, A.I.; Hefny, M.; Abdel Maksoud, M.I.; Elgarahy, A.M.; Rooney, D.W. Recent Advances in Carbon Capture Storage and Utilisation Technologies: A Review. Environ. Chem. Lett. 2021. [Google Scholar] [CrossRef]
  5. Orchard, K.; Hay, M.; Ombudstvedt, I.; Skagestad, R.; Joos, M.; Nysæter, G.; Sjøbris, C.; Gimnes Jarøy, A.; Durusut, E.; Craig, J. The Status and Challenges of CO2 Shipping Infrastructures. Ssrn Electron. J. 2021. [Google Scholar] [CrossRef]
  6. Porter, R.T.; Fairweather, M.; Pourkashanian, M.; Woolley, R.M. The range and level of impurities in CO2 streams from different carbon capture sources. Int. J. Greenh. Gas Control 2015, 36, 161–174. [Google Scholar] [CrossRef]
  7. Luna-Ortiz, E.; Yao, C.; Barnes, J.; Winter, M.; Healey, M. Development of a CO2 Specification for Industrial CCS Transport Networks: Methodology, Limitations and Opportunities. In Proceedings of the Annual Offshore Technology Conference, Houston, TX, USA, 2–5 May 2022. [Google Scholar] [CrossRef]
  8. Simonsen, K.R.; Hansen, D.S.; Pedersen, S. Challenges in CO2 transportation: Trends and perspectives. Renew. Sustain. Energy Rev. 2024, 191, 114149. [Google Scholar] [CrossRef]
  9. Vitali, M.; Corvaro, F.; Marchetti, B.; Terenzi, A. Thermodynamic challenges for CO2 pipelines design: A critical review on the effects of impurities, water content, and low temperature. Int. J. Greenh. Gas Control 2022, 114, 103605. [Google Scholar] [CrossRef]
  10. Loria, P.; Bright, M.B. Lessons captured from 50 years of CCS projects. Electr. J. 2021, 34, 106998. [Google Scholar] [CrossRef]
  11. Sonke, J.; Morland, B.H.; Moulie, G.; Franke, M.S. Corrosion and chemical reactions in impure CO2. Int. J. Greenh. Gas Control 2024, 133, 104075. [Google Scholar] [CrossRef]
  12. Morland, B.H.; Tjelta, M.; Dugstad, A.; Svenningsen, G. Corrosion in CO2 systems with impurities creating strong acids. Corrosion 2019, 75, 1307–1314. [Google Scholar] [CrossRef]
  13. Simonsen, K.R.; Hansen, D.S.; Pedersen, S. Experimental Approaches for Emulating CO2 Transportation. In Proceedings of the 17th Greenhouse Gas Control Technologies Conference (GHGT-17), Calgary, AB, Canada, 20–24 October 2024. [Google Scholar] [CrossRef]
  14. Sun, H.; Wang, H.; Zeng, Y.; Liu, J. Corrosion challenges in supercritical CO2 transportation, storage, and utilization—A review. Renew. Sustain. Energy Rev. 2023, 179, 113292. [Google Scholar] [CrossRef]
  15. Gao, K.; Shang, S.; Zhang, Z.; Gao, Q.; Ma, J.; Liu, W. Effect of Temperature on Corrosion Behavior and Mechanism of S135 and G105 Steels in CO2/H2S Coexisting System. Metals 2022, 12, 1848. [Google Scholar] [CrossRef]
  16. Simonsen, K.R.; Goebel, J.; Hansen, D.S.; Pedersen, S. The influence of temperature, H2O, and NO2 on corrosion in CO2 transportation pipelines. Process Saf. Environ. Prot. 2025, 198, 107190. [Google Scholar] [CrossRef]
  17. Dugstad, A.; Morland, B.; Clausen, S. Corrosion of transport pipelines for CO2—Effect of water ingress. Energy Procedia 2011, 4, 3063–3070. [Google Scholar] [CrossRef]
  18. Morland, B.H.; Norby, T.; Tjelta, M.; Svenningsen, G. Effect of SO2, O2, NO2, and H2O concentrations on chemical reactions and corrosion of carbon steel in dense phase CO2. Corrosion 2019, 75, 1327–1338. [Google Scholar] [CrossRef]
  19. Barker, R.; Hua, Y.; Neville, A. Internal corrosion of carbon steel pipelines for dense-phase CO2 transport in carbon capture and storage (CCS)—A review. Sage J. 2017, 62, 1176306. [Google Scholar] [CrossRef]
  20. Xiang, Y.; Xu, M.; Choi, Y.S. State-of-the-art overview of pipeline steel corrosion in impure dense CO2 for CCS transportation: Mechanisms and models. Sage J. 2017, 62, 1304690. [Google Scholar] [CrossRef]
  21. Morland, B.H.; Svenningsen, G.; Dugstad, A. The challenge of monitoring impurity content of CO2 streams. Processes 2021, 9, 570. [Google Scholar] [CrossRef]
  22. Yan, Y.; Lyu, W.; Yu, H.; Lv, W.; Wei, K.; Jiang, L. Advances in Synergistic Corrosion Mechanisms of and Management Strategies for Impurity Gases During Supercritical CO2 Pipeline Transportation. Molecules 2025, 30, 4094. [Google Scholar] [CrossRef] [PubMed]
  23. Sonke, J.; Bos, W.; Paterson, S. Materials challenges with CO2 transport and injection for carbon capture and storage. Int. J. Greenh. Gas Control 2022, 114, 103601. [Google Scholar] [CrossRef]
  24. Yu, X.; Liu, G.; Rao, S.; Fan, X.; Li, Y.; Hu, Q.; Wang, C.; Zhang, L.; Chen, Z. A critical review on corrosion challenges and prospects in supercritical offshore CO2 pipelines. J. Pipeline Sci. Eng. 2025, 5, 100401. [Google Scholar] [CrossRef]
  25. Simonsen, K.R. CO2 Transportation in CCUS: Impurity Monitoring and Corrosion Challenges; Aalborg University: Aalborg, Denmark, 2025. [Google Scholar] [CrossRef]
  26. Zero Emissions Platform (ZEP). The Costs of CO2 Transport Post-Demonstration CCS in the EU; Technical Report; Global CCS Institute: Melbourne, Australia, 2011; Available online: https://zeroemissionsplatform.eu/publication/the-costs-of-co2-transport/ (accessed on 30 October 2025).
  27. Doctor, R.; Palmer, A.; Coleman, D.; Davison, J.; Hendriks, C.; Kaarstad, O.; Ozaki, M.; Austell, M. Transport of CO2. In IPCC Special Report on Carbon dioxide Capture and Storage; Chapter 4; Cambridge University Press: Cambridge, UK, 2005; pp. 179–194. [Google Scholar]
  28. de Waard, C.; Milliams, D.E. Carbonic Acid Corrosion of Steel. Corrosion 1975, 31, 177–181. [Google Scholar] [CrossRef]
  29. Nesic, S.; Nordsveen, M.; Maxwell, N.; Vrhovac, M. Probabilistic modelling of CO2 corrosion laboratory data using neural networks. Corros. Sci. 2001, 43, 1373–1392. [Google Scholar] [CrossRef]
  30. Wang, G.; Wang, C.; Shi, L. CO2 Corrosion Rate Prediction for Submarine Multiphase Flow Pipelines Based on Multi-Layer Perceptron. Atmosphere 2022, 13, 1833. [Google Scholar] [CrossRef]
  31. Dong, Z.; Zhang, M.; Li, W.; Wen, F.; Dong, G.; Zou, L.; Zhang, Y. Development of a Predictive Model for Carbon Dioxide Corrosion Rate and Severity Based on Machine Learning Algorithms. Materials 2024, 17, 4046. [Google Scholar] [CrossRef]
  32. Aghaaminiha, M.; Mehrani, R.; Colahan, M.; Brown, B.; Singer, M.; Nesic, S.; Vargas, S.M.; Sharma, S. Machine learning modeling of time-dependent corrosion rates of carbon steel in presence of corrosion inhibitors. Corros. Sci. 2021, 193, 109904. [Google Scholar] [CrossRef]
  33. Cui, J. Deep learning-based modeling of CO2 corrosion rate prediction in oil and gas pipelines. Appl. Math. Nonlinear Sci. 2025, 10. [Google Scholar] [CrossRef]
  34. Yang, H.; Lu, L.; Tsai, K.; Sidahmed, M. A Hybrid Physics and Active Learning Model For CFD-Based Pipeline CO2 and O2 Corrosion Prediction. In Proceedings of the International Petroleum Technology Conference, Bangkok, Thailand, 1–3 March 2023. [Google Scholar] [CrossRef]
  35. Nyborg, R. CO2 corrosion models for oil and gas production systems. In Proceedings of the Nace Corrosion, NACE, San Antonio, TX, USA, 14–16 March 2010; p. 10371. [Google Scholar]
  36. Nyborg, R. Guidelines for Prediction of CO2 Corrosion in Oil and Gas Production Systems; Technical Report; Institute for Energy Technology (IFE): Kjeller, Norway, 2009. [Google Scholar]
  37. Kahyarian, A.; Singer, M.; Nesic, S. Modeling of uniform CO2 corrosion of mild steel in gas transportation systems: A review. J. Nat. Gas Sci. Eng. 2016, 29, 530–549. [Google Scholar] [CrossRef]
  38. Cui, G.; Yang, Z.; Liu, J.; Li, Z. A comprehensive review of metal corrosion in a supercritical CO2 environment. Int. J. Greenh. Gas Control 2019, 90, 102814. [Google Scholar] [CrossRef]
  39. Nesic, S.; Postlethwaite, J.; Vrhovac, M. CO2 Corrosion of Carbon Steel—From Mechanistic to Empirical Modelling. Corros. Rev. 1997, 15, 211–240. [Google Scholar] [CrossRef]
  40. Fosbøl, P.L. Carbon Dioxide Corrosion: Modelling and Experimental Work Applied to Natural Gas Pipelines. Ph.D. Thesis, IVC-SEP Department of Chemical Engineering Technical University of Denmark, Kongens Lyngby, Denmark, 2008. [Google Scholar]
  41. Fontes, E.; Nistad, B. Modeling Corrosion and Corrosion Protection. Comsol White Pap. Ser. 2019, 4, 1–11. [Google Scholar]
  42. Fonseca, D.; Tagliari, M.; Guaglianoni, W.; Tamborim, S.; Borges, M. Carbon Dioxide Corrosion Mechanisms: Historical Development and Key Parameters of CO2-H2O Systems. Int. J. Corros. 2024, 2024, 5537767. [Google Scholar] [CrossRef]
  43. Bachega Cruz, J.P.; Veruz, E.G.; Aoki, I.V.; Schleder, A.M.; de Souza, G.F.M.; Vaz, G.L.; de Barros, L.O.; Orlowski, R.T.C.; Martins, M.R. Uniform corrosion assessment in oil and gas pipelines using corrosion prediction models—Part 1: Models performance and limitations for operational field cases. Process Saf. Environ. Prot. 2022, 167, 500–515. [Google Scholar] [CrossRef]
  44. Hua, Y.; Barker, R.; Neville, A. The effect of O2 content on the corrosion behaviour of X65 and 5Cr in water-containing supercritical CO2 environments. Appl. Surf. Sci. 2015, 356, 499–511. [Google Scholar] [CrossRef]
  45. Sun, C.; Sun, J.; Wang, Y.; Lin, X.; Li, X.; Cheng, X.; Liu, H. Synergistic effect of O2, H2S and SO2 impurities on the corrosion behavior of X65 steel in water-saturated supercritical CO2 system. Corros. Sci. 2016, 107, 193–203. [Google Scholar] [CrossRef]
  46. Xiang, Y.; Wang, Z.; Li, Z.; Ni, W.D. Effect of exposure time on the corrosion rates of X70 steel in supercritical CO2/SO2/O2/H2O environments. Corrosion 2013, 69, 251–258. [Google Scholar] [CrossRef] [PubMed]
  47. Hoa, L.Q.; Baessler, R.; Bettge, D. On the Corrosion Mechanism of CO2 Transport Pipeline Steel Caused by Condensate: Synergistic Effects of NO2 and SO2. Materials 2019, 12, 364. [Google Scholar] [CrossRef] [PubMed]
  48. Wu, G.; Li, C.; Vera, J.R.; Kovacs, W., III; Evans, K. Effect of CO on Stress Corrosion Cracking of Carbon Steel Pipelines in CCS Environments. In Proceedings of the AMPP Annual Conference + Expo, New Orleans, LO, USA, 3–7 March 2024. [Google Scholar] [CrossRef]
  49. Span, R.; Wagner, W.; Span, R.; Wagner, W. A New Equation of State for Carbon Dioxide Covering the Fluid Region from the Triple-Point Temperature to 1100 K at Pressures up to 800 MPa. J. Phys. Chem. Ref. Data 1996, 25, 1509–1596. [Google Scholar] [CrossRef]
  50. Simonsen, K.R.; Hansen, D.S.; Pedersen, S. Framework for CO2 impurity monitoring in CCUS infrastructure. Carbon Capture Sci. Technol. 2025, 16, 100453. [Google Scholar] [CrossRef]
  51. Luna Ortiz, E. CO2cktails in a pipeline’: The Phase Behaviour of CO2 with >20 impurities. In Proceedings of the 11th Trondheim Conference on Carbon Capture, Transport and Storage, Virtual Conference, 22–23 June 2021; pp. 10–18. [Google Scholar]
  52. Lu, H.; Ma, X.; Huang, K.; Fu, L.; Azimi, M. Carbon dioxide transport via pipelines: A systematic review. J. Clean. Prod. 2020, 266, 121994. [Google Scholar] [CrossRef]
  53. Cole, I.S.; Paterson, D.A.; Corrigan, P.; Sim, S.; Birbilis, N. State of the aqueous phase in liquid and supercritical CO2 as relevant to CCS pipelines. Int. J. Greenh. Gas Control 2012, 7, 82–88. [Google Scholar] [CrossRef]
  54. Equinor. Northern Lights FEED Report; Technical Report; Equinor: Stavanger, Norway, 2020. [Google Scholar]
  55. Peletiri, S.P.; Rahmanian, N.; Mujtaba, I.M. CO2 Pipeline design: A review. Energies 2018, 11, 2184. [Google Scholar] [CrossRef]
  56. Brescini, M.; Antomarioni, S.; Ciarapica, F.E.; Bevilacqua, M. Techno-economic and environmental assessment of carbon capture solutions in maritime transportation. Ocean. Eng. 2025, 330, 121252. [Google Scholar] [CrossRef]
  57. National Institute of Standards and Technology. Thermophysical Properties of Fluid Systems; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2025.
  58. Danish Energy Agency. Carbon Capture, Transport and Storage: Technology Descriptions and Projections for Long-Term Energy System Planning; Danish Energy Agency: København, Denmark, 2024; pp. 1–198. [Google Scholar]
  59. Bai, F.; Lu, Y. Effects of Impurities on Anthropogenic CO2 Pipeline Transport. Energy Fuels 2024, 38, 9958–9966. [Google Scholar] [CrossRef]
  60. Raju, D.; Ramdin, M.; Vlugt, T.J. Thermophysical Properties and Phase Behavior of CO2 with Impurities: Insight from Molecular Simulations. J. Chem. Eng. Data 2024, 69, 2735–2755. [Google Scholar] [CrossRef]
  61. Onyebuchi, V.E.; Kolios, A.; Hanak, D.P.; Biliyok, C.; Manovic, V. A systematic review of key challenges of CO2 transport via pipelines. Renew. Sustain. Energy Rev. 2018, 81, 2563–2583. [Google Scholar] [CrossRef]
  62. Cole, I.S.; Corrigan, P.; Sim, S.; Birbilis, N. Corrosion of pipelines used for CO2 transport in CCS: Is it a real problem? Int. J. Greenh. Gas Control 2011, 5, 749–756. [Google Scholar] [CrossRef]
  63. Sun, C.; Sun, J.; Luo, J.L. Unlocking the impurity-induced pipeline corrosion based on phase behavior of impure CO2 streams. Corros. Sci. 2020, 165, 108367. [Google Scholar] [CrossRef]
  64. Peng, S.; Zhang, Z.; Liu, E.; Liu, W.; Qiao, W. A new hybrid algorithm model for prediction of internal corrosion rate of multiphase pipeline. J. Nat. Gas Sci. Eng. 2021, 85, 103716. [Google Scholar] [CrossRef]
  65. Brown, J.; Graver, B.; Gulbrandsen, E.; Dugstad, A.; Morland, B. Update of DNV recommended practice RP-J202 with focus on CO2 corrosion with impurities. Energy Procedia 2014, 63, 2432–2441. [Google Scholar] [CrossRef]
  66. Morland, B.H.; Tjelta, M.; Norby, T.; Svenningsen, G. Acid reactions in hub systems consisting of separate non-reactive CO2 transport lines. Int. J. Greenh. Gas Control 2019, 87, 246–255. [Google Scholar] [CrossRef]
  67. Ahmad, M.; Gersen, S. Water Solubility in CO2 Mixtures: Experimental and Modelling Investigation. Energy Procedia 2014, 63, 2402–2411. [Google Scholar] [CrossRef]
  68. Kairy, S.K.; Zhou, S.; Turnbull, A.; Hinds, G. Corrosion of pipeline steel in dense phase CO2 containing impurities: A critical review of test methodologies. Corros. Sci. 2023, 214, 110986. [Google Scholar] [CrossRef]
  69. Zhao, F.; Cui, C.; Dong, S.; Xu, X.; Liu, H. An overview on the corrosion mechanisms and inhibition techniques for amine-based post-combustion carbon capture process. Sep. Purif. Technol. 2023, 304, 122091. [Google Scholar] [CrossRef]
  70. Halseid, M.; Dugstad, A.; Morland, B. Corrosion and bulk phase reactions in CO2 transport pipelines with impurities: Review of recent published studies. Energy Procedia 2014, 63, 2557–2569. [Google Scholar] [CrossRef]
  71. Choi, Y.S.; Colahan, M.; Nesic, S. Effect of Flow on the Corrosion Behavior of Pipeline Steel in Supercritical CO2 Environments with Impurities. Corrosion 2023, 79, 497–508. [Google Scholar] [CrossRef] [PubMed]
  72. Xiang, Y.; Li, C.; Hesitao, W.; Long, Z.; Yan, W. Understanding the pitting corrosion mechanism of pipeline steel in an impure supercritical CO2 environment. J. Supercrit. Fluids 2018, 138, 132–142. [Google Scholar] [CrossRef]
  73. Villarreal, J.; Laverde, D.; Fuentes, C. Carbon-steel corrosion in multiphase slug flow and CO2. Corros. Sci. 2006, 48, 2363–2379. [Google Scholar] [CrossRef]
  74. Ruhl, A.S.; Kranzmann, A. Investigation of Pipeline Corrosion in Pressurized CO2 Containing Impurities. Energy Procedia 2013, 37, 3131–3136. [Google Scholar] [CrossRef]
  75. Hua, Y.; Barker, R.; Neville, A. The influence of SO2 on the tolerable water content to avoid pipeline corrosion during the transportation of supercritical CO2. Int. J. Greenh. Gas Control 2015, 37, 412–423. [Google Scholar] [CrossRef]
  76. He, G.; Zou, Q.; Liao, K.; Leng, J.; Zhao, S. Corrosion mechanism of high temperature and O2 content in steamed CO2/O2/SO2 system and failure behavior of 20G steel on steam-injection pipelines. Process Saf. Environ. Prot. 2022, 163, 528–542. [Google Scholar] [CrossRef]
  77. Barker, R.; Burkle, D.; Charpentier, T.; Thompson, H.; Neville, A. A review of iron carbonate (FeCO3) formation in the oil and gas industry. Corros. Sci. 2018, 142, 312–341. [Google Scholar] [CrossRef]
  78. Yan, T.; Xu, L.C.; Zeng, Z.X.; Pan, W.G. Mechanism and anti-corrosion measures of carbon dioxide corrosion in CCUS: A review. iScience 2024, 27, 108594. [Google Scholar] [CrossRef]
  79. ASTM International. ASTM G1-03; Standard Practice for Preparing, Cleaning, and Evaluating Corrosion Test Specimens. ASTM International: West Conshohocken, PA, USA, 2003. [CrossRef]
  80. ASTM International. ASTM G31-21; Standard Guide for Laboratory Immersion Corrosion Testing of Metals. ASTM International: West Conshohocken, PA, USA, 2021. [CrossRef]
  81. Elgaddafi, R.; Ahmed, R.; Shah, S. Corrosion of carbon steel in CO2 saturated brine at elevated temperatures. J. Pet. Sci. Eng. 2021, 196, 107638. [Google Scholar] [CrossRef]
  82. Agrawal, A. Corrosion Monitoring. In Encyclopedia of Materials: Science and Technology; Buschow, K.J., Cahn, R.W., Flemings, M.C., Ilschner, B., Kramer, E.J., Mahajan, S., Veyssière, P., Eds.; Elsevier: Oxford, UK, 2001; pp. 1698–1701. [Google Scholar] [CrossRef]
  83. Kahyarian, A.; Achour, M.; Nesic, S. Mathematical modeling of uniform CO2 corrosion. Trends Oil Gas Corros. Res. Technol. 2017, 34, 805–849. [Google Scholar]
  84. Nesic, S. Key issues related to modelling of internal corrosion of oil and gas pipelines—A review. Corros. Sci. 2007, 49, 4308–4338. [Google Scholar] [CrossRef]
  85. Nyborg, R.; Andersson, P.; Nordsveen, M. Implementation of CO2 Corrosion Models in a Three-Phase Fluid Flow Model 2000. In Proceedings of the CORROSION 2000, Orlando, FL, USA, 26–31 March 2000; pp. 1–10. [Google Scholar] [CrossRef]
  86. Olsen, S. CO2 Corrosion Prediction by Use of the NORSOK M-506 Model—Guidelines and Limitations. In Proceedings of the CORROSION 2003, San Diego, CA, USA, 16–21 March2003; pp. 1–12. [Google Scholar] [CrossRef]
  87. Abbas, M.H. Modelling CO2 Corrosion of Pipeline Steels. Ph.D. Thesis, Newcastle University, Newcastle, UK, 2016. [Google Scholar]
  88. Nyborg, R. Overview of CO2 Corrosion Models for Wells and Pipelines 2002. In Proceedings of the CORROSION 2002, Denver, CO, USA, 7–11 April 2002. [Google Scholar]
  89. Zhao, S.; Liao, K.; Liu, Y.; Miao, C.; Wei, C.; He, G. Corrosion prediction model of submarine mixed pipeline X65 steel under a CO2/Cl- synergistic system. Arab. J. Sci. Eng. 2022, 47, 11673–11685. [Google Scholar] [CrossRef]
  90. John, R.C.; Jordan, K.G.; Young, A.; Kapusta, S.; Thompson, W. SweetCor: An information system for the analysis of corrosion of steels by water and carbon dioxide. In Proceedings of the CORROSION 1998. Association for Materials Protection and Performance, San Diego, CA, USA, 22–27 March 1998; pp. 1–19. [Google Scholar]
  91. Nesic, S.; Li, H.; Huang, J.; Sormaz, D. An open source mechanistic model for CO2/H2S corrosion of carbon steel. In NACE–International Corrosion Conference Series; AMPP: Houston, TX, USA, 2009; pp. 1–19. [Google Scholar] [CrossRef]
  92. Pots, B. Mechanistic Models for the Prediction of CO2 Corrosion Rates under Multi-phase Flow Conditions. In Proceedings of the Corrosion 1995, Orlando, FL, USA, 26–31 March 1995; pp. 1–20. [Google Scholar] [CrossRef]
  93. Zhang, R.; Gopal, M.; Jepson, W.P. Development of a Mechanistic Model for Predicting Corrosion Rate in Multiphase Oil/Water/Gas Flows. In Proceedings of the CORROSION 1997, New Orleans, LA, USA, 9–14 March 1997; pp. 1–30. [Google Scholar] [CrossRef]
  94. Song, F. A comprehensive model for predicting CO2 corrosion rate in oil and gas production and transportation systems. Electrochim. Acta 2010, 55, 689–700. [Google Scholar] [CrossRef]
  95. Xiang, Y.; Wang, Z.; Xu, M.; Li, Z.; Ni, W. A mechanistic model for pipeline steel corrosion in supercritical CO2–SO2–O2–H2O environments. J. Supercrit. Fluids 2013, 82, 1–12. [Google Scholar] [CrossRef]
  96. Choi, Y.S.; Duan, D.; Jiang, S.; Nesic, S. Mechanistic modeling of carbon steel corrosion in a methyldiethanolamine (MDEA)-based carbon dioxide capture process. Corrosion 2013, 69, 551–559. [Google Scholar] [CrossRef] [PubMed]
  97. Nesic, S.; Zheng, Y.; Jing, N.; Esmaeely, S.N.; Ma, Z. Freecorp Tm 2.0 Theoretical Background and Verification; Technical Report; Institute for Corrosion and Multiphase Technology Ohio University: Athens, OH, USA, 2018. [Google Scholar]
  98. OLI Systems Inc. OLI Studio: Corrosion Analyzer Reference Guide; Technical Report; OLI Systems Inc.: Parsippany, NJ, USA, 2023. [Google Scholar]
  99. de Waard, C.; Lotz, U.; Milliams, D.E. Predictive Model for CO2 Corrosion Engineering in Wet Natural Gas Pipelines. Corrosion 1991, 47, 976–985. [Google Scholar] [CrossRef]
  100. Marbun, B.T.H.; Fadholi, B.Z.; Sinaga, S.Z.; Purbantanu, B.A. Integrated Design and Analysis of Corrosion Rate Model on Tubular Pipe of Oil and Gas Production Well. IOP Conf. Ser. Earth Environ. Sci. 2022, 1031, 012013. [Google Scholar] [CrossRef]
  101. Thomas, D.C.; Benson, S.M. (Eds.) Carbon Dioxide Capture for Storage in Deep Geologic Formations—Results from the CO2 Capture Project; Elsevier: Amsterdam, The Netherlands, 2005; p. 1358. [Google Scholar]
  102. Erosion/Corrosion Research Center. Erosion/Corrosion Research Center (E/CRC)—SPPS Modules. Available online: https://www.ecrc-tusmp.org/ecrc (accessed on 1 October 2025).
  103. Nassef, A.; Keller, M.; Hassani, S.; Shirazi, S.; Roberts, K. A Review of Erosion-Corrosion Models for the Oil and Gas Industry Applications. In Recent Developments in Analytical Techniques for Corrosion Research; Springer International Publishing: Cham, Switzerland, 2022; pp. 205–233. [Google Scholar] [CrossRef]
  104. de Waard, C.; Lotz, U.; Dugstad, A. Influence of Liquid Flow Velocity on CO2 Corrosion: A Semi-Empirical Model. In Proceedings of the CORROSION 1995, Orlando, FL, USA, 26–31 March 1995; pp. 1–19. [Google Scholar] [CrossRef]
  105. de Waard, C.; Lotz, U. Prediction of CO2 Corrosion of Carbon Steel. In Proceedings of the CORROSION 1993, New Orleans, LA, USA, 7–12 March 1993; pp. 1–17. [Google Scholar] [CrossRef]
  106. Dugstad, A.; Lunde, L.; Videm, K. Parametric Study of CO2 Corrosion of Carbon Steel. In Proceedings of the CORROSION 1994, Baltimore, MD, USA, 27 February–4 March 1994; pp. 1–16. [Google Scholar] [CrossRef]
  107. Nesic, S.; Kahyarian, A.; Choi, Y.S. Implementation of a Comprehensive Mechanistic Prediction Model of Mild Steel Corrosion in Multiphase Oil and Gas Pipelines. Corrosion 2018, 75, 274–291. [Google Scholar] [CrossRef] [PubMed]
  108. Qu, Z.; Zou, X.; Xiong, G.; Yue, X.; Zhang, L. Hybrid Intelligent Model for Predicting Corrosion Rate of Carbon Steel in CO2 Environments. Mater. Corros. 2025, 76, 14840. [Google Scholar] [CrossRef]
  109. Nordsveen, M.; Nesic, S.; Nyborg, R.; Stangeland, A. A Mechanistic Model for Carbon Dioxide Corrosion of Mild Steel in the Presence of Protective Iron Carbonate Films—Part 1: Theory and Verification. Corrosion 2003, 59, 443–456. [Google Scholar] [CrossRef]
  110. Nesic, S.; Nordsveen, M.; Nyborg, R.; Stangeland, A. A Mechanistic Model for Carbon Dioxide Corrosion of Mild Steel in the Presence of Protective Iron Carbonate Films—Part 2: A Numerical Experiment. Corrosion 2003, 59, 489–497. [Google Scholar] [CrossRef]
  111. Nesic, S.; Lee, K. A Mechanistic Model for Carbon Dioxide Corrosion of Mild Steel in the Presence of Protective Iron Carbonate Films—Part 3: Film Growth Model. Corrosion 2003, 59, 616–628. [Google Scholar] [CrossRef]
  112. Wang, W.; Shen, K.; Tang, S.; Shen, R.; Parker, T.; Wang, Q. Synergistic effect of O2 and SO2 gas impurities on X70 steel corrosion in water-saturated supercritical CO2. Process Saf. Environ. Prot. 2019, 130, 57–66. [Google Scholar] [CrossRef]
  113. Coelho, L.B.; Zhang, D.; Van Ingelgem, Y.; Steckelmacher, D.; Nowé, A.; Terryn, H. Reviewing machine learning of corrosion prediction in a data-oriented perspective. Npj Mater. Degrad. 2022, 6, 8. [Google Scholar] [CrossRef]
Figure 1. The progression of CO2 corrosion models and software tools from 1975 to 2025 [20,28,30,31,32,33,34,36,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106]. Blue, purple, and green labels correspond to empirical, semi-empirical, and mechanistic predictions, respectively. Machine learning approaches are labeled with dark blue. The owner of the model or software is shown in parentheses, followed by the year. The subscript denotes whether it is a model (m) or a software tool (s) or both (ms). Arrows denote inspiration or continuation of work.
Figure 1. The progression of CO2 corrosion models and software tools from 1975 to 2025 [20,28,30,31,32,33,34,36,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106]. Blue, purple, and green labels correspond to empirical, semi-empirical, and mechanistic predictions, respectively. Machine learning approaches are labeled with dark blue. The owner of the model or software is shown in parentheses, followed by the year. The subscript denotes whether it is a model (m) or a software tool (s) or both (ms). Arrows denote inspiration or continuation of work.
Processes 14 00170 g001
Table 1. Comparison of selected CO2 corrosion models with their scope and limitations.
Table 1. Comparison of selected CO2 corrosion models with their scope and limitations.
ModelTypeTemp. RangePressure RangeInputsLimitationsAssumptionsSource
De Waard-Milliams (1975)Semi-empirical<90 °C-CO2 partial pressure and temperature.Does not include bicarbonate or hydrogen ions. pH only defined by CO2 equilibrium. Disregard reactions with carbonate.Relationship between temperature and CO2 partial pressure. Linking potential to corrosion current with pH-dependent expressions.[28,88,107]
LIPUCOR (Total) (1979)Semi-empirical20–150 °C<250 bar, <50 bar CO2H2S in mole%, total pressure, temperature, diameter, water chemistry, and production rate.-Based on lab and field data. Calculates pH from CO2 concentration, water chemistry, and temperature.[20,88,101]
CORMED (Elf) (1989)Semi-empirical<120 °C--Does not include oil wetting or liquid flow velocity.Accepts higher temperatures and ionic strengths but displays a warning, as the pH calculation becomes uncertain. The corrosion risk prediction is still valid. pH calculated from pressure, CO2 concentration, temperature, water chemistry, and free acetic acid.[20,88,101]
De Waard-Lotz-Milliams (1991)Semi-empirical---No flow velocity.Correction factors for the effect of corrosion product formation and pH were added.[99]
De Waard-Lotz (1993)Semi-empirical--pH, effect of corrosion product film fugacity of CO2, temperature.No flow velocity.Builds on top of the de Waard-Milliams equations. The correction factors were adjusted from the 1993 version. Uses correction factors to avoid overestimation of corrosion.[105]
De Waard-Lotz-Dugstad (1995)Semi-empirical0–140 °C<10 bar CO2 partial pressurepH, temperature, velocity, CO2 partial pressure.-The effects of fluid velocity, mass transport, and composition of the steel were added. This model often gives a lower corrosion rate than the 1991 and 1993 models at low flow rates.[88,104]
NORSOK (1995)Semi-empirical20–150 °C<1000 bar, <10 bar CO2Temperature, total pressure, CO2 concentration, wall shear stress, glycol concentration, and pH.pH between 3.5–6.5.Spreadsheet that is open access. Fitted to data from IFE.[20,88,101]
HYDROCOR (Shell) (1995)Mechanistic0–150 °C<200 bar, <20 bar CO2Multi-phase flow. Pressure, temperature, diameter, concentration, H2S, and CO2 in mole%, organic acid, glycol, and bicarbonate concentration, and production rate.-The LCR model by Pots is used for the prediction of corrosion. Determines the corrosion rate along the pipeline. Calculation of pH takes the iron and bicarbonate production into account.[20,88,101]
TULSA (1995)Mechanistic38–116 °C<17 bar CO2 partial pressureTemperature, liquid flow velocity, pipe diameter, CO2 partial pressure, and pH.Single-phase flow. Does not take effect from oil wetting.Electrochemical reaction kinetics and mass transfer. It can be used in straight and elbow pipes. pH can be input directly or calculated from water chemistry.[20,88,101]
PREDICT (InterCorr) (1996)Semi-empirical20–200 °C<100 bar CO2 partial pressureTemperature, H2S and CO2 partial pressure, and flow velocity.pH between 2.5–7. Does not give any limits, either in the software or in the manual.Based on the de Waard model, with other correction factors. Calculates CO2 partial pressure from the pH. pH can be input directly or calculated from water chemistry or from bicarbonate concentration and ionic strength.[20,88,101]
Dream (1996)Semi-empirical--Temperature and pressure at wellhead, separator, and bottomhole, well depth, diameter, water chemistry, gas composition, gas and water production rates.Does not include hydrocarbon condensates.Corrosion prediction in annular flow gas wells. Includes effects from protective corrosion films.[88]
Ohio model (1997)Mechanistic10–110<20 barTemperature, pressure, CO2 concentration, flow rate, pipe diameter, water chemistry.Protective film formation is not predicted.pH is either an input or calculated from water chemistry.[88,93,101]
Cassandra (BP) (1998)Semi-empirical<140 °C200 bar, <10 bar CO2Water chemistry, total pressure, temperature, velocity, CO2 in mole% Glycol concentration, oil type, diameter.-Implements the de Waard model and BP’s own experience. pH calculated on CO2 content, water chemistry, and temperature. Accepts input outside these values but displays a warning.[20,88,101]
KSC (IFE) (1998)Mechanistic5–150 °C<250 bar, <50 bar CO2Temperature, pressure, partial pressure of CO2, liquid flow velocity, and pH.Flow velocity between 0.2 and 30 m/s. Does not take oil wetting into account.Based on electrochemistry. Calculates corrosion rate with and without protective film formation. pH calculated on ionic strength and bicarbonate.[20,88,101]
SweetCor (Shell) (1998)Empirical5–121 °C0.2–170 bar CO2 partial pressureTemperature, pressure, pipe diameter, CO2 mole%, water chemistry, and gas/liquid superficial velocities.-Large corrosion dataset from laboratory and field data. Data is grouped using statistics to make correlations and predict the corrosion rate, or by filtering data that closely matches the conditions. pH and corrosion rate are calculated from inputs.[20,88,90,101]
Corpos (CorrOcea) (1999)Semi-empirical--Needs calculations of pH, probability of water wetting, and an external fluid flow model.-Norsok model is used to calculate corrosion rate at several points along the pipeline.[88]
OLI model (1999)Mechanistic--Temperature, pressure, molar composition, flow velocity.Does not account for oil wetting.Combines thermodynamic and electrochemical corrosion modeling. Modeling of formation and dissolution of sulfide and carbonate scales. Scale formation has been calibrated from laboratory data.[88]
Electronic Corrosion Engineer (ECE) (early 2000s)Semi-empirical Software--O2, H2S, and CO2, and NaCl concentration, water density, liquid density, and length of tubing.Quantitatively assess corrosion by using thermodynamic simulations to predict free water content for dense phase or supercritical CO2 systems.Based on de Waard 95 model. Calculates pH from bicarbonate and water chemistry.[88,100]
ULL (2002)Semi-empirical--Temperature and pressure at wellhead, separator, and bottomhole, well depth, diameter, water chemistry, gas composition, condensate density, gas, condensate, and water production rates.-Developed for gas condensate wells, predicting corrosion rate along the well. Predicts corrosion only when water condenses.[88]
Freecorp 1.0 (2008)Semi-empirical1–120 °C-Concentration of H2S, acetic acid, O2, Fe2+, temperature, pressure.Pipe diameter: 0.01 to 1 m Liquid velocity: 0.001 to 20 m/s Fe2+: 0 to 100 ppm acetic acid: 0 to 1000 ppm pH: 3 to 7, O2: 0 to 10,000 ppm.Point model: predicts corrosion at a single location with known conditions. Single-phase flow only. Uniform corrosion only. Ideal water chemistry (infinite dilution assumed). No interaction between diffusing species. Salinity effects are ignored. Iron carbonate film growth based on simple supersaturation correlation. The dominant mechanism between CO2 and H2S corrosion is chosen based on the higher predicted corrosion rate. Fe2+ is required for FeCO3 film growth kinetics. Total time must be provided if H2S is present.[91]
Freecorp 2.0 (2018)Semi-empirical--H2S, organic acids, flow rate, time.Ignore effects of high salinity, oxygen, elemental sulfur, etc.Includes formation and protection by corrosion product layer. Uniform corrosion only. Bulk water chemistry is used, which may differ from surface conditions. No corrosion product layer formation included. Steady-state model, no time-dependent effects. Ideal water chemistry (infinite dilution assumed). Single-phase flow only, uses empirical mass transfer correlations.[97]
OLI Corrosion analyzer (2019)Mechanistic--Temperature, pressure, impurity concentration, salt concentration.Aqueous systems only, not valid for concentrated acids, alcohols, etc. No film shear removal, erosion-corrosion is not modeled. Thin films are not fully included. Transport and gas/liquid equilibrium are simplified.Thermophysical and electrochemical simulation showing effects of temperature, pressure, pH, concentration, and velocity on corrosion. Inclusion of half-reactions for elements, alloys, and solution species. Electrochemical parameters such as the Tafel slope and the exchange current density are calibrated from the literature. Rigorous transport predictions include diffusion, electrical conductivity, and viscosity. Supports many alloys, including carbon steel, nickel alloys, stainless steel, and aluminum.[98]
Tubular Corrosion Desktop (TCD) (2022) (Bandung Institute of Technology)Semi-empirical Software created in C++.--O2, H2S, N2, bicarbonate, CO2, and NaCl concentration, water density, and liquid density.-Based on work by de Waard, Liane Smith, and Mike Billingham.[100]
Yang et al. (2023) (Shell)Machine learning model--CO2 and O2 partial pressure, pH, and temperature.-Uses data generated from CFD simulations. A total of 1122 data points for training and testing of the LightGBM model. The hybrid model is 106 times faster than CFD simulations alone.[34]
Dong et al. (2024)Machine learning model0–200 °C-Material, chromium content, H2O, O2, SO2, NO2, and H2S concentration, pressure, temperature, and time.-Six different models using SVM, LightGBM, XGBoost, GBDT, KNN, and RF made on 248 data points.[31]
Juian Cui (2025)Deep learning model--Temperature, pressure, pH, material, flow rate, water content, carbonic acid concentration, and CO2 concentration.-Deep confidence network model optimized with Adam algorithm, made on 75 sets of data.[33]
Table 2. Comparison of model types for CO2 corrosion prediction [33,37].
Table 2. Comparison of model types for CO2 corrosion prediction [33,37].
EmpiricalMachine LearningSemi-EmpiricalMechanistic
BasisData correlationData-drivenPhysics-informedPhysics-based
InterpretabilityLowLowMediumHigh
ExtrapolationPoorDependent on trainingPoor to reasonableGood
Computation timeLowLow after trainingLowModerate to high
Data needHighHighModerateLittle or none
Number of InputsFewModerateModerateMany
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Simonsen, K.R.; Ostadi, M.; Zychowski, M.; Pedersen, S.; Bram, M.V. Review of CO2 Corrosion Modeling for Carbon Capture, Utilization and Storage (CCUS) Infrastructure. Processes 2026, 14, 170. https://doi.org/10.3390/pr14010170

AMA Style

Simonsen KR, Ostadi M, Zychowski M, Pedersen S, Bram MV. Review of CO2 Corrosion Modeling for Carbon Capture, Utilization and Storage (CCUS) Infrastructure. Processes. 2026; 14(1):170. https://doi.org/10.3390/pr14010170

Chicago/Turabian Style

Simonsen, Kenneth René, Mohammad Ostadi, Maciej Zychowski, Simon Pedersen, and Mads Valentin Bram. 2026. "Review of CO2 Corrosion Modeling for Carbon Capture, Utilization and Storage (CCUS) Infrastructure" Processes 14, no. 1: 170. https://doi.org/10.3390/pr14010170

APA Style

Simonsen, K. R., Ostadi, M., Zychowski, M., Pedersen, S., & Bram, M. V. (2026). Review of CO2 Corrosion Modeling for Carbon Capture, Utilization and Storage (CCUS) Infrastructure. Processes, 14(1), 170. https://doi.org/10.3390/pr14010170

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop