Next Article in Journal
Numerical Study on Glaze Ice Accretion Characteristics over Time for a NACA 0012 Airfoil
Previous Article in Journal
Obtaining of Antibacterial Nanoporous Layer on Ti7.5Mo Alloy Surface Combining Alkaline Treatment and Silver: In Vitro Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Water in Epoxy Coatings: Basic Principles of Interaction with Polymer Matrix and the Influence on Coating Life Cycle

by
Nikita A. Bratasyuk
1,
Alexandr V. Latyshev
2 and
Vjacheslav V. Zuev
1,3,*
1
Chemical Engineering Center, ITMO University, Kronverkskiy pr. 49, 197101 Saint-Petersburg, Russia
2
Gazprom VNIIGAZ LLC, Malookhtinsky pr. 45 A, 195112 Saint-Petersburg, Russia
3
Institute of Macromolecular Compounds of the Russian Academy of Sciences, Bolshoi pr. 31, 199004 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Coatings 2024, 14(1), 54; https://doi.org/10.3390/coatings14010054
Submission received: 5 December 2023 / Revised: 19 December 2023 / Accepted: 21 December 2023 / Published: 29 December 2023

Abstract

:
This review describes the interaction of epoxy coatings with water, the kinetics of water uptake in epoxy resins, and the methods of studying the latter (gravimetric, IR (Infrared) and NMR (Nuclear magnetic resonance) spectroscopies). The analysis of experimental data requires using mathematical models which simulate the processes that occur during water uptake in a polymer matrix. This review classifies the numerous models applied for the study of water absorption into two large groups: Fickian and non-Fickian models. Fickian models involve the Fickian diffusion model and its development for the case of anomalous diffusion, as well as the models of Jacobs–Jones, Berens–Hopfenberg, and the time-varying diffusion coefficient model. The Carter–Kibler approach is provided as an example of a non-Fickian or Langmuir-type model. A critical analysis of the experimental methods is provided, including advantages and possible experimental errors. The methods used for processing experimental data are discussed as well as the effect of water on the exploitation properties and life cycle of epoxy coatings.

1. Introduction

Polymer coatings are widely used as the most accessible tool to protect metal and wood. Water on Earth ensures the existence of life on Earth; on the other hand, it is one of the main sources that destroy objects produced by humankind. The infusion of water stimulates corrosion processes in the materials underlying protective coatings. In polar regions, it can result in the destruction of coatings and coated materials as a result of multiple freeze–thaw cycles.
In this study, water is considered a complex liquid. The polarity of its molecules in combination with the possibility of a hydrogen bond network being formed results in a multilevel structure of liquid water. Such structures may persist during the infusion of water into the polymer matrix, which influences the diffusion processes and the infiltrate in the polymer bulk. Hence, the study of water infusion in coatings requires a multimethod approach using modeling and spectroscopic structural analysis.
Epoxy resins, with their network structure and excellent mechanical performance, high barrier properties, and thermal and chemical stability, are widely used as coatings and matrices for composite materials. Their high rigidity leads to their fragility and void formation during their preparation, which largely affects their water permeation.
The aim of most studies in this field is to examine the water permeation of common epoxy-based coatings across a range of commercially relevant temperatures (from room temperature to +80 °C). Water absorption can induce damage to the epoxy network, induce distinctive cavity formation in the coating films, and increase the aggregative porosity of the coatings. Hence, the analysis of the processes of water uptake is crucial for protecting transport pipelines, tanks, industrial constructions, and other objects. Vapor transport analysis in a broad range of temperatures from deep freezing to high sunshine is important for epoxy composites used in the aircraft industry because any damage to an aircraft is crucial. Other important engineering plastics are polyamides 6 and 6.6, which are widely used as materials for bearings, shafts, and other construction elements in the ship-building industry. The high uptake of water by these plastics and its influence on the performance characteristics of products is a topic that is not directly related to this review and would be the subject of another research. However, the main methods of studying water absorption appear to be common for all plastics and are referred to in the papers discussed.
The study of water uptake to epoxies has a long history [1]. The performance of epoxy coatings during moisture absorption has been described to decrease as a result of swelling [2], the fluctuation of bulk mechanical properties [3], the plasticization effect [4], the chemical destruction of the epoxy matrix [5], and other factors. In this study, these changes are considered to involve both experimental [6] and molecular simulation [7] methods. The complexity of this task requires new insights to advance the molecular-level understanding of the water interaction with the epoxy matrix and justifies the need for new strategies, including 2D IR spectroscopy [8] and solid-state NMR using multi-quantum filter techniques [9]. Methods based on electrochemical noise measurement have been shown to be promising as reliable quantitative indicators of corrosion [10,11,12,13].
This review includes the following sections: the experimental methods used for studying the water uptake of epoxy resins, the kinetics of this process, the water state in the epoxy matrix, the influence of water absorption on the material properties of epoxy coatings, and the molecular origin of these changes.

2. Methods for Studying Water Uptake

2.1. Gravimetric Analysis

Gravimetric analysis is one of the most popular methods used to characterize water absorption in polymers and polymer composites [14,15,16]. Rectangular samples (10 mm wide, 10 mm long, and 1 mm thick) are immersed in distilled water at room temperature or higher. Water absorption is monitored as a function of the immersion time. The samples are regularly removed from the liquid and wiped with filter paper to measure the water uptake. The change in the water content, Mt (%), is calculated using Equation (1):
M t = w t w d w d × 100
where wt is the weight of the sample immersed in water at the time t, and wd is the weight of the dry sample.
The resulting diffusion plots normally represent curves with an oblique linear initial portion and a subsequent horizontal representation of the equilibrium statement. On the other hand, the immersion time is a very important parameter that affects the type of curve that is produced. Often, researchers have achieved pseudo-equilibrium due to the measurement time not being long enough.
In many cases, the nature of water absorption at the initial stages is consistent with Fick’s law, since the absorption of water molecules in the volume occurs due to their penetration into the free volume of the polymer network. Further absorption is usually related to hydrogen bond formation, polymer network degradation, segmental relaxation, swelling, and many other negative factors. Therefore, a gradual increase in water content after reaching a pseudo-equilibrium state is typical for long-term tests [17,18,19,20,21,22,23,24,25].
The diffusion plots are easy to calculate and refer to, and thus can fit various mathematical models. These models make it possible to define the water diffusion coefficient and some other diffusion parameters. There are multiple mathematical models, and they have different solutions for calculating the kinetic parameters of water absorption [14,26,27,28].
Today, commercial epoxy coatings are very complex polymer systems containing many components including an epoxy binder, hardener, fillers, anti-corrosion pigments, solvents, reactive diluents, plasticizers, dispersants, coupling agents, thixotropic additives, and other substances of various nature. The behaviors of these compositions under water diffusion will vary depending on the morphology, structure, crosslink density, and free volume of each polymer system [29,30]. Thus, the behavior of diffusion can be described by any of the two large groups of models: Fickian and non-Fickian models.

2.1.1. Fickian Diffusion Model

The Fickian diffusion model is a straightforward approach to evaluate the water absorption of epoxy compositions [15,16,31,32,33]. Fick’s law describes the transfer of water into a polymer by diffusion only without considering intermolecular interactions and/or changes in the topology of the polymer network. In a one-dimensional system with a substance concentration gradient dC/dx in the x direction, the rate of change in the substance concentration at a given point is a function of diffusion, and it is determined by Fick’s second law (Equation (2)):
dC dt = D d 2 C dx 2
Equation (2) can be presented for a three-dimensional system:
dC dt = D x d 2 C dx 2 + D y d 2 C dy 2 + D z d 2 C dz 2
where C is the concentration of the diffused substance; D, Dx, Dy, and Dz are the diffusion coefficients; and x, y, and z are the diffusion directions.
Fick’s second law has more than one analytical solution, but the most common equation is given by Crank [16,27,34,35,36]:
M t M = 1 8 π 2 n = 0 1 2 n + 1 2 exp D 2 n + 1 2 π 2 t l 2
where M t is the water absorption at time t, M is the equilibrium water absorption, n is an integer representing the number of terms in the series, l is the sample thickness, and t is the diffusion time.
An analytical simplification of Equation (4) for a short exposure time takes the following form for M t / M < 0.5:
M t M = 4 l Dt π
and for M t / M > 0.5, it takes the following form:
M t M = 1 exp 7.3 Dt l 2 0.75
The Fickian behavior of water diffusion in cured epoxy resin is only typical for its early stages. However, many epoxy compositions exhibit behaviors that tend to deviate from Fick’s law after a long-term immersion [6,17,35,37]. The movement of water molecules into the polymer depends not only on the penetration of water into the free volume, but also on the complexity of the intermolecular interactions between the polymer network and the water. Polar groups in the structure, the crosslink density, free volume, and the results of the water–polymer interactions (including segmental relaxation, plasticization, swelling, and the destruction of the macromolecular structure) make it impossible to completely describe the diffusion process using Fick’s law [24,38,39].

2.1.2. Non-Fickian (Anomalous) Diffusion Models

Polar functional groups in cured epoxy resins and the relaxation processes resulting from the interaction with water molecules is the cause of an eventual deviation from the classical Fick’s law. There has been ample research to achieve a deeper understanding of the water absorption process in epoxy materials and to propose more efficient models. As a result, the following four models were proposed for anomalous diffusion in epoxies: two-phase polymer dual-diffusivity model (Jacobs–Jones model) [14,40], coupled diffusion–relaxation model (Berens–Hopfenberg model) [41], time-varying diffusion coefficient model [42,43], and “Langmuir-type” model (Carter–Kibler model) [44]. They include some additional factors: bound and unbound phases of water; phases with different structures of macromolecules formed within polymers; relaxation effects due to swelling; and variability of diffusion coefficients over time, to name a few. All of these models rely on experimental observations to determine their parameters and coefficients. This paper focuses on the four above-mentioned models and does not regard those used for describing nano- and reinforced composites (for example, the hindered diffusion model [28,45], barrier models [46,47], and the model proposed by Shen and Springer [48]).

Jacobs–Jones Model (Two-Phase Polymer Dual-Diffusivity Model)

The Jacobs–Jones model assumes that a polymer has a two-phase structure, including highly crosslinked and low crosslinked regions. The first stage of water absorption is suggested to correspond to the rapid absorption of water. In this case, the diffusion of molecules occurs both in the highly crosslinked phase and in the low crosslinked one. Further, the diffusion process slows down due to water absorption only into the highly crosslinked phase. Both stages follow Fick’s law, and the resulting equation is as follows [14,40]:
M t M = V d 1 exp 7.3 D d t l 2 0.75 + 1 V d 1 exp 7.3 D l t l 2 0.75
where D is the diffusion coefficient of the dense (highly crosslinked) phase, Dl is the diffusion coefficient of the less dense (low crosslinked) phase, and Vd is the volume fraction of the highly crosslinked phase. This model neglects the interaction between the water and the polymer network.
Maggana and Pissis [14] suggested a modified Jacobs–Jones model. The main difference is the final interpretation of the two phases above. Thus, there is the main (the first) phase, in which most of the water is absorbed, and the second phase with a different density and/or hydrophilic characteristics.

Berens–Hopfenberg Model (Coupled Diffusion-Relaxation Model)

The Behrens-Hopfenberg model is based on the relaxation phenomena resulting from the polymer swelling during the diffusion. The diffusion of water in a glassy polymer is facilitated by Fickian diffusion and relaxation changes in the polymer network occurring with the free volume changes [41,49]. Thus, the main equation of the model can be expressed as follows (Equation (8)):
M t = M F + M R = M F 1 exp 7.3 D d t l 2 0.75 + M R 1 exp kt 2 ,
where M F is the amount of absorbed water, related to the Fickian diffusion process, M R is the amount of absorbed water, related to relaxation, M F is the water saturation value disregarding any stress relaxation, M R is the water saturation value related to the relaxation effect, and k is a constant related to the relaxation rate of the polymer.
Research on absorption in polymers has revealed that the rate-controlling process depends on the size of the penetrant [41]. When the absorption starts rapidly but slows down at the later stages, the diffusion outpaces the relaxation. However, if these processes operate simultaneously, it becomes challenging to differentiate them based on the absorption curve.

Time-Varying Diffusion Coefficient Model

A one-dimensional time-varying diffusion coefficient model is a modified Fickian model (Equation (2)) interpreting the diffusion coefficient as a time-dependent function. Weistman [42] suggested the following equation for a diffusion model with a time-varying diffusivity:
C t = D 0 exp B θ t 2 C x 2
where D0 and B are material constants and θ t is the absolute temperature.
According to Roy et al. [43], moisture uptake can be expressed by Equation (10):
M t M = 1 8 π 2 n = 0 1 2 n + 1 2 exp 2 n + 1 2 π 2 l 2 × D 0 t + r = 1 R D r t + τ r e t τ r 1
The main parameters, M , D0, Dr, and τ r , are determined by fitting Equation (10) to the experimental data.
This model is often used for water diffusion in epoxy composites and adhesives [34,42,43,47,50,51,52,53]. Despite obtaining good approximation results for composites and adhesives, the model with a time-varying diffusion coefficient has been rarely found to describe the water absorption process in epoxy coatings.

Carter–Kibler Model (“Langmuir-Type” Model)

The Langmuir model and its modifications consider chemical interactions between water molecules and polar groups of the epoxy polymer in contrast to the above-mentioned models. The Langmuir-type model assumes that there are two states of absorbed water molecules: bound and free. According to the model, the state of water molecules at any time and in any place in the polymer unit may regularly change from bound to free (this can happen with probability β) and from free to bound (with probability γ). The coefficients β and γ are also referred to as the desorption and absorption coefficients, respectively. The Langmuir equation for the one-dimensional case has the following form [33,54,55,56,57]:
dC f dt + dC b dt = D d 2 C dx 2
dC b dt = γ C f β C b
γ C f = β C b
where C f and C f are the concentration of «free» water and its maximum concentration, respectively, and C b and C b are the concentration of «bound» water and its maximum concentration, respectively.
Carter and Kibler [44] derived Equation (14) for calculating water absorption:
M t M = 1 8 π 2 i = 1 r i + exp r i t r i exp r i + t i 2 r i + r i + 8 π 2 κ β β + γ i = 1 exp r i t exp r i + t r i + r i
r i ± = 1 2 κ i 2 + γ + β ± κ i 2 + γ + β 2 4 κ β i 2
κ = D γ π l 2
These equations can be simplified for ease of calculation:
M t M = β γ + β e γ t y t + 1 e β t + β γ + β e β t e γ t
y t = 1 8 π 2 n = 0 1 2 n + 1 2 exp 2 n + 1 2 κ t
A typical water absorption curve can be presented as the sum of free (unbound) and bound water (Figure 1). There are normally two stages of absorption—the first one (characteristic of unbound absorption) and the second one (characteristic of bound absorption).
The absorption coefficient γ and the desorption coefficient β are very important parameters of the Carter–Kibler model. Scott and Lees [55] reported some difficulties in selecting the method for calculating these parameters and compared a number of methods used to estimate the coefficients β and γ. The authors chose a simplex algorithm to evaluate the non-linear regression corresponding to the experimental data.
The Langmuir-type model showed the greatest efficiency among those described for non-Fickian diffusion in epoxy coatings [9,26,34,37,39,51,54,55,56,57,58,59]. For example, Glaskova et al. [51] quantified the performance of a number of the above models for a selected commercial epoxy resin and found that Langmuir-type and relaxation models provide results that have adequate agreement with the experimental data. Other researchers [55,58,59] also showed that Langmuir-type models provide highly convergent results and can properly predict a further increase in mass from a set of obtained experimental data.
On the other hand, a good approximation of the data does not always confirm the validity of the model. Two models with different permissible physical justifications may equally approximate the same water absorption data. Therefore, the true applicability of the model has to be confirmed by other control tests (e.g., absorption–desorption test and elevated temperatures) or by studying water absorption using spectral analysis methods.
Using gravimetric measurements without any comparison with other methods (e.g., IR or NMR) may also lead to an interpretation error. Unfortunately, gravimetric analysis does not allow for the study of the diffusion mechanism, the relaxation behavior, and the interactions between the penetrant and the polymer. Nevertheless, the absorption kinetics obtained by gravimetric analysis is essential for studying the water absorption of epoxy compositions.

2.2. Infrared Spectroscopy (IR)

Another widely used method is infrared (IR) spectroscopy in the middle and near-infrared ranges [19,31,60,61,62,63,64,65,66,67,68,69]. FTIR spectroscopy has been successfully used to investigate water diffusion in a thermoset and, in particular, in epoxy materials. In contrast to gravimetric analysis, it is not only possible to observe the concentration dependencies of water molecules in the polymer volume, but also to determine the interaction of the penetrant and the polymer network at a molecular level.
The equipment available today contributes to a more accurate investigation of the diffusion processes in polymer systems compared to the facilities that existed 30–40 years ago. For example, the development of in situ measurement technology has improved the accuracy of analyzing the changes in polymer systems during moisture absorption. Measurements are performed in real time with continuous spectrum removal using special tooling or techniques (in situ time-resolved spectroscopy, difference spectroscopy, or 2D correlation spectroscopy). Many studies are performed using FTIR-ATR (attenuated total reflectance) spectroscopy. This technique is often criticized due to a number of drawbacks: insufficient contact between the specimen and the crystal; a more complex relationship between absorbance and water concentration compared to transmission; and the “skin effect” and multilayer structures in thin films that make interpreting the resulting data difficult [20]. However, the influence of these factors is negligible. The most common research approach is to collect spectra in the middle range (4000–400 cm−1) of dry and wet samples at different time intervals [8,70,71,72]. The further quantitative analysis is performed by the deconvolution of the spectra in the ranges of 3700–3100 cm−1 (stretching vibrations of hydroxyl groups) and of 1700–1600 cm−1 (deformation vibrations of hydroxyl groups). These ranges contain information about both intermolecular and intramolecular hydrogen bonding. The region from 3700 to 3100 cm−1 is more intense and contains more information about the hydrogen bonds, while that of 1700–1600 cm−1 is less informative due to the overlapping with other functional groups and low signal intensity [9,60,66].
According to the research [9,15,60,66,73,74], the intensity of the bands corresponding to OH-group vibrations increases, and the peak broadens with the increasing diffusion time. The selected region has a single overlapping broad peak that cannot be used in diffusion studies without some additional processing. Therefore, according to the results of deconvolution, it is customary to distinguish the following ranges: 3800–3500 cm−1 (S0, free/mobile water molecules), 3600–3400 cm−1 (S1, water molecules bound by weak hydrogen bonds), 3500–3200 cm−1 (S2L, moderately bound water molecules), and 3300–3100 cm−1 (S2T, strongly bound water molecules) [9,15,60,62,63,66,71,75]. Figure 2 illustrates the possible deconvolution peaks.
The number of water molecule states differs from one study to another; however, most researchers agree on the definitions of free, weakly related, and strongly related states.
In addition to the middle range, the near infrared (NIR) range is of particular interest for studying water diffusion by evaluating the stretching and bending vibrations during hydrogen bond formation. The NIR region of the infrared spectrum has a wavelength range of 8000–4000 cm−1. The current study focuses on two ranges, 7300–6300 cm−1 and 5400–4800 cm−1, that correspond to the -OH stretching vibration band and the –OH bending vibration band, respectively. Although the 5500–4800 cm−1 range provides qualitative information at the molecular level, quantitative analysis is very difficult, as the hydrogen bond has opposite effects on stretching and bending vibrations, which reduces the ability to separate different components of a complex profile [73,76]. Therefore, the interval of the hydroxyl group stretching vibration is often analyzed in more detail [62,76,77,78,79,80,81].
The deconvolution of the 7300–6300 cm−1 range shows three overlapping peaks (Figure 3): a peak at 7080–7050 cm−1, attributed to the free state of water molecules (type S0); a peak at 6850–6810 cm−1, assigned to singly hydrogen-bound water molecules (type S1); and a peak at 6580–6535 cm−1, attributed to strong hydrogen-bound interactions (type S2).
As water is absorbed, the width and the height of the peaks change due to many factors. These changes allow for the determination of the exact nature of hydrogen bonding as well as the resulting changes in properties. Polar groups, the structure and branching of the polymer network, its defectiveness, and the compatibility of the system components are among the factors that will determine the diffusion rate and the ways water molecules will penetrate into the polymer system volume.
An isolated spectrum of absorbed water is required for adequate data analysis. Using difference spectroscopy, the spectra of absorbed water are isolated by eliminating the interference of the epoxy network spectra [15,75]:
A dif = A wet KA dry
where Adif, Awet, and Adry are the absorbances of the difference spectrum, wet spectrum, and dry spectrum, respectively, and K is the subtraction factor, i.e., an adjustable parameter used to compensate for possible thickness changes between the sample and the reference spectra [15,75,79]. Usually, K is equal to 1 because the thickness changes upon water absorption are negligible for most specimens.
The resulting time-resolved spectra are studied using generalized 2D correlation spectroscopy. This method better distinguishes the overlapping peaks and determines the correlation between them. Two-dimensional correlation FTIR spectroscopy has a number of advantages: it can detect overlapping spectroscopic information, it can easily search for minor changes in peaks, and it can identify specific sequence of events occurring in the system under the influence of external factors [9,15,60,62,65].
Noda [82] proposed the basic principles of 2D IR spectroscopy based on the 2D NMR correlation method. The generalized 2D method considers signals as arbitrary functions of time, temperature, pressure, concentration, or any other physical variable, and defines the intensity of synchronous and asynchronous correlations (Equation (20)):
Φ υ 1 , υ 2 + i Ψ υ 1 , υ 2 = 1 π T max T min 0 Y 1 ω · Y 2 * ω d ω
where Φ υ 1 , υ 2 is the synchronous two-dimensional correlation intensity; Ψ υ 1 , υ 2 is the asynchronous two-dimensional correlation intensity; Y 1 ω = y υ 1 , t e i ω t dt is the Fourier transformation of changes in spectral intensity, observed for a certain spectral variable y υ 1 , t relative to the external variable t; Y 2 * ω = y υ 2 , t e + i ω t dt is the variation of spectral intensity y υ 2 , t at another spectral variable υ 2 ; ω is the Fourier frequency; and t is the fixed interval of some external variable between T min and T max .
The synchronous spectrum characterizes the degree of coherence between the dynamic fluctuations of the IR signals measured at two different wavenumbers and shows a general similarity between the two separate spectral intensity changes. In contrast, the asynchronous spectrum describes the independent and inconsistent fluctuations and indicates that the bands υ 1 and υ 2 vary in phases relative to each other during the diffusion. The typical synchronous and asynchronous spectra are shown in Figure 4.
According to Noda’s rule, if Φ υ 1 , υ 2 > 0 and Ψ υ 1 , υ 2 > 0, the band υ 1 changes before the band υ 2 . And if Φ υ 1 , υ 2 > 0 and Ψ υ 1 , υ 2 < 0, the band υ 1 changes after the band υ 2 . The red regions in the 2D correlation spectra generally indicate a positive correlation, and the blue regions indicate a negative one. The coordinates of the cross-peaks, e, located below the correlation diagonal, are to be indicated as υ 1 , υ 2 , where υ 1 and υ 2 refer to the spectrum frequencies by the X-axis and Y-axis. The coordinate values also have a “+” or “−“ sign depending on the positive or negative value of the peak, respectively.
These approaches can be used for both medium and near-infrared ranges. More details on the test methodology and interpretation can be found in [15,31,33,60,62,64,65,66,67].
Besides the features described above, spectroscopy also provides the diffusion coefficient, absorption kinetic plots and profiles, concentration plots, volume change during water uptake, and other parameters [19,20,31,64,70,74,75,83].

2.3. Nuclear Magnetic Resonance (NMR)

Nuclear Magnetic Resonance (NMR) spectroscopy is a powerful technique for investigating water diffusion in epoxy coatings [4,84,85]. NMR exploits the behavior of atomic nuclei in a magnetic field and provides valuable information about molecular mobility, diffusion behavior, and the interaction between water molecules and the polymer matrix. Using various NMR methods, researchers can quantitatively analyze water diffusion, probe water–polymer interactions, and evaluate coating performance. By applying magnetic field gradients and analyzing the resulting changes in the NMR signal, researchers can determine the diffusion coefficient and the spatial and temporal distribution of water molecules [86].
Using NMR techniques can provide an understanding of the effects of temperature, humidity, and of the coating structure on diffusion. Moreover, NMR spectroscopy can be combined with other NMR-based methods, such as relaxometry and imaging, to gain some further information about water diffusion processes and their impacts on coating properties [34,58,87,88,89].
When studying the water–epoxide interaction using NMR spectroscopy, it is deuterium oxide, D2O, that is used as the diffusing liquid, rather than hydrogen oxide, H2O. This is explained by the fact that the 1H isotope makes an essential proportion of the epoxy composition, making the 1H signal from the polymer dominant in the NMR signal, while the content of deuterium, 2H, is very small. Therefore, using D2O and 2H-NMR will provide the data that refers only to diffusing heavy water [34,59,90,91].
The sensitivity of 2H-NMR experiments allows for the assessment of molecular mobility and, consequently, the physical state of water molecules. The 2H NMR spectra (Figure 5) can be interpreted as a superposition of only two subspectra corresponding to water molecules reorienting either very rapidly or very slowly. The decomposition of the spectra may be related to the mobile/“fast moving” phase (narrow peak) and bound/“slow-moving” phase (wide peak) of heavy water [9,59,90]. The relative proportions of highly mobile and less mobile water molecules can be determined by analyzing the intensity of narrow and broad components of the 2H NMR spectrum.
The interpretation and comparison of absorption data obtained using different methods of analysis have revealed an interesting fact. Some authors [59,90] obtained time dependencies of water absorption in epoxy compositions for each type of water by integrating the two peaks of the NMR signal. Despite good convergence between the results obtained using NMR spectroscopy and using the Carter and Kibler model (gravimetric analysis), there was a striking difference in the interpretation of the terms “mobile” and “bound” water for each of them. Popineau et al. [59] noted that the “mobile” water in the Langmuir-type model is actually the “bound” phase in NMR spectroscopy, and vice versa. From a kinetic perspective, the following explanation can be given: the “free” phase refers to individual water molecules with high mobility that can easily permeate into the polymer’s bulk. On the other hand, the “bound” phase represents the water molecules that have formed hydrogen bonds and occupy the microcavities and channels within the polymer. These water molecules are constrained by the available free volume within the polymer matrix. The authors emphasize that the “mobile” and “bound” states are related to kinetic concepts, and not directly to intermolecular water–polymer interactions.

NMR Imaging

NMR imaging, including Magnetic Resonance Imaging (MRI), provides spatially resolved information on water diffusion patterns within epoxy coatings [34,87,92]. These imaging techniques provide a visualization and mapping of the water distribution and the penetration depth. By acquiring multiple images over time, researchers can also monitor the progression of water diffusion. NMR imaging provides valuable insights into the spatial heterogeneity of water diffusion and can guide the optimization of coating designs for improved water resistance.
The principles of NMR imaging are similar to NMR spectroscopy, with spatial encoding added. In NMR imaging, a strong, homogeneous magnetic field is applied to the sample, which aligns the nuclear spins of the water molecules. Radiofrequency pulses are then applied to perturb the spins, followed by the detection of the resulting NMR signal. Spatially varying magnetic field gradients applied during the experiment provide the NMR signal encoding.
The reconstructed NMR images provide visual information about the transport and distribution of water within the epoxy matrix. A quantitative analysis can be performed on the images to extract parameters including water diffusion coefficients, water uptake profiles, water concentration gradients, and other parameters [34,87,92].
NMR imaging allows for the non-destructive observation of dynamic processes, including water diffusion and flow, in real time. By varying the imaging parameters, it is possible to study various aspects of water transport in a polymer, including the diffusion rates, spatial distribution, and the effects of temperature, humidity, and composition on water transport behavior.
To summarize, NMR spectroscopy offers essential opportunities for studying water diffusion in epoxy coatings. Through quantitative analysis, the investigation of water–polymer interactions, and the assessment of coating performance, NMR techniques provide valuable information on diffusion kinetics, molecular dynamics, and coating behavior. These data contribute to the development of more water-resistant epoxy coatings for various applications.

2.4. Electrochemical Impedance Spectroscopy (EIS)

The basic principle of EIS involves applying an AC electrical signal to the polymer–water system and measuring the resulting impedance response across a range of frequencies. The impedance (Z) is a complex quantity that consists of two components: resistance (R) and reactance (X). The resistance component represents the ohmic losses in the system, while the reactance component reflects the capacitive and inductive effects.
EIS can be utilized to assess the transport properties of water within the polymer matrix. As water diffuses into the polymer, it influences the electrical conductivity and impedance of the system. The analysis of the impedance spectra obtained through EIS provides valuable information about the water diffusion process [93] that increases the material’s polarizability. This enhanced polarizability raises the polymer permittivity. Measuring impedance quantifies the rise in permittivity through the derivation of capacitance. The relationship between the permittivity of the polymer film and the capacitance is expressed by Equation (21) [19,94].
C = ε ε 0 A d
where A is the area of the sample, d is the sample thickness, ε is the dielectric constant of the medium, and ε 0 is the permittivity of free space.
Assuming that the changes in capacitance are directly proportional to the changes in the dielectric constant (while keeping the geometrical parameters constant), using Equation (6) with any given time “t” and at time zero estimates the volume water uptake (Brasher–Kingsbury equation):
χ V = log C t / C 0 log ε w
where Ct is the capacitance at time t, C0 is the dry coating capacitance, which is generally obtained by extrapolating t to zero, and ε w is the dielectric constant of water. The volume water uptake χ V obtained from the EIS experiments on coatings can be compared to the mass water uptake χ m determined through gravimetric measurements. This comparison was made possible by following the procedure described in [25,95]. The resulting absorption curves can be fitted to calculate the diffusion coefficient using various Fickian and non-Fickian models that have been considered above for gravimetric analysis [19,95,96,97].
The impedance spectra obtained from EIS measurements provide valuable information about water transport features, including the diffusion of water molecules through the epoxy matrix, the water-filled voids or channels present, and the interaction of water with the epoxy surface or interfaces. The spectra can be analyzed using equivalent circuit models. These models consist of various circuit elements that represent the electrochemical processes occurring during water diffusion, including charge transfer at interfaces, Warburg impedance for diffusion, and solution resistance. Fitting the experimental data to the equivalent circuit model provides the information about the underlying transport mechanisms and diffusion coefficients [69,98,99].
In addition, EIS provides an evaluation of the electrochemical barrier performance of epoxy coatings against water and corrosion. Measuring the impedance response of the coated system assesses the resistance to ion transport and the effectiveness of the epoxy coating in preventing the penetration of corrosive species and other corrosion processes. The impedance modulus at the low-frequency domain (0.1 or 1 Hz) can generally be regarded as an indicator of systematic barrier properties of a coated metal [62,99,100]. The double-layer capacitance Cdl and the charge transfer resistance Rct were the electrochemical parameters related to the corrosion reaction to the metal/coatings interface, and they were obtained from fitting the EIS spectra [68,69,97,98].
EIS offers a versatile and non-destructive method for studying water transport in epoxy materials and assessing their anticorrosive performance. By providing information about water diffusion, coating integrity, and electrochemical barrier properties, EIS provides an understanding of the behavior of epoxy coatings in corrosive environments and supports the development of improved corrosion protection strategies.

2.5. Other Methods

2.5.1. Fluorescence

Fluorescence-based techniques offer a powerful approach to study water diffusion in polymer materials [101,102,103]. Fluorescent probes or dyes can help to track and visualize the movement of water molecules within the polymer matrix.
One common method is to incorporate a fluorescent dye or probe into the epoxy material. The dyes or probes are sensitive to water, and their fluorescence properties change upon the interaction with water molecules. Trans-4-nitro-4′-aminostilbene (NAS) [104], tricyanovinyljulolidene (TCJ) [103], 5-Dimethylamino-1-naphthalene-sulfonylchloride (DNS-Cl) [105,106], 5-Dimethylaminonaphthalene-1-[N-(2-aminoethylenesulfonamide)] (DNS-2), and 5-Dimethylaminonaphthalene-1-[N-(2-aminopentylenesulfonamide)] (DNS-5) [107] were used as external fluorophores to control the absorption of water molecules into epoxy resins. As water diffuses into the epoxy, the fluorescence signal can be monitored and analyzed to understand the dynamics of water diffusion.
Fluorescence microscopy or spectroscopy can be employed to observe and quantify the diffusion of water in a polymer. Exciting the fluorescent dye with a specific wavelength of light allows for the measurement of the emission intensity or lifetime of the fluorescence signal. As water molecules diffuse, the fluorescence signal can change, providing information about the spatial and temporal distribution of water within the material.
Furthermore, fluorescence correlation spectroscopy (FCS) can be used to measure the diffusion coefficients and dynamics of water in coatings at the microscopic level. FCS analyzes the fluctuation in the fluorescence signal caused by the diffusion of fluorescently labeled water molecules. By analyzing these fluctuations, it is possible to determine the diffusion coefficient and investigate the mechanisms of water diffusion in the polymer matrix [108].

2.5.2. Atomic Force Microscope-Infrared Spectroscopy (AFM-IR)

When investigating water diffusion in polymers using AFM-IR, water molecules’ distribution and movement within the polymer matrix can be visualized and characterized. AFM provides topographical information, identifying surface features and defects. Meanwhile, IR spectroscopy provides chemical identification and mapping based on the absorption of infrared light by various molecular species [109].
The AFM-IR setup allows us to directly observe and analyze the diffusion of water molecules in polymers. Water absorption and diffusion in a polymer can cause changes in its chemical composition and structure. These changes can be detected and monitored through variations in the infrared absorption spectra obtained by AFM-IR. By mapping these changes, it is possible to visualize the movement and diffusion of water at a high spatial resolution.
Thus, Morsch and colleagues [8,110,111] were able to directly visualize the water absorption in epoxy-phenolic compositions on a submicron scale using the AFM-IR method, as well as characterize the distribution of absorbed water and study the morphology of the coating.

2.5.3. Positron Annihilation Lifetime Spectroscopy (PALS)

PALS involves the use of positrons, which are positively charged antiparticles of electrons. When positrons are implanted into a material, they rapidly annihilate with electrons, resulting in the emission of gamma rays. Measuring the time delay between the positron injection and annihilation provides information about the free volume and defects within the material. PALS not only determines the amount of free volume, but also the distribution of free-volume holes [112,113]. As mentioned earlier [18,30,85], polarity, the topology of the polymer structure, and molecular mobility determine the absorption kinetics. Polarity has the greatest influence on the equilibrium moisture absorption; the greater the polarity of the polymer units, the more water is absorbed [22].
It is often a very challenging task to estimate the topological and polar contributions to the absorption process. However, two separate research groups [22,114] managed to isolate and study the influence of topological contribution on the nature of absorption. One of the groups used various amine isomers to obtain polymer networks that were fundamentally different in terms of their structure and packing density while maintaining identical chemical compositions and network polarity [114]. The second group selected the “epoxy resin-hardener-chain extender” system and synthesized the epoxy compositions, in which the polarity remained constant, while the topology systematically changed [18]. In both cases, the kinetics of the penetrant absorption was found to strongly depend on the size of the free-volume holes.

3. Influence of Water on Glass Transition Temperatures of Epoxy Coatings

The glass transition temperature (Tg) is a crucial parameter for epoxy coatings as it strongly affects the material’s mechanical, thermal, electrical, and chemical properties. Understanding and controlling the Tg is essential for achieving desirable coating performance and durability. The influence of water on the glass transition temperature of epoxy coatings is one of the most important considerations for designing and formulating coatings for various applications. Understanding the plasticization effect of water can help engineers and coating manufacturers optimize formulations to ensure long-lasting performance, even in water-exposed environments. Effective mitigation strategies produce highly durable and protective epoxy coatings for a wide range of applications.
When water diffuses into an epoxy coating, this can lead to changes in the Tg of the epoxy. Water molecules can affect the mobility of polymer chains. Water molecules can act as plasticizers (so that polymer chains can move more easily) and reduce the Tg of the epoxy coating. As water diffuses into the coating, it disrupts the polymer chains, increasing their mobility. This increased mobility lowers the Tg and can result in a softer and more flexible material. Usually, the decrease in Tg due to the plasticizing effect ranges from 10 to 20 °C for every percent of water molecules in the epoxy material [6,53,115,116,117,118].
Water diffusion highly depends on the polymer structure, crosslink density, environmental conditions (including temperature, pressure, humidity, and presence of salts), steric effect, polymer–penetrant interaction, and thermal and mechanical history. Since the relaxation phenomena also have significant effects on the changes in the nature of absorption, Fickian diffusion predominates in an amorphous polymer at temperatures much higher than Tg, while non-Fickian behavior regularly occurs in polymers at the glass transition temperature and below. This is explained by the relaxation of the polymer, which is essential for penetrant molecules to diffuse into the molecular network. This rearrangement results in a significant macroscopic swelling related to the diffusion of vapors and liquids into polymers [119,120,121].
When water penetrates into the epoxy matrix, there is a change in the effective crosslink density as a result of the interaction of the polymer and water molecules. This leads to an increase in the distribution of molecular weights and in the relaxation time. The effect of this process can be observed on thermograms when measuring the glass transition temperature using the DMA method. As the water content increases, the tan delta (tan δ) peak in the glass transition region broadens towards lower temperatures [61,122,123]. Often, the tan δ peak can become separated into two peaks with low- and high-temperature maxima [5,6,122,124]. The high-temperature maximum corresponds to a less plasticized network, while the low-temperature one corresponds to a highly plasticized network. The tan δ peak will change until the water saturation is reached. After that point, no evolution in the shape of the peak is observed. This is often explained by the plasticizing effect. Nogueria et al. [5] explained the observed phenomenon by the plasticizing effect and two additional factors: the free volume and the epoxy–water interactions. The formation of hydrogen bonds between water molecules and polymer chains requires breaking the interchain hydrogen bonds, which leads to an increased mobility of the chain in the glass transition region and a corresponding decrease in the effective crosslink density.
The extent of the Tg change depends on the concentration of water diffusing into the epoxy coating. Higher water concentrations typically result in more important reductions in Tg. The relationship between the water concentration and Tg change may not be linear, and different epoxy systems may exhibit varying sensitivity to water plasticization. Intramolecular interactions, post-cure, swelling, microcracking, physical aging, hydrolysis, and chain scission at higher temperatures also result in a different Tg [39,49,125,126,127].
Understanding the effect of water uptake on Tg and subsequent chemical resistance is essential for selecting appropriate coatings for environments where exposure to chemicals is a concern. Prediction models for the glass transition temperature of epoxy coatings during water diffusion are based on various approaches, ranging from empirical equations to more complex theoretical models. The goal of these models is to estimate how Tg changes with the water content in the increasing epoxy matrix. Some of the most commonly used methods are the Fox, Kelley-Beuche (free volume model) and Simha–Boyer equations.
The Fox equation is an empirical model that relates the glass transition temperature of a polymer to the weight fractions of various components. The equation can be modified for epoxy–water systems to consider the effect of water (or other low-molecular-weight diluent) uptake [114,125,128]:
1 T g = w p T gp + w w T gw
where Tgp and Tgw are the glass transition temperature of the polymer and water, respectively, and wp and ww are mass fractions of polymer and water, respectively. According to [129,130], the experimentally obtained value of Tg of the water is in the range between 124 and 138 K.
The Fox equation is formulated based on the assumption of ideal volumetric additivity at Tg, and it neglects any specific interactions between the two components. Therefore, this model, in most cases, poorly describes the change in the glass transition temperature during water absorption as there are specific interactions between the water and the polar groups within the epoxy network.
The free volume theory (Kelley-Bueche equation) suggests that the glass transition is related to the available free volume in the polymer matrix. By considering the effect of water uptake on the free volume, the Tg changes can be estimated [131,132,133]:
T g = α gp V p T gp + α lw 1 V p T gw α gp V p + α lw 1 V p
where αgp is the difference in the coefficient of thermal expansion (CTE) of a polymer between the glassy and rubbery states; αlw is the CTE of water; and Vp is the volume fraction of the polymer. The value of Vp can be calculated using the following equation [134]:
V p = 1 1 + 0.01 M m ρ p ρ w
where Mm is the equilibrium water content, and ρp and ρw are the densities of the polymer and water, respectively.
This approach requires knowledge of the epoxy’s free volume properties and their variation with water uptake. All parameters required to calculate the glass transition temperature can be obtained from the literature (for example, the CTE and Tg of water are taken as 4 × 10−3/K and 277 K, respectively [133,134,135]) or experimentally (for example, the thermal expansion coefficients αlw and αgp are determined with a TMA experiment). This model has been used by many authors [124,132,133,134,135,136] and has proven to have good agreement with experimental data.
The Simha–Boyer equation is another empirical model that describes Tg changes with the water uptake. It is given in [137] as follows:
1 T g = 1 T gp + 1 T gw 1 T gp V w
where V is the volume fraction of water in the polymer. As mentioned above, the Tg of water is in the range between 124 and 138 K. It was also mentioned [117,118] that this method is able to adequately describe the decrease in the glass transition temperature as a result of the water penetrating into the polymer.
In addition to empirical and theoretical models, there are molecular simulation methods and techniques that can be used to predict the glass transition temperature during water uptake in the epoxy. Molecular dynamics (MD) simulations can provide insights into the behavior of polymers at the molecular level. By simulating the interactions between epoxy and water molecules, MD simulations can offer predictions of Tg changes during water uptake. However, MD simulations require detailed knowledge of the epoxy molecular structure and parameters, as well as ample computational resources [135,138].
It is important to note that each method has its own limitations and may not accurately capture the specific behavior of the epoxy system selected. Therefore, accurate predictions may require a combination of approaches along with an experimental validation. Additionally, a specific formulation, curing agents, and environmental conditions of the epoxy system can influence the Tg changes, and accounting for these factors would improve the accuracy of the predictions.

4. The Effect of Water Absorption on Mechanical Properties of Epoxy Materials

Epoxy materials are the basis for preparing composites used in the production of aircrafts, wind, and tidal turbine blades [139]. They are in direct contact with atmospheric water (moisture), rain, or sea water. As a result, they undergo humidity damages during their operation. Composite materials were shown to absorb more water from the environment compared to common polymers [140]. This has a detrimental effect on the mechanical, electrical, and thermal properties of epoxy composites and reduces their service lives [141].
A high moisture uptake can lead to obvious mechanical degradation and even severe failure and damage [142]. The inclusion of fillers was found to decrease the moisture absorption, increase the glass transition temperature, and slightly reduce the detrimental effect on the mechanical properties after hygrothermal conditioning [143]. However, no long-term durability of epoxy resins seems to have been comprehensively documented.
Water attacks epoxy-based composites at every level. The fibers, matrix, fiber–matrix interface, and adhesives are all susceptible to deterioration. Absorbed water can increase creep and relaxation, introduce residual stresses, cause osmotic pressure, and degrade the epoxy matrix, fillers, and filler/matrix interfaces via hydrolysis and chemical nucleophilic attack and acid or alkali catalysis. Water accelerates the fatigue degradation of composites and shortens their fatigue life. Additionally, fatigue damage offers new paths for moisture ingress and significantly increases the rate of moisture-related damage. However, the effect of fluids tends to be contradictory. For example, fluids result in improved impact resistance. This may be due to the impacting object being resisted by a larger volume of fibers. However, over time, impact damage may allow more routes for moisture to penetrate into the composite, and, thus, impair its properties over time [144].
Moisture damage begins near the surface of the material and spreads inward over time, with cracks tending to grow parallel to the free surface. This damage is often localized, resulting in a small number of large cracks [145]. Crack growth is dominated by different effects depending on the level of loading. At lower load levels, cracking is most influenced by chemical reactions. At moderate load levels, cracking is most affected by diffusion. At higher load levels, stress-assisted corrosion controls crack growth [146]. The rate of mechanical performance degradation of epoxy-based composites has been observed to be directly correlated with the rate of moisture absorption [147]. Moisture is attracted to areas of air entrainment such as voids and delaminations. Thus, these areas can collect water over time [148]. Cracks and voids, even the microscopic ones, allow for easier penetration of water into the composite system via capillary action and diffusion. Therefore, a generous application of epoxy resin can potentially render moisture-related effects negligible. After an initial period of seeking out and filling cracks and voids, moisture begins to swell the composite. One study observed a linear relationship between strain and water uptake from the beginning of swelling [149]. Notably, water has been shown to diffuse more slowly through epoxy composites than through polyester composites, to give an example [150].
Water uptake results in microcrack formation in the bulk of the epoxy and in the amplified voids, which tend to become linked together, generating paths, which promotes further penetration of water in the epoxy depth. The absorbed water may prove damaging to installed wraps. The tendency of composite materials to suffer damage when operating in harsh humidity environments shows that the humidity issue for epoxy-based composites should not be underestimated. It needs, instead, a rigorous and deep study to evaluate the impact on the system performance and assess the reduction in the operative life.
The moisture history of the composite is important. Both the maximum load at failure and the fracture toughness have been observed to decrease linearly as the time of moisture exposure increases. A higher average moisture content over the life of the material will result in a higher level of damage, and desorption appears to be more destructive than absorption. Additionally, wet–dry cycling has been observed to decrease the ultimate load of various epoxy-based composites. Typically, glass-fiber-reinforced composites suffer much more than carbon-fiber-reinforced composites.
Epoxy matrices are the primary victims of moisture absorption. If unchecked, moisture absorption by the resin will result in plasticization, swelling, hydrolysis, and fiber debonding from the matrix [45]. By plasticizing the polymer and lowering the glass transition temperature, moisture can soften a polymer and increase creep effects. This is typically reversible with drying. The hydrolysis of the epoxy resin, however, is irreversible, and can change both the stiffness and the strength of the polymer. Degradation is substantially slowed if the resin is fully cured prior to exposure [7]. Additionally, water-resistant epoxies have been shown to increase the reliability of glass-fiber-reinforced epoxy composites by 200%–300%. Moisture absorption is known to encourage fibers to debond from the surrounding matrix. Matrix cracking allows further paths for moisture penetration, thus accelerating damage growth. Additionally, fluid ingress can result in swelling, which can cause stresses between the fibers and the matrix and deteriorate the bond between them, resulting in tiny debondings. Epoxy resin swelling and osmotic pressure due to moisture uptake can cause debonding stresses between the fibers and resin.

5. Conclusions

This review discusses water uptake in epoxy thermosets and its effect on their physical properties. The absorption of water strongly affects the exploitation characteristics of epoxy coatings and their lifecycles. The models used for describing water absorption to a polymer matrix involve many parameters that can adequately describe the process with parameter tuning. However, when using multiple variable parameters, the physical essence of absorption can be obscured. Hence, one of the promising directions in this field is the development of new models with fixed physically transparent parameters, which will adequately describe the water uptake process in epoxy resins.
Another challenge is the development of experimental techniques that would determine the water state in an epoxy matrix in a straightforward and unambiguous manner. Conventional and amply used experimental methods (IR and NMR spectroscopies) do not appear to address this issue directly because the interpretation of the data obtained is based on spectra simulation. Hence, the models underlying the calculation predetermine the results.
Therefore, new investigations are required to understand the mechanisms of water uptake for advanced epoxy materials or any incidental effects related to it. Like all polymer materials, epoxy resins are expected to become recyclable or reusable. As they are used in sophisticated technological devices, self-healing and recyclable materials have to be readily accessible. The controlled water uptake can be considered for this challenge. Epoxy resins offer an excellent platform for materials and systems that integrate stimuli-responsive motives to obtain life-like and intelligent materials with dynamically adaptive and interactive functions.

Author Contributions

Conceptualization, N.A.B., A.V.L. and V.V.Z.; writing—review and editing, N.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

Alexandr V. Latyshev was employed by the company Gazprom VNIIGAZ LLC. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Moy, P.; Karasz, F.E. The interactions of water with epoxy resins. In Water in Polymers; Rowland, S.P., Ed.; American Chemical Society: Washington, DC, USA, 1980; pp. 505–513. [Google Scholar]
  2. Krauklis, A.E.; Gagani, A.I.; Echtermeyer, A.T. Prediction of Orthotropic Hygroscopic Swelling of Fiber-Reinforced Composites from Isotropic Swelling of Matrix Polymer. J. Compos. Sci. 2019, 3, 10. [Google Scholar] [CrossRef]
  3. Starkova, O.; Gagani, A.I.; Karl, C.W.; Rocha, I.B.C.M.; Burlakovs, J.; Krauklis, A.E. Modelling of Environmental Ageing of Polymers and Polymer Composites—Durability Prediction Methods. Polymers 2022, 14, 907. [Google Scholar] [CrossRef] [PubMed]
  4. Moy, P.; Karasz, F.E. Epoxy-water interactions. Polym. Eng. Sci. 1980, 20, 315–319. [Google Scholar] [CrossRef]
  5. Nogueira, P.; Torres, A.; Abad, M.-J.; Cano, J.; Barral, L. Effect of water sorption on the structure and mechanical properties of an epoxy resin system. J. Appl. Polym. Sci. 2001, 80, 71–80. [Google Scholar] [CrossRef]
  6. De’Nève, B.; Shanahan, M.E.R. Water absorption by an epoxy resin and its effect on the mechanical properties and infra-red spectra. Polymer 1993, 34, 5099–5105. [Google Scholar] [CrossRef]
  7. Masoumi, S.; Valipour, H. Effects of moisture exposure on the crosslinked epoxy system: An atomistic study. Model. Simul. Mater. Sci. Eng. 2016, 24, 035011. [Google Scholar] [CrossRef]
  8. Morsch, S.; Lyon, S.; Greensmith, P.; Smith, S.D.; Gibbon, S.R. Mapping water uptake in organic coatings using AFM-IR. Faraday Discuss. 2015, 180, 527–542. [Google Scholar] [CrossRef]
  9. Bratasyuk, N.; Ostanin, S.A.; Mokeev, M.V.; Zuev, V.V. Water transport in epoxy/polyurethane interpenetrating networks. Polym. Adv. Technol. 2022, 33, 3173–3191. [Google Scholar] [CrossRef]
  10. Tan, Y. Sensing localised corrosion by means of electrochemical noise detection and analysis. Sens. Actuators B Chem. 2009, 139, 688–698. [Google Scholar] [CrossRef]
  11. Homborg, A.M.; Tinga, T.; Zhang, X.; van Westing, E.P.M.; Oonincx, P.J.; Ferrari, G.M.; de Wit, J.H.W.; Mol, J.M.C. Transient analysis through Hilbert spectra of electrochemical noise signals for the identification of localized corrosion of stainless steel. Electochim. Acta 2013, 104, 84–93. [Google Scholar] [CrossRef]
  12. Emeron, C.; Rios, E.C.; Zimer, A.M.; Mendes, P.C.D.; Freitas, M.B.J.; de Castro, E.V.R.; Mascaro, L.H.; Pereira, E.C. Corrosion of AISI 1020 steel in crude oil studied by the electrochemical noise measurements. Fuel 2015, 150, 325–333. [Google Scholar] [CrossRef]
  13. Homborg, A.M.; van Westing, A.M.; Tinga, T.; Zhang, X.; Oonincx, P.J.; Ferrari, G.M.; de Wit, J.H.W.; Mol, J.M.C. Novel time–frequency characterization of electrochemical noise data in corrosion studies using Hilbert spectra. Corros. Sci. 2013, 66, 97–110. [Google Scholar] [CrossRef]
  14. Maggana, C.; Pissis, P. Water sorption and diffusion studies in an epoxy resin system. J. Polym. Sci. Part B Polym. Phys. 1999, 37, 1165–1182. [Google Scholar] [CrossRef]
  15. Zhang, D.; Li, K.; Li, Y.; Sun, H.; Cheng, J.; Zhang, J. Characteristics of water absorption in amine-cured epoxy networks: A molecular simulation and experimental study. Soft Matter 2018, 14, 8740–8749. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, Y.; Adams, R.D.; da Silva, L.F.M. Absorption and glass transition temperature of adhesives exposed to water and toluene. Int. J. Adhes. Adhes. 2014, 50, 85–92. [Google Scholar] [CrossRef]
  17. Nayak, R.K.; Mahato, K.K.; Ray, B.C. Water absorption behavior, mechanical and thermal properties of nano TiO2 enhanced glass fiber reinforced polymer composites. Compos. Part A Appl. Sci. Manuf. 2016, 90, 736–747. [Google Scholar] [CrossRef]
  18. Soles, C.L.; Chang, F.T.; Gidley, D.W.; Yee, A.F. Contributions of the nanovoid structure to the kinetics of moisture transport in epoxy resins. J. Polym. Sci. Part B Polym. Phys. 2000, 38, 776–791. [Google Scholar] [CrossRef]
  19. Philippe, L.V.S.; Lyon, S.B.; Sammon, C.; Yarwood, J. Validation of electrochemical impedance measurements for water sorption into epoxy coatings using gravimetry and infra-red spectroscopy. Corros. Sci. 2008, 50, 887–896. [Google Scholar] [CrossRef]
  20. Cotugno, S.; Larobina, D.; Mensitieri, G.; Musto, P.; Ragosta, G. A novel spectroscopic approach to investigate transport processes in polymers: The case of water-epoxy system. Polymer 2001, 42, 6431–6438. [Google Scholar] [CrossRef]
  21. Wong, T.; Broutman, L. Moisture diffusion in epoxy resins Part I. Non-Fickian sorption processes. Polym. Eng. Sci. 1985, 25, 521–528. [Google Scholar] [CrossRef]
  22. Wong, K.J.; Low, K.O.; Israr, H.A.; Tamin, M.N. Thickness-dependent non-Fickian moisture absorption in epoxy molding compounds. Microelectron. Reliab. 2016, 65, 160–166. [Google Scholar] [CrossRef]
  23. Vanlandingham, M.R.; Eduljee, R.F.; Gillespie, J.W., Jr. Moisture diffusion in epoxy systems. J. Appl. Polym. Sci. 1999, 71, 787–798. [Google Scholar] [CrossRef]
  24. Coniglio, N.; Nguyen, K.; Kurji, R.; Gamboa, E. Characterizing water sorption in 100% solids epoxy coatings. Prog. Org. Coat. 2013, 76, 1168–1177. [Google Scholar] [CrossRef]
  25. Miszczyk, A.; Darowicki, K. Water Uptake in Protective Organic Coatings and Its Reflection in Measured Coating Impedance. Prog. Org. Coat. 2018, 124, 296–302. [Google Scholar] [CrossRef]
  26. Melo, R.; Santos, W.; de Lima, A.G.B.; Lima, W.; Silva, J.; Farias, R. Water Absorption Process in Polymer Composites: Theory Analysis and Applications. In Transport Phenomena in Multiphase Systems; Delgado, J.M.P.Q., Lima, A.G.B., Eds.; Springer: Cham, Switzerland, 2018; pp. 219–249. [Google Scholar] [CrossRef]
  27. Apeagyei, A.K.; Grenfell, J.R.A.; Airey, G.D. Application of Fickian and non-Fickian diffusion models to study moisture diffusion in asphalt mastics. Mater. Struct. 2015, 124, 1461–1474. [Google Scholar] [CrossRef]
  28. Guloglu, G.E.; Altan, M.C. Moisture Absorption of Carbon/Epoxy Nanocomposites. J. Compos. Sci. 2020, 4, 21. [Google Scholar] [CrossRef]
  29. Capiel, G.; Miccio, L.A.; Montemartini, P.E.; Schwartz, G.A. Water Diffusion and Hydrolysis Effect on the Structure and Dynamics of Epoxy-Anhydride Networks. Polym. Degrad. Stab. 2017, 143, 57–63. [Google Scholar] [CrossRef]
  30. Wind, M.M.; Lenderink, H.J.W. A capacitance study of pseudo-fickian diffusion in glassy polymer. Prog. Org. Coat. 1996, 28, 239–250. [Google Scholar] [CrossRef]
  31. Li, L.; Yu, Y.; Wu, Q.; Zhan, G.; Li, S. Effect of chemical structure on the water sorption of amine-cured epoxy resins. Corros. Sci. 2009, 51, 3000–3006. [Google Scholar] [CrossRef]
  32. Frank, K.; Childers, C.; Dutta, D.; Gidley, D.; Jackson, M.; Ward, S.; Maskell, R.; Wiggins, J. Fluid uptake behavior of multifunctional epoxy blends. Polymer 2013, 54, 403–410. [Google Scholar] [CrossRef]
  33. Gupta, V.B.; Drzal, L.T.; Rich, M.J. The physical basis of moisture transport in a cured epoxy resin system. J. Appl. Polym. Sci. 1985, 30, 4467–4493. [Google Scholar] [CrossRef]
  34. LaPlante, G.; Ouriadov, A.V.; Lee-Sullivan, P.; Balcom, B.J. Anomalous moisture diffusion in an epoxy adhesive detected by magnetic resonance imaging. J. Appl. Polym. Sci. 2008, 109, 1350–1359. [Google Scholar] [CrossRef]
  35. Legghe, E.; Aragon, E.; Bélec, L.; Margaillan, A.; Melot, D. Correlation between water diffusion and adhesion loss: Study of an epoxy primer on steel. Prog. Org. Coat. 2009, 66, 276–280. [Google Scholar] [CrossRef]
  36. Pineda, A.F.E.; Garcia, F.G.; Simões, A.Z.; da Silva, E.L. Mechanical properties, water absorption and adhesive properties of diepoxy aliphatic diluent-modified DGEBA/Cycloaliphatic amine networks on 316 L stainless steel. Int. J. Adhes. Adhes. 2016, 68, 205–211. [Google Scholar] [CrossRef]
  37. Gillet, C.; Tamssaouet, F.; Hassoune-Rhabbour, B.; Tchalla, T.; Nassiet, V. Parameters Influencing Moisture Diffusion in Epoxy-Based Materials during Hygrothermal Ageing—A Review by Statistical Analysis. Polymers 2022, 14, 2832. [Google Scholar] [CrossRef] [PubMed]
  38. Leger, R.; Roy, A.; Grandidier, J.C. Non-classical water diffusion in an industrial adhesive. Int. J. Adhes. Adhes. 2010, 30, 744–753. [Google Scholar] [CrossRef]
  39. Kim, J.-K.; Hu, C.; Woo, R.S.; Sham, M.-L. Moisture barrier characteristics of organoclay–epoxy nanocomposites. Compos. Sci. Technol. 2005, 65, 805–813. [Google Scholar] [CrossRef]
  40. Jacobs, P.M.; Jones, E.R. Diffusion of moisture into two-phase polymers. J. Mater. Sci. 1989, 24, 2343–2347. [Google Scholar] [CrossRef]
  41. Berens, A.; Hopfenberg, H. Diffusion and relaxation in glassy polymer powders: 2. Separation of diffusion and relaxation parameters. Polymer 1978, 19, 489–496. [Google Scholar] [CrossRef]
  42. Weitsman, Y. Diffusion with Time-Varying Diffusivity, with Application to Moisture-Sorption in Composites. J. Compos. Mater. 1976, 10, 193–204. [Google Scholar] [CrossRef]
  43. Roy, S.; Xu, W.X.; Park, S.J.; Liechti, K.M. Anomalous Moisture Diffusion in Viscoelastic Polymers: Modeling and Testing. J. Appl. Mech. 1999, 67, 391–396. [Google Scholar] [CrossRef]
  44. Carter, H.G.; Kibler, K.G. Langmuir-Type Model for Anomalous Moisture Diffusion In Composite Resins. J. Compos. Mater. 1978, 12, 118–131. [Google Scholar] [CrossRef]
  45. Grace, L.R.; Altan, M.C. Characterization of anisotropic moisture absorption in polymeric composites using hindered diffusion model. Compos. Part A Appl. Sci. Manuf. 2012, 43, 1187–1196. [Google Scholar] [CrossRef]
  46. Sun, L.; Boo, W.J.; Clearfield, A.; Sue, H.J.; Pham, H.Q. Barrier properties of model epoxy nanocomposites. J. Membr. Sci. 2008, 318, 129–136. [Google Scholar] [CrossRef]
  47. Liu, W.; Hoa, S.; Pugh, M. Water uptake of epoxy–clay nanocomposites: Model development. Compos. Sci. Technol. 2007, 67, 3308–3315. [Google Scholar] [CrossRef]
  48. Shen, C.H.; Springer, G.S. Moisture absorption and desorption of composite materials. J. Compos. Mater. 1976, 10, 2–20. [Google Scholar] [CrossRef]
  49. Starkova, O.; Chandrasekaran, S.; Schnoor, T.; Sevcenko, J.; Schulte, K. Anomalous water diffusion in epoxy/carbon nano-particle composites. Polym. Degrad. Stab. 2019, 164, 127–135. [Google Scholar] [CrossRef]
  50. Joannès, S.; Mazé, L.; Bunsell, A.R. A concentration-dependent diffusion coefficient model for water sorption in composite. Compos. Struct. 2014, 108, 111–118. [Google Scholar] [CrossRef]
  51. Glaskova, T.I.; Guedes, R.M.; Morais, J.J.; Aniskevich, A.N. A comparative analysis of moisture transport models as applied to an epoxy binder. Polym. Mech. 2007, 43, 377–388. [Google Scholar] [CrossRef]
  52. Kumar, A.; Roy, S. Modeling of anomalous moisture diffusion in nanographene reinforced thermoset polymers. Compos. Struct. 2015, 122, 1–7. [Google Scholar] [CrossRef]
  53. Guadagno, L.; Vertuccio, L. Resistive Response of Carbon Nanotube-Based Composites Subjected to Water Aging. Nanomaterials 2021, 11, 2183. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, W.; Hoa, S.; Pugh, M. Water uptake of epoxy–clay nanocomposites: Experiments and model validation. Compos. Sci. Technol. 2008, 68, 2066–2072. [Google Scholar] [CrossRef]
  55. Scott, P.; Lees, J.M. Water, saltwater and alkaline solution uptake in epoxy thin films. J. Appl. Polym. Sci. 2013, 130, 1898–1908. [Google Scholar] [CrossRef]
  56. Gao, C.; Zhou, C. Moisture Absorption and Cyclic Absorption–Desorption Characters of Fibre-Reinforced Epoxy Composites. J. Mater. Sci. 2019, 54, 8289–8301. [Google Scholar] [CrossRef]
  57. Sugiman, S.; Salman, S.; Maryudi, M. Effects of volume fraction on water uptake and tensile properties of epoxy filled with inorganic fillers having different reactivity to water. Mater. Today Commun. 2020, 24, 101360. [Google Scholar] [CrossRef]
  58. Suri, C.; Perreux, D. The effects of mechanical damage in a glass fibre/epoxy composite on the absorption rate. Compos. Eng. 1995, 5, 415–424. [Google Scholar] [CrossRef]
  59. Popineau, S.; Rondeau-mouro, C.; Sulpice-gaillet, C.; Shanahan, M.E.R. Free/bound water absorption in an epoxy adhesive. Polymer 2005, 46, 10733–10740. [Google Scholar] [CrossRef]
  60. Wu, P.Y.; Siesler, H.W. Water diffusion into epoxy resin: A 2D correlation ATR-FTIR investigation. Chem. Phys. Lett. 2003, 374, 74–78. [Google Scholar] [CrossRef]
  61. Wang, M.; Xu, X.; Ji, J.; Yang, Y.; Shen, J.; Ye, M. The hygrothermal aging process and mechanism of the novolac epoxy resin. Compos. Part B Eng. 2016, 107, 1–8. [Google Scholar] [CrossRef]
  62. Wang, H.; Liu, Y.; Zhang, J.; Li, T.; Hu, Z.; Yu, Y. Effect of curing conversion on the water sorption, corrosion resistance and thermo-mechanical properties of epoxy resin. RSC Adv. 2015, 5, 11358–11370. [Google Scholar] [CrossRef]
  63. Liu, Z.; Huo, J.; Yu, Y. Water absorption behavior and thermal-mechanical properties of epoxy resins cured with cardanol-based novolac resins and their esterified ramifications. Mater. Today Commun. 2017, 10, 80–94. [Google Scholar] [CrossRef]
  64. Liu, M.; Wu, P.; Ding, Y.; Li, S. Study on diffusion behavior of water in epoxy resins cured by active ester. Phys. Chem. Chem. Phys. 2003, 5, 1848–1852. [Google Scholar] [CrossRef]
  65. Liu, M.; Wu, P.; Ding, Y.; Chen, G.; Li, S. Two-Dimensional (2D) ATR−FTIR Spectroscopic Study on Water Diffusion in Cured Epoxy Resins. Macromolecules 2002, 35, 5500–5507. [Google Scholar] [CrossRef]
  66. Li, L.; Yu, Y.; Su, H.; Zhan, G.; Li, S.; Wu, P. The diffusion mechanism of water transport in amine-cured epoxy networks. Appl. Spectrosc. 2010, 64, 458–465. [Google Scholar] [CrossRef] [PubMed]
  67. Li, L.; Liu, M.; Li, S. Dynamic process of water sorption in a thermoplastic modified epoxy resin system. J. Phys. Chem. B 2004, 108, 4601–4606. [Google Scholar] [CrossRef]
  68. Ji, W.-G.; Hu, J.-M.; Liu, L.; Zhang, J.-Q.; Cao, C.-N. Water uptake of epoxy coatings modified with γ-APS silane monomer. Prog. Org. Coat. 2006, 57, 439–443. [Google Scholar] [CrossRef]
  69. Ha, H.M.; Alfantazi, A. On the role of water, temperature, and glass transition in the corrosion protection behavior of epoxy coatings for underground pipelines. J. Coat. Technol. Res. 2015, 12, 1095–1110. [Google Scholar] [CrossRef]
  70. Feng, J.; Berger, K.R.; Douglas, E.P. Water vapor transport in liquid crystalline and non-liquid crystalline epoxies. J. Mater. Sci. 2004, 39, 3413–3423. [Google Scholar] [CrossRef]
  71. Takeshita, Y.; Becker, E.; Sakata, S.; Miwa, T.; Sawada, T. States of water absorbed in water-borne urethane/epoxy coatings. Polymer 2014, 55, 2505–2513. [Google Scholar] [CrossRef]
  72. Xiao, G.Z.; Delamar, M.; Shanahan, M.E.R. Irreversible interactions between water and DGEBA/DDA epoxy resin during hygrothermal aging. J. Appl. Polym. Sci. 1997, 65, 449–458. [Google Scholar] [CrossRef]
  73. Choi, S.; Phantu, A.; Douglas, E.P. Evaluation of the complex hygrothermal behaviors of epoxy-amine systems. J. Appl. Polym. Sci. 2012, 125, 3778–3787. [Google Scholar] [CrossRef]
  74. Cabanelas, J.C.; Prolongo, S.G.; Serrano, B.; Bravo, J.; Baselga, J. Water absorption in polyaminosiloxane-epoxy thermosetting polymers. J. Mater. Process. Technol. 2003, 143–144, 311–315. [Google Scholar] [CrossRef]
  75. Cotugno, S.; Mensitieri, G.; Musto, P.; Sanguigno, L. Molecular Interactions in and Transport Properties of Densely Cross-Linked Networks: A Time-Resolved FT-IR Spectroscopy Investigation of the Epoxy/H2O System. Macromolecules 2005, 38, 801–811. [Google Scholar] [CrossRef]
  76. Musto, P.; Ragosta, G.; Scarinzi, G.; Mascia, L. Probing the molecular interactions in the diffusion of water through epoxy and epoxy-bismaleimide networks. J. Polym. Sci. Pol. Phys. 2002, 40, 922–938. [Google Scholar] [CrossRef]
  77. Idolor, O.; Guha, R.D.; Berkowitz, K.; Geiger, C.; Davenport, M.; Grace, L. Polymer-water interactions and damage detection in polymer matrix composites. Compos. B Eng. 2021, 211, 108637. [Google Scholar] [CrossRef]
  78. Musto, P.; Ragosta, G.; Mascia, L. Vibrational Spectroscopy Evidence for the Dual Nature of Water Sorbed into Epoxy Resins. Chem. Mater. 2000, 12, 1331–1341. [Google Scholar] [CrossRef]
  79. Musto, P.; Mascia, L.; Ragosta, G.; Scarinzi, G.; Villano, P. The transport of water in a tetrafunctional epoxy resin by near-infrared Fourier transform spectroscopy. Polymer 2000, 41, 565–574. [Google Scholar] [CrossRef]
  80. Linde, E.; Giron, N.H.; Celina, M.C. Water diffusion with temperature enabling predictions for sorption and transport behavior in thermoset materials. Polymer 2018, 153, 653–667. [Google Scholar] [CrossRef]
  81. Zargarnezhad, H.; Asselin, E.; Wong, D.; Lam, C.N.C. A Critical Review of the Time-Dependent Performance of Polymeric Pipeline Coatings: Focus on Hydration of Epoxy-Based Coatings. Polymers 2021, 13, 1517. [Google Scholar] [CrossRef]
  82. Noda, I. Two-Dimensional Infrared (2D IR) Spectroscopy: Theory and Applications. Appl. Spectrosc. 1990, 44, 550–561. [Google Scholar] [CrossRef]
  83. Li, L.; Chen, Y.; Li, S. Water Diffusion Behavior in Epoxy Resins with Various Fluorine Contents. Appl. Spectrosc. 2006, 60, 392–397. [Google Scholar] [CrossRef]
  84. Luo, S.; Leisen, J.; Wong, C. Study on Mobility of Water and Polymer Chain in Epoxy and Its Influence on Adhesion. J. Appl. Polym. Sci. 2002, 85, 1–8. [Google Scholar] [CrossRef]
  85. Soles, C.L.; Yee, A.F. A Discussion of the Molecular Mechanisms of Moisture Transport in Epoxy Resins. J. Polym. Sci. Part B Polym. Phys. 2000, 38, 792–802. [Google Scholar] [CrossRef]
  86. Spiess, H.W. 50th Anniversary Perspective: The Importance of NMR Spectroscopy to Macromolecular Science. Macromolecules 2017, 50, 1761–1777. [Google Scholar] [CrossRef]
  87. Bouznik, V.M.; Morozov, E.V.; Avilova, I.A.; Volkov, V.I. NMR Applications for Polymer Composite Materials Moisture Uptake Investigation. Appl. Magn. Reson. 2016, 47, 321–334. [Google Scholar] [CrossRef]
  88. Baukh, V.; Huinink, H.P.; Adan, O.C.G.; Erich, S.J.F.; Van Der Ven, L.G. Water–Polymer Interaction during Water Uptake. Macromolecules 2011, 44, 4863–4871. [Google Scholar] [CrossRef]
  89. Baukh, V.; Huinink, H.; Adan, O.; Erich, S.; van der Ven, L.; Ondergrond, T.B. NMR Imaging of Water Uptake in Multilayer Polymeric Films: Stressing the role of Mechanical Stress. Macromolecules 2010, 43, 3882. [Google Scholar] [CrossRef]
  90. Klotz, J.; Brostow, W.; Hess, M.; Veeman, W.S. Epoxy and glass composites in water studied with 2H-NMR. Polym. Eng. Sci. 1996, 36, 1129–1133. [Google Scholar] [CrossRef]
  91. Jelinski, L.W.; Dumais, J.J.; Cholli, A.L.; Ellis, T.S.; Karasz, F.E. Nature of the water-epoxy interaction. Macromolecules 1985, 18, 1091–1095. [Google Scholar] [CrossRef]
  92. Dutheillet, Y.; Mantle, M.; Vesely, D.; Gladden, L. Diffusion of water-acetic acid mixture in epoxy. J. Polym. Sci. B 1999, 37, 3328–3336. [Google Scholar] [CrossRef]
  93. Yuan, X.; Yue, Z.F.; Chen, X.; Wen, S.F.; Li, L. Effect of mixture ratio on water uptake and corrosion performance of silicone-epoxy hybrid coatings coated 2024 Al-alloy. Prog. Org. Coat. 2015, 78, 168–175. [Google Scholar] [CrossRef]
  94. Shreepathi, S.; Naik, S.M.; Vattipalli, M.R. Water transportation through organic coatings: Correlation between electrochemical impedance measurements, gravimetry, and water vapor permeability. J. Coat. Technol. Res. 2012, 9, 411–422. [Google Scholar] [CrossRef]
  95. Elkebir, Y.; Mallarino, S.; Trinh, D.; Touzain, S. Effect of physical ageing onto the water uptake in epoxy coatings. Electrochim. Acta 2020, 337, 135766. [Google Scholar] [CrossRef]
  96. Ding, R.; Jiang, J.; Gui, T. Study of impedance model and water transport behavior of modified solvent-free epoxy anticorrosion coating by EIS. J. Coat. Technol. Res. 2016, 13, 501–515. [Google Scholar] [CrossRef]
  97. Fan, C.; Shi, J.; Dilger, K. Water uptake and interfacial delamination of an epoxy-coated galvanized steel: An electrochemical impedance spectroscopic study. Prog. Org. Coat. 2019, 137, 105333. [Google Scholar] [CrossRef]
  98. Ding, R.; Cong, W.; Jiang, J.; Gui, T. Study of impedance model and water transport behavior of modified solvent-free epoxy anti-corrosion coating by EIS II. J. Coat. Technol. Res. 2016, 13, 981–997. [Google Scholar] [CrossRef]
  99. Zhang, J.T.; Hu, J.M.; Zhang, J.Q.; Cao, C.N. Studies of water transport behavior and impedance models of epoxy-coated metals in NaCl solution by EIS. Prog. Org. Coat. 2004, 51, 145–151. [Google Scholar] [CrossRef]
  100. Yang, H.-Q.; Zhang, Q.; Li, Y.-M.; Liu, G.; Huang, Y. Effects of Immersion Temperature on the Performance of a Marine Epoxy-Based Organic Coating for Ballast Tanks. J. Mater. Eng. Perform. 2021, 30, 4458–4465. [Google Scholar] [CrossRef]
  101. Hakala, K.; Vatanparast, R.; Vuorimaa, E.; Lemmetyinen, H. Monitoring water uptake of polyurethanes by in situ fluorescence technique. J. Appl. Polym. Sci. 2001, 82, 1593–1599. [Google Scholar] [CrossRef]
  102. Sung, C.S.P.; Sung, N.H. Fluorescence characterization of cure and water uptake in polymers and composites. Mater. Sci. Eng. A 1993, 162, 241–247. [Google Scholar] [CrossRef]
  103. Miller, K.E.; Krueger, R.H.; Torkelson, J.M. Mobility-sensitive fluorescence probes for quantitative monitoring of water sorption and diffusion in polymer coatings. J. Polym. Sci. Part B Polym. Phys. 1995, 33, 2343–2349. [Google Scholar] [CrossRef]
  104. Martin, O.; Pastoriza, A.; Mikes, F.; Baselga, J. Water absorption in epoxy resins by electronic energy transfer. Polym. Int. 2002, 51, 1207–1210. [Google Scholar] [CrossRef]
  105. Olmos, D.; Lopez-Moron, R.; Gonzalez-Benito, J. The nature of the glass fiber surface and its effect in the water absorption of glass fiber/epoxy composites. The use of fluorescence to obtain information at the interface. Compos. Sci. Technol. 2006, 66, 2758–2768. [Google Scholar] [CrossRef]
  106. González-Benito, J.; Mikes, F.; Bravo, J.; Aznar, A.J.; Baselga, J. Fluorescence monitoring of curing process and water accessibility at glass fiber/epoxy interphase on composite materials. J. Macromol. Sci. Part B 2001, 40, 429–441. [Google Scholar] [CrossRef]
  107. Gonzales-Benito, J.; Bravo, J.; Mikes, F.; Baselga, J. Fluorescence labels to monitor water absorption in epoxy resins. Polymer 2003, 44, 653–659. [Google Scholar] [CrossRef]
  108. Wöll, D. Fluorescence correlation spectroscopy in polymer science. RSC Adv. 2014, 4, 2447–2465. [Google Scholar] [CrossRef]
  109. Dazzi, A.; Prater, C.B. AFM-IR: Technology and applications in nanoscale infrared spectroscopy and chemical imaging. Chem. Rev. 2017, 117, 5146–5173. [Google Scholar] [CrossRef]
  110. Morsch, S.; Lyon, S.; Greensmith, P.; Smith, S.D.; Gibbon, S.R. Water transport in an epoxy-phenolic coating. Prog. Org. Coat. 2015, 78, 293–299. [Google Scholar] [CrossRef]
  111. Morsch, S.; Stuart, S.E.; Lyon, B.; Gibbon, S.R.; Irwin, M. The location of adsorbed water in pigmented epoxy-amine coatings. Prog. Org. Coat. 2022, 173, 107223. [Google Scholar] [CrossRef]
  112. Wang, J.; Gong, J.; Gong, Z.; Yan, X.; Wang, B.; Wu, Q.; Li, S. Effect of curing agent polarity on water absorption and free volume in epoxy resin studied by PALS. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2010, 268, 2355–2361. [Google Scholar] [CrossRef]
  113. Suzuki, T.; Oki, Y.; Numajiri, M.; Miura, T.; Kondo, K. Free-volume characteristics and water absorption of novolac epoxy resins investigated by positron annihilation. Polymer 1996, 37, 3025–3030. [Google Scholar] [CrossRef]
  114. Jackson, M.; Kaushik, M.; Nazarenko, S.; Ward, S.; Maskell, R.; Wiggins, J. Effect of free volume hole-size on fluid ingress of glassy epoxy networks. Polymer 2011, 52, 4528–4535. [Google Scholar] [CrossRef]
  115. Roggero, A.; Caussé, N.; Dantras, E.; Villareal, L.; Santos, A.; Pébère, N. In situ study of the temperature activated kinetics of water sorption in an epoxy varnish. Polymer 2021, 213, 123206. [Google Scholar] [CrossRef]
  116. Zinck, F.; Gerard, J.-F. Polyepoxide-water interactions: Influence of the chemical structure of the network. Polym. Degrad. Stab. 2008, 93, 1231–1237. [Google Scholar] [CrossRef]
  117. Le Guen-Geffroy, A.; Le Gac, P.-Y.; Habert, B.; Davies, P. Physical ageing of epoxy in a wet environment: Coupling between plasticization and physical ageing. Polym. Degrad. Stab. 2019, 168, 108947. [Google Scholar] [CrossRef]
  118. Ilioni, A.; Le Gac, P.-Y.; Badulescu, C.; Thévenet, D.; Davies, P. Prediction of Mechanical Behaviour of a Bulk Epoxy Adhesive in a Marine Environment. J. Adhes. 2019, 95, 64–84. [Google Scholar] [CrossRef]
  119. Wilmers, J.; Bargmann, S. Simulation of non-classical diffusion in polymers. Heat Mass Transfer. 2014, 50, 1543–1552. [Google Scholar] [CrossRef]
  120. Damian, C.; Espuche, E.; Escoubes, M. Influence of three ageing types (thermal oxidation, radiochemical and hydrolytic ageing) on the structure and gas transport properties of epoxy-amine networks. Polym. Degrad. Stab. 2001, 72, 447–458. [Google Scholar] [CrossRef]
  121. Frisch, H.L. Sorption and transport in glassy polymer—A review. Polym. Eng. Sci. 1980, 20, 2–13. [Google Scholar] [CrossRef]
  122. Toscano, A.; Pitarresi, G.; Scafidi, M.; Di Filippo, M.; Spadaro, G.; Alessi, S. Water diffusion and swelling stresses in highly crosslinked epoxy matrices. Polym. Degrad. Stabl. 2016, 133, 255–263. [Google Scholar] [CrossRef]
  123. Han, S.O.; Drzal, L.T. Water absorption effects on hydrophilic polymer matrix of carboxyl functionalized glucose resin and epoxy resin. Eur. Polym. J. 2003, 39, 1791–1799. [Google Scholar] [CrossRef]
  124. Chateauminois, A.; Chabert, B.; Soulier, J.P.; Vincent, L. Dynamic-mechanical analysis of epoxy composites plasticized by water—Artifact and reality. Polym. Compos. 1995, 16, 288–296. [Google Scholar] [CrossRef]
  125. Perrin, F.X.; Nguyen, M.H.; Vernet, J.L. Water transport in epoxy–aliphatic amine networks—Influence of curing cycles. Eur. Polym. J. 2009, 45, 1524–1534. [Google Scholar] [CrossRef]
  126. Yang, B.; Huang, W.M.; Li, C.; Li, L. Effects of moisture on the thermomechanical properties of a polyurethane shape memory polymer. Polymer 2006, 47, 1348–1356. [Google Scholar] [CrossRef]
  127. Xian, G.; Karbhari, V.M. Segmental relaxation of water-aged ambient cured epoxy. Polym. Degrad. Stab. 2007, 92, 1650–1659. [Google Scholar] [CrossRef]
  128. Ivanova, K.; Pethrick, R.; Affrossman, S. Investigation of hydrothermal ageing of a filled rubber toughened epoxy resin using dynamic mechanical thermal analysis and dielectric spectroscopy. Polymer 2000, 41, 6787–6796. [Google Scholar] [CrossRef]
  129. Sugisaki, M.; Suga, H.; Seki, S. Calorimetric Study of the Glassy State. IV. Heat Capacities of Glassy Water and Cubic Ice. Bull. Chem. Soc. Jpn. 1968, 41, 2591–2599. [Google Scholar] [CrossRef]
  130. Hallbrucker, A.; Mayer, E.; Johari, G.P. The heat capacity and glass transition of hyperquenched glassy water. Philos. Mag. 1989, 60, 179–187. [Google Scholar] [CrossRef]
  131. Kelley, F.N.; Bueche, F. Viscosity and glass temperature relations for polymer-diluent systems. J. Polym. Sci. 1961, 50, 549–556. [Google Scholar] [CrossRef]
  132. McKague, E.L.; Reynolds, J.D.; Halkias, J.E. Swelling and glass transition relations for epoxy matrix material in humid environments. J. Appl. Polym. Sci. 1978, 22, 1643–1654. [Google Scholar] [CrossRef]
  133. Carfagna, C.; Apicella, A.; Nicolais, L. The effect of the prepolymer composition of amino-hardened epoxy 520 resins on the water sorption behavior and plasticization. J. Appl. Polym. Sci. 1982, 27, 105–112. [Google Scholar] [CrossRef]
  134. Zhou, J.; Lucas, J.P. Hygrothermal effects of epoxy resin. Part II: Variations of glass transition temeprature. Polymer 1999, 40, 5505–5512. [Google Scholar] [CrossRef]
  135. Park, H.; Yang, S.; Han, J.; Cho, M. Prediction of quasistatic constitutive equations of moisture-absorbed epoxy polymers using atomistic simulations. Extrem. Mech. Let. 2020, 41, 100983. [Google Scholar] [CrossRef]
  136. Thomson, K.W.; Wong, T.; Broutman, L.J. The plasticization of an epoxy resin by dibutylphthalate and water. Polym. Eng. Sci. 1984, 24, 1270–1276. [Google Scholar] [CrossRef]
  137. Simha, R.; Boyer, R.F. On a General Relation Involving the Glass Temperature and Coefficients of Expansion of Polymers. J. Chem. Phys. 1962, 37, 1003–1007. [Google Scholar] [CrossRef]
  138. Sharp, N.; Li, C.H.; Strachan, A.; Adams, D.; Pipes, R.B. Effects of water on epoxy cure kinetics and glass transition temperature utilizing molecular dynamics simulations. J. Polym. Sci. Part B Polym. Phys. 2017, 55, 1150–1159. [Google Scholar] [CrossRef]
  139. Ray, B.C.; Prusty, R.; Rathore, D.K. Fibrous Polymeric Composites; Environmental Degradation and Damage; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  140. Niu, Y.-F.; Yan, Y.; Yao, J.-W. Hygrothermal aging mechanism of carbon fiber/epoxy resin composites based on quantitative characterization of interface structure. Polym. Test. 2021, 94, 107019. [Google Scholar] [CrossRef]
  141. Kececi, E.; Asmatulu, R. Effects of moisture ingressions on mechanical properties of honeycomb-structured fiber composites for aerospace applications. Int. J. Adv. Manuf. Technol. 2017, 88, 459–470. [Google Scholar] [CrossRef]
  142. Khotbehsara, M.M.; Manalo, A.; Thiru Aravinthan, T.; Ferdous, W.; Nguyen, K.T.Q.; Hota, G. Ageing of particulate-filled epoxy resin under hygrothermal conditions. Constr. Build. Mater. 2020, 249, 118846. [Google Scholar] [CrossRef]
  143. Böer, P.; Holliday, L.; Kang, T.H.-K. Independent environmental effects on durability of fiber-reinforced polymer wraps in civil applications: A review. Constr. Build. Mater. 2013, 48, 360–370. [Google Scholar] [CrossRef]
  144. Woo, M.S.W.; Piggott, M.R. Water absorption of resins and composites: II. Diffusion in carbon and glass reinforced epoxies. J. Compos. Technol. Res. 1987, 9, 162–166. [Google Scholar] [CrossRef]
  145. Jones, F.R. Durability of reinforced plastics in liquid environments. In Reinforced Plastics Durability; CRC Press: Boca Raton, FL, USA, 1999; pp. 70–110. [Google Scholar] [CrossRef]
  146. Schutte, C.L. Environmental durability of glass–fiber composites. Mater. Sci. Eng. 1994, R13, 265–324. [Google Scholar] [CrossRef]
  147. Marom, G.; Broutman, L.J. Moisture penetration into composites under external stress. Polym. Compos. 1981, 2, 132–136. [Google Scholar] [CrossRef]
  148. Youssef, G.; Freour, S.; Jacquemin, F. Effects of moisture-dependent properties of constituents on the hygroscopic stresses in composite structures. Mech. Compos. Mater. 2009, 45, 369–380. [Google Scholar] [CrossRef]
  149. Freour, S.; Jacquemin, F. Mechanical States Induced by Moisture Diffusion in Organic Matrix Composites: Coupled Scale Transition Models. In Composite Materials in Engineering Structures; Nova Science: New York, NY, USA, 2010. [Google Scholar]
  150. Bledzki, A.R.; Spaude, R.; Ehrenstein, G.W. Corrosion phenomena in glass fibers and glass fiber reinforced thermosetting resins. Comp. Sci. Technol. 1985, 23, 263–285. [Google Scholar] [CrossRef]
Figure 1. Typical curve for Langmuir-type model.
Figure 1. Typical curve for Langmuir-type model.
Coatings 14 00054 g001
Figure 2. Band deconvolution corresponding to OH-group vibrations in the range of 3800–3000 cm−1.
Figure 2. Band deconvolution corresponding to OH-group vibrations in the range of 3800–3000 cm−1.
Coatings 14 00054 g002
Figure 3. Band deconvolution corresponding to OH-group vibrations in the range of 7800–6000 cm−1.
Figure 3. Band deconvolution corresponding to OH-group vibrations in the range of 7800–6000 cm−1.
Coatings 14 00054 g003
Figure 4. Typical 2D correlation spectra of water absorption into epoxy composition in the 3700–3100 cm−1 range.
Figure 4. Typical 2D correlation spectra of water absorption into epoxy composition in the 3700–3100 cm−1 range.
Coatings 14 00054 g004
Figure 5. Example of a deconvoluted 2H NMR peak.
Figure 5. Example of a deconvoluted 2H NMR peak.
Coatings 14 00054 g005
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

Bratasyuk, N.A.; Latyshev, A.V.; Zuev, V.V. Water in Epoxy Coatings: Basic Principles of Interaction with Polymer Matrix and the Influence on Coating Life Cycle. Coatings 2024, 14, 54. https://doi.org/10.3390/coatings14010054

AMA Style

Bratasyuk NA, Latyshev AV, Zuev VV. Water in Epoxy Coatings: Basic Principles of Interaction with Polymer Matrix and the Influence on Coating Life Cycle. Coatings. 2024; 14(1):54. https://doi.org/10.3390/coatings14010054

Chicago/Turabian Style

Bratasyuk, Nikita A., Alexandr V. Latyshev, and Vjacheslav V. Zuev. 2024. "Water in Epoxy Coatings: Basic Principles of Interaction with Polymer Matrix and the Influence on Coating Life Cycle" Coatings 14, no. 1: 54. https://doi.org/10.3390/coatings14010054

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