Skip to main content
JACS Au logoLink to JACS Au
. 2022 Jun 9;2(7):1522–1542. doi: 10.1021/jacsau.2c00120

Network Formation and Physical Properties of Epoxy Resins for Future Practical Applications

Atsuomi Shundo †,*, Satoru Yamamoto , Keiji Tanaka †,‡,*
PMCID: PMC9327093  PMID: 35911459

Abstract

graphic file with name au2c00120_0025.jpg

Epoxy resins are used in various fields in a wide range of applications such as coatings, adhesives, modeling compounds, impregnation materials, high-performance composites, insulating materials, and encapsulating and packaging materials for electronic devices. To achieve the desired properties, it is necessary to obtain a better understanding of how the network formation and physical state change involved in the curing reaction affect the resultant network architecture and physical properties. However, this is not necessarily easy because of their infusibility at higher temperatures and insolubility in organic solvents. In this paper, we summarize the knowledge related to these issues which has been gathered using various experimental techniques in conjunction with molecular dynamics simulations. This should provide useful ideas for researchers who aim to design and construct various thermosetting polymer systems including currently popular materials such as vitrimers over epoxy resins.

Keywords: thermoset, cross-linking, network, glass, interface, adhesive

1. Introduction

In 1907, Leo Baekeland synthesized a phenol–formaldehyde resin, so-called Bakelite,1 as the first example of a class of thermosetting polymers.2 Thermosets are generally formed from a liquid mixture of monomer molecules, having multifunctional groups, which can react with each other to cross-link three-dimensionally. One of the advantages of thermosets over thermoplastics is that their precursor can be a reaction mixture having low viscosity, which offers good processability in injection and molding.3 Epoxy resins are one of the most versatile categories of thermosets derived from the precursor having oxirane or epoxy groups. They were discovered in 1939 by Prileschajew.3 Thanks to their excellent thermal and mechanical properties, epoxy resins have been widely applied in a range of fields.46 The global market for epoxy resins is driven by the increasing demand in the fields of chemistry, automotive, aerospace, civil engineering, leisure, electrical, marine, and many others.7

Epoxy resins can be commonly obtained by chemical reactions of epoxy compounds with initiators or curing agents (hardeners). The high reactivity of the epoxy groups toward a wide variety of functional groups has attracted much attention from chemists as well as chemical engineers. So far, the reaction kinetics for various types of epoxy compounds, initiators, and curing agents have been examined.8 On the other hand, researchers working with polymer materials have focused on the relationship between the network structure and physical properties. In particular, gaining an understanding of the network formation and the change in physical states (liquid, rubbery, and glassy solids) involved in the reaction process has been the subject of intensive research for many years.9

Epoxy resins are generally a glass-forming material and are often in a glassy state at room temperature. The mechanical relaxation associated with the network architecture of fully cured epoxy resins has been extensively studied.10 Notably, direct characterization of the network architecture is difficult or even impossible because of their infusibility at higher temperatures and insolubility in organic solvents. These difficulties are particularly acute for the “buried” interface, at which the epoxy resins come into contact with a solid. Therefore, atomistic and coarse-grained molecular dynamics (MD) simulations are considered to be a powerful analytical tool.

From this perspective, we highlight important aspects that should be considered in current research trends. First, we briefly introduce the chemistry of epoxy groups. Then the reaction kinetics associated with the network formation and the physical state change are presented. The network and physical properties including dynamic heterogeneity of the fully cured epoxy resins are discussed. Finally, we summarize recent applications in the practical and industrial fields.

2. Chemistry of Epoxy Resins

2.1. Classification of Epoxy Resins

The term of “epoxy” is used to describe a range of monomers containing an epoxy group, while “epoxy resins” refers to a class of molecules containing at least two epoxy groups. The material obtained after the curing reaction is commonly referred to as “epoxy resin” even if it no longer contains epoxy groups. Figure 1 shows examples of widely used epoxy monomers. Diglycidyl ether of bisphenol A (DGEBA), which can be obtained from the reaction between bisphenol A and epichlorohydrin in the presence of sodium hydroxide, is one of the most common precursors for epoxy resins.11 Multifunctional epoxy monomers are also widely used because they tend to increase the cross-linking density.1215 For instance, N,N,N',N'-tetraglycidyl-4,4'-methylenedianiline (TGMDA) is a typical monomer extensively used in aerospace composites.16 Polyglycidyl derivatives of phenolic prepolymers are also common, known to yield epoxy resins with a higher glass transition temperature (Tg) and high resistance to thermal degradation.17,18 Cycloaliphatic resins are another class of epoxy resins of great interest.19,20 They have less tendency toward yellowing than aromatic resins.21 In addition, their low viscosity and electrical loss properties have made them useful commercially in electrical and electronic applications.22,23 Such examples can be seen for 3,4-epoxycyclohexylmethyl 3,4-epoxycyclohexanecarboxylate (ECC) and bis[3,4-epoxycyclohexylmethyl] adipate (BECHMA).

Figure 1.

Figure 1

Typical monomers used as a precursor for epoxy resins.

2.2. Type of Reactions and Curing Agents

An oxirane in an epoxy monomer is a class of three-membered ring. Such a small ring exhibits high reactivity dominated by the effect of ring strain, whose strain energy is estimated to be about 115 kJ·mol–1.24 Thus, epoxy monomers can generate a cross-linked network structure by either chain-growth ring-opening polymerization or step-growth polymerization, depending on the type of curing agent.25,26 Chain-growth ring-opening polymerization can be further classified into cationic and anionic polymerizations.

Panel a of Figure 2 shows the scheme for cationic polymerization of epoxy monomers. The propagation reaction proceeds via an active oxonium at the end of the growing chain. Common initiators are boron trifluoride (BF3) complexes27,28 and onium salts including diaryliodonium, triarylsulfonium, or phosphonium salts.2932 Panel b of Figure 2 shows the scheme for anionic polymerization which is generally initiated by imidazoles33 or highly reactive tertiary amines.34 It is based on the formation of alkoxide, which then reacted with an epoxy monomer, leading to the generation of another alkoxide. In recent years, several authors have reported that ionic liquids such as dialkylimidazolium ions can behave as an initiator.35 Here, it is noteworthy that epoxies can be co-polymerized with other cyclic monomers. In particular, the curing of epoxy–anhydride formulations, using tertiary amines as an initiator, has been demonstrated as an alternating epoxy–anhydride anionic polymerization.3638

Figure 2.

Figure 2

Reaction schemes of chain-growth ring-opening polymerization with (a) cationic and (b) anionic modes.

The step-growth ring-opening polymerization of epoxy monomers can be performed using amines, acids, isocyanates, and mercaptans.25,3943 Of these, amines have been widely used. In this case, the reactivity and thereby the kinetics of the curing reaction are determined by the electrophilicity of the epoxy group and the nucleophilicity of the amino group. Panel a of Figure 3 shows the reaction scheme of epoxy monomers with an amine. A primary amino group first reacts with an epoxy group. This produces a secondary amino group that further reacts with another epoxy group and then a tertiary amino group is generated.44 Since the resultant tertiary amino group gives a branching or cross-linking structure, the amines used are often referred to as “curing agents” or “hardeners”. Here, it should be noted that a hydroxy group generated by the ring-opening reaction can form a hydrogen bond with an oxygen atom in unreacted epoxy monomers.45 The formation of the hydrogen bond promotes the nucleophilic attack of an amino group to an epoxy group, as shown in Figure 3b. Such a reaction has often been regarded as an “autocatalytic reaction” and has been extensively studied.46,47

Figure 3.

Figure 3

Reaction scheme of step growth ring opening polymerization of epoxy and amine (a) without and (b) with autocatalytic reaction. Panel b only shows the first step reaction.

A wide variety of amines have been used as curing agents for epoxy resins,25 and Figure 4 shows some common examples. Basically, aromatic amines such as 4,4'-diaminodiphenylmethane (DDM) and 4,4'-diaminodiphenylsulfone (DDS) are less reactive than aliphatic ones (diethylenetriamine, DETA) because of the weaker nucleophilicity of the epoxy groups. Dicyandiamide (DICY) is one of the most commonly used “latent” curing agents, which do not react with an epoxy monomer unless the temperature increases.48 DICY is a solid with a high melting temperature of ca. 460 K and is insoluble in most epoxy monomers at room temperature. Thus, a mixture of DICY and epoxy monomer has excellent stability and can be cured once the temperature exceeds the melting point.49 Such a feature leads to a long storage lifetime and makes it easier to handle.50 As an alternative to DICY, dihydrazides are also known as a latent curing agent. Most dihydrazides can be dispersed as a solid in a liquid epoxy monomer because they possess a high crystalline feature. In addition, the nucleophilicity of amino groups is moderately reduced by the directly adjacent NH group.51 Notably, dihydrazides with various chemical structures, which are easily obtained from the corresponding diacids, are available.25

Figure 4.

Figure 4

Typical hardeners used for epoxy resins.

To obtain a single-phase system with latent properties, several researchers have proposed the protection and deprotection of amino groups in the curing agents.52,53 For instance, a ketone-based imine has been developed as a water-initiated latent agent.54Figure 5 depicts the regeneration of an amine from a ketimine. The nucleophilicity of the ketimine is low enough for a long shelf life although it is an enamine–imine tautomerism.55 Once the ketimine is exposed to atmospheric moisture, the imine hydrolysis regenerates amine, which can react with epoxies followed by the curing reactions.56 The nucleophilicity control of amines based on the protection/deprotection approach has been demonstrated with 2-nitrobenzyl carbamates,57,58O-acyloximes,59 and N-aryl-N,N′-dialkyl urea.60 One of the interesting approaches is a thermally activated single-component system proposed by Fréchet and co-workers, as shown in Figure 6.61 The single precursor for epoxy resin contains both epoxy and diamine groups held by thermally degradable carbamate linkages. The precursor molecule has a long shelf life at room temperature. Upon heating, carbamate decomposition provides the eliminated alkene with epoxy groups, primary diamines, and carbon dioxide, leading to the curing reaction of epoxies with amine.61

Figure 5.

Figure 5

Regeneration of an amine by imine hydrolysis after exposure to atmospheric humidity.

Figure 6.

Figure 6

Single-component curing system which can generate both epoxy and amine upon heating.

3. Curing Process

3.1. Reaction Kinetics

As stated in the previous section (Figure 3), the curing process based on the step growth polymerization of epoxy and amine involves two steps. At the initial stage of the curing process, small chains with linear or branched structures are formed, and this is accompanied by a gradual increase in molecular weight. As the reaction proceeds, the branching of the chains becomes more pronounced, resulting in gelation.62 The curing time, or reaction conversion, at which gelation occurs, is called the gel point. At this point, the three-dimensional network expands over the entire system and thus the average molecular weight can be regarded to be infinity.63 When the curing temperature is sufficiently low, the transition from a liquid or gel to a glass, so-called vitrification, is thought to take place.64 Both the gelation and vitrification suppress further curing reactions due to a lowering of local mobility of unreacted functional groups and/or chain segments. Therefore, the curing proceeds via the chemically controlled reactions at the initial stage, followed by the diffusion-controlled reactions.65,66

The chemical reaction and concurrent increase in the average molecular weight at the initial stage of the curing process can be discussed on the basis of nuclear magnetic resonance (NMR) spectroscopy67,68 and gel permeation chromatography (GPC).69,70 However, these techniques cannot be applied to the late stage because of the insolubility of the curing product into organic solvents. Thus, differential scanning calorimetry (DSC) has been shown to be a valuable tool for studying the reaction kinetics of epoxy resins.71,72 It is known that the reactions involving the ring-opening of epoxy groups are exothermic. Given that the reaction rate is proportional to the heat flow, the degree of the curing reaction, namely the reaction conversion (α), can be examined as a function of time. There are essentially two types of experiments to determine the α value. One is an isothermal experiment where temperature is kept constant.73 The other is a “non-isothermal” or “dynamic” experiment, which involves ramping the temperature to a given value at a constant rate. For the former experiment, the α can be determined by the ratio of the heat recorded up to a certain time (Q(t)) to the total heat recorded over the entire reaction (Q).73 Hence, the conversion curve, which is a plot of the α value against curing time, can be analyzed with a kinetic model, as described later. Figure 7 shows typical examples of the conversion curves at various curing temperatures.

Figure 7.

Figure 7

Typical conversion curves for a mixture of DGEBA and DDM at various curing temperatures. Data are taken from ref (73) with a style modification. Copyright 2013 Elsevier.

Fourier-transform infrared (FT-IR) spectroscopy has also been widely used for the study of epoxy–amine reaction kinetics.74,75 In general, the reaction is monitored on the basis of the change in the intensity of the absorption bands in the near IR wavenumber range, typically 4000–7500 cm–1.76Figure 8 shows an example of FT-IR spectra for an epoxy–amine mixture at various stages of the curing process. At the initial stage, the spectrum provides an absorption band due to the combination of the stretching and bending vibration modes of epoxy groups at ∼4530 cm–1epoxy).75,76 Also, two bands are observed at ∼4940 and ∼6540 cm–1. The former is assignable to the combination of the stretching and bending vibrations of primary amino groups (νamino), while the latter includes the overtones of the stretching vibration for both primary and secondary amino groups (νps).77 As the reaction proceeds, the absorbance for both νepoxy and νamino bands decreases, meaning that the primary amino and epoxy groups reacted with each other. The reaction is also accompanied by a change in the νps band. The absorbance of the νps band initially decreases and subsequently shifts toward the lower wavenumber side because of the generation of secondary amino groups. Then the absorbance decreases as a result of the transformation of secondary into tertiary amino groups.77 Notably, a new broad band at ∼7100 cm–1 appears as the reaction proceeds. This band corresponds to the overtone of the stretching vibration for hydroxy groups, which was generated as a result of the ring opening of epoxy groups.77

Figure 8.

Figure 8

FT-IR spectra obtained for the reaction mixture of hydrogenated DGEBA (HDGEBA) and poly(3-aminopropylmethyl)siloxane at 343 K. Data are taken from ref (76) with a style modification.

Based on the absorbance change in the νepoxy band, the α value can be extracted. The plot of α against curing time can be analyzed on the basis of the kinetic models, which are also applied to the DSC data.78 The basic rate equation in the kinetic analysis can be expressed by dα/dt = k(T)f(α), where dα/dt is the rate of conversion, k(T) is a reaction rate constant, and f(α) is a function of α. The k(T) is dependent on the temperature and is generally assumed to be of the Arrhenius form with a pre-exponential factor, A, and an apparent activation energy, Ea. To date, various types of kinetic models with modified k(T) and f(α) have been proposed.79 Of these, the Kamal–Sourour model is the most widely used for the epoxy–amine reactions.80 In this model, the autocatalytic reaction of epoxy groups with primary amino ones is considered. This model represents the chemically controlled kinetics at the initial state of the curing reaction. To account for the shift from chemically controlled to diffusion-controlled reactions, the model has to be modified.81 For instance, Dušek and Havlíček introduced a dependency of the reaction rate on the Tg for the curing system.82 Another approach to take the diffusion effect into account is expanding the reaction kinetics model with a diffusion factor.83,84

An advantage of the kinetic study of the curing reaction with FT-IR over DSC is that the concentration of epoxy groups (CE), primary (CA1), secondary (CA2), and tertiary amino groups (CA3) can be determined. The analytical method proposed is based on the assumption that there are two step reactions as shown in Figure 3a without any side reactions and the reactivity ratio (R), which is defined as the reaction rate constant ratio between first and second steps (k2/k1), is independent of the reaction path.85,86 Thus, considering the mass balance of the functional groups, the concentration of each group can be estimated on the basis of the absorbance change for the νepoxy and νamino bands. Figure 9 shows an example of the time course of CE, CA1, CA2, and CA3 during the curing reaction process. As the reaction proceeds, the CE and CA1 values decrease while the CA2 and CA3 values increase. Then the CA2 value starts to decrease, while CA3 keeps increasing before finally reaching a plateau. The plateau region corresponds to the diffusion-controlled reaction due to the gelation and/or vitrification.87,88 According to a suggestion by PazAbuin et al., the R value can be estimated from the concentration ratio between CA1 and CA2 at the curing time, at which CA2 is maximized.89 The R values so far reported are less than 0.5, meaning that the reactivity of secondary amino groups is much lower than that of primary ones.89 This is explained by the reduced nucleophilicity and the increased steric hindrance of a secondary amino group relative to a primary one.85,86

Figure 9.

Figure 9

Time course for CE, CA1, CA2, and CA3 during the curing process at 323 K. Data are taken from ref (88). Copyright 2020 Royal Society of Chemistry.

3.2. Evolution of Network Structure

As mentioned before, FT-IR spectroscopy provides information on the change in the concentration of functional groups during the curing process. Based on such a change, the network formation can be discussed. It has been pointed out that the generation of secondary and tertiary amino groups in an epoxy–amine mixture depends on the curing temperature.87 At a lower temperature, secondary amines are initially generated and then converted to tertiary ones. At a higher temperature, on the other hand, the secondary and tertiary amines are concurrently generated at the initial stage of the curing. To explain such conversion behaviors, two different types of epoxy–amine network formation were proposed by Morgan and Sahagun.90Figure 10 shows a schematic illustration showing the two different network formations. At a lower curing temperature, linear chains initially grow until a low-density network expands over the system (skeleton network), and then unreacted groups cross-link within the skeleton network. The curing at a higher temperature provides the initial generation of the cross-linked domains, or microgels, followed by the interconnection with one another. The former and latter types would lead to less heterogeneous and heterogeneous networks, respectively.88

Figure 10.

Figure 10

Pictures for network formation for the curing at (a) lower and (b) higher temperatures. Reproduced with permission from ref (88). Copyright 2020 Royal Society of Chemistry.

Direct investigation of how the network structure in an epoxy resin develops during the curing process is quite difficult or even impossible because of the infusibility at higher temperatures and the insolubility in organic solvents. Generally, the resin has to be fractured and then characterized. Atomic force microscopy (AFM) and scanning electron microscopy (SEM) and have been commonly used to analyze the aggregation states for the fracture surface.9196 It has been pointed out that the fracture surface of an epoxy resin contained nodules with a size ranging from tens to hundreds of nanometers.9294 The nodular structure has been regarded as being at relatively high cross-linking regions with an interstitial phase of low cross-linking density. Also, it has been found that the characteristic length of the nodules decreases with increasing curing time.95 Such a morphological change was thought to reflect the increase in the cross-linking density as a result of the curing. Here, it should be noted that there is another interpretation for the nodular structure. That is, the nodular structure itself is not proof for the difference in the cross-linking density of the network because a similar structure is also found for the fracture surface of polymers without any cross-links.91,96 However, the latter interpretation seems to have been invalidated by a recently developed technique, nanoscale infrared analysis (AFM-IR). In this technique, the deflection of an AFM probe is used as a local sensor to detect photothermal expansion in response to infrared excitation, and the nanoscale lateral variations are detected in response to the illumination at different wavenumbers.97 AFM-IR measurement has revealed that the nodular structure corresponds to the chemical heterogeneity associated with the heterogeneous cross-linking structure.98,99

Recently, Izumi and Shibayama et al. proposed a nondestructive method using small-angle X-ray and neutron scattering (SAXS and SANS) to characterize the network at various state of the curing for a phenolic resin.100102 In this method, the resin is swollen in a good solvent to enhance a spatial difference in the cross-linking density.103 Through a series of works, it was found that tightly cross-linked domains initially appeared, and then the size of the domains increased as a result of incorporating other polymer chains into the domain.98 Such a network formation was also the case for epoxy resins.104

3.3. Change in Physical Properties

Gillham et al. proposed time–temperature-transformation (TTT) diagram, where gelation, vitrification, as well as degradation are represented.105Figure 11 shows a typical TTT phase diagram for thermosets including epoxy resins. It contains the gelation (solid, red), vitrification (dotted, blue), and degradation curves (dashed-dotted, black) at borders between different phases as functions of curing time and temperature. Liquid (before gelation), rubber (after gelation but above Tg), gelled glass (after gelation and below Tg), and degraded polymer are possible phases. The curing temperature is commonly chosen on the basis of the TTT diagram. No matter what temperature is chosen, the system will go through the gelation and/or the vitrification curve(s) after a certain time. Thus, great efforts have so far been made to gain a better understanding of gelation and vitrification during the curing process.106,107

Figure 11.

Figure 11

Typical time–temperature transformation (TTT) diagram for thermosetting polymers including an epoxy resin. Reproduced with permission from ref (105). Copyright 2003 John Wiley and Sons.

The reaction conversion at which the macroscopic gelation occurs, or the gel point, is usually determined by gel fraction measurement. Once a reaction system reaches the gel point, the insoluble component appears even in a good solvent, and then the fractional amount increases with increasing curing time.108 Viscosity measurement is also common for determining the gel point. The steady-state shear viscosity is measured as a function of curing time. Once the system undergoes gelation, the viscosity increases with increasing time. By extrapolating the viscosity to infinity, the gel point can be determined.109

Dynamic shear oscillatory measurements have become more common in characterization of the gel point upon the curing process.110112Figure 12 shows a typical example for the time course of shear storage and loss moduli (G′ and G″) during the curing process of an epoxy resin. The measurement was performed at an angular frequency of 10 rad s–1, which has been commonly used according to the suggestion by the American Society for Testing and Materials (ASTM).113 At the initial curing stage, G″ was larger than G′, indicating a liquid state. As curing proceeded, both G′ and G″ increased before reaching to a crossover point, at which G′ and G″ became identical with each other. After that, G″ reached a maximum while G′ kept increasing and then approached a plateau. The crossover point of G′ and G″ is often regarded as the gel point.114 However, since the crossover point depends on the measurement frequency, determination of the gel point using tan δ (= G″/G′) should be made with care. Actually, the gel point was determined to be the time at which the tan δ curves acquired at various frequencies intersect one another.115,116 Here, it is noteworthy that the decrease in G″ with increasing time is an indication of vitrification, where the segments are frozen in terms of mobility, leading to the lesser contribution of the energy dissipation.115,116 In fact, when the curing temperature is substantially higher than the Tg of the system, no decrease in G″ is observed during the curing process.117 The Tg value, which generally increases with increasing curing time, can be determined as a heat capacity change detected by temperature-modulated differential scanning calorimetry (TMDSC).118,119

Figure 12.

Figure 12

Time course of G′ and G′′ during the reaction between HDGEBA and 1,4-cyclohexanebis(methylamine) (CBMA) at 296 K. Data are taken from ref (104). Copyright 2019 American Chemical Society.

Characterization methods for the change in physical properties of the epoxy resins during the curing process are often limited to bulk measurements, which provide ensemble-averaged information over the entire region of the system. Recently, we applied a particle tracking experiment, which is one of the techniques for microrheology, to an epoxy–amine curing system.104 In this technique, probe particles are embedded in the medium to be measured. Since the thermal motion of the particles reflects the physical properties of the surrounding medium, tracking the movement provides insights into the local properties of the medium.120,121 Information on the spatial heterogeneity can be obtained by detecting the particles located at different positions in the medium.122,123 Also, by changing the particle size, the length scale of the observation can be altered.124128Figure 13 shows an illustration of the curing process drawn after a particle tracking study. Observation of individual particles at different locations revealed that, at the initial stage of the curing, the heterogeneous structure with a mesoscopic scale was generated and the characteristic length scale decreased with increasing curing time.104 Such a picture was recently confirmed by bimodal AFM, where both modulus and dissipated energy quantities describing the elastic and adhesive responses, respectively, were simultaneously obtained.129 Notably, the length scale of the heterogeneity decreased to a value of ∼10 nm at the fully cured stage.

Figure 13.

Figure 13

Schematic illustration showing the evolution of the heterogeneity during the curing process. Reproduced from ref (104). Copyright 2019 American Chemical Society.

3.4. Molecular Picture of Structure Formation

Atomistic and coarse-grained MD simulations have been applied to model the cross-linked structure of epoxy resins since around the year 2000. These developments have enabled us to gain access to the reaction kinetics, heterogeneity, and thermal and mechanical properties of the resins.130133 The basic procedure for creating the cross-linked structure is as follows. Sites where epoxy and amine molecules react are specified in advance, and then, when they approach within a certain distance during the MD simulation, the chemical bonds are reorganized and reacted. Yarovsky and Evans developed a modeling method for the cross-linked structure of a water-soluble phosphated epoxy resin.134 The distance to create new chemical bonds was set to within 0.6 nm, while at the same time, produced water or alcohol molecules were removed from the system. Wu and Xu proposed a method of reacting in order from the closest pair existing in the range of 0.4–1.0 nm and achieved a high reaction conversion of 90% or more.135 Komarov et al. proposed a method to improve computational efficiency by running the coarse-grained (CG) model once. In their method, the all-atom model is first converted to a CG model to reduce the computational load, and the curing reaction is simulated using CGMD. The obtained structure is then remapped to the all-atom model.136 Varshney et al. proposed a procedure to obtain a highly dense system with a reaction conversion of about 90% by differentiating the reactivity of primary and secondary amines.137 Bandyopadhyay et al. used a united atom model for a computationally efficient method.138

The methods described so far proceed in many steps, that is, in order from the closest one, and are effective for reducing the local stress in the system.139 On the other hand, there is another way in which pairs that may chemically react are specified in the initial structure and reactions are carried out in a single step.140 Although the computational efficiency of this approach has advantages, it is difficult to obtain a sufficiently relaxed structure. Lin and Khare proposed a single step method to obtain a relaxed structure by using a simulated annealing algorithm to minimize the sum of the bond length.141

The charge of each atom must be renewed when the chemical bond recombination occurs due to the cross-linking reaction. Generally, for the sake of simplicity, the electrical charge of atoms determined by the force field is assigned or calculated via a simple method like charge equilibration (QEq).142 Li and Strachan used the electronegativity equalization method (EEM), a fast empirical method that imparts a charge dependent on the surrounding environment. This made it possible to create a more precise cross-linked structure.143 Since the cross-linking reaction between epoxy and amine is exothermic, the temperature rises as the reaction progresses. Okabe et al. proposed an elegant algorithm that considers the activation energy of the reaction and the heat generated by the curing reaction, as shown in Figure 14.144 The Arrhenius-type reaction probability including the activation energy and the local temperature is defined, and whether or not the reaction occurs is determined by comparing with a random number. Once the reaction occurs, the kinetic energy corresponding to the heat of formation is applied to the site involved in the reaction, the temperature rises momentarily, and the subsequent reaction is accelerated. This method has made it possible to investigate differences in kinetics depending on the molecular structure.145

Figure 14.

Figure 14

DGEBA-based network structure and temperature distribution showing the rise in local temperature due to the exothermic reaction. Reproduced with permission from ref (144). Copyright 2013 Elsevier.

Based on the heat generated by the reaction using Okabe’s method, the heterogeneity of the network observed in the reaction process can be explained. We discussed the origin of the mesoscopic heterogeneity generated in an epoxy resin on the basis of the MD simulation.104 That is, once a reaction occurs, the temperature at the site is locally elevated, and a subsequent reaction is accelerated, resulting in the formation of the spatial heterogeneity.

We studied the effect of the molecular size of epoxies and amines on the reaction kinetics.146 In the combination of larger and smaller molecules of epoxy and amine, it was seen that the smaller the epoxy the faster the reaction. This is because when a primary amine reacts to become a secondary amine it is incorporated into the network, so that even if the initial diffusion is fast, the movement becomes slow. Also, we showed that the density increased due to the shrinkage as the reaction progressed; however, the shrinkage hardly occurred beyond the gel point.147 On the other hand, it was found that many free spaces in which water molecules can enter are formed beyond the gel point, as shown in Figure 15. It was ascertained that absorbed water molecules exist in the free space forming hydrogen bonds and diffuse in the free space one after another.

Figure 15.

Figure 15

Changes in density and free space occupancy during the curing process. Data are taken from ref (147). Copyright 2021 Royal Society of Chemistry.

4. Network and Physical Properties

4.1. Cross-Linking Density and Thermal Motions

One of the outstanding features of epoxy resins is the facile tunability of the network architecture by changing the cross-linking density. For example, the cross-linking density has been adjusted by changing the stoichiometric ratio of epoxy and amine in the initial reaction mixture.148,149 However, this method alters not only the cross-linking density but also the number density of residual functional groups and thereby the network defects including dangling chains, in which one end attaches to the network and the other is free.150 An alternative approach to changing the cross-linking density is the use of a mixture of mono- and difunctional amines as a curing agent. In this method, monofunctional amines behave as a chain extender, and thus, the cross-linking density can be systematically varied on the basis of the ratio between the mono- and difunctional amines.151,152 Furthermore, the cross-linking density can be simply tailored by varying the molecular weight of difunctional epoxies and/or amines. If a full conversion is ideally reached, this method should yield a network which contains no dangling chain.153 Thus, the distance between the two functional groups in epoxies or amines corresponds to the chain length between the cross-linking points.154,155

Cross-linking density affects the physical properties of epoxy resins. For example, as the cross-linking density increases, the Tg also increases.155,156 This is generally explained in terms of the dense glassy state, in which network chains tightly pack together, leading to a reduction in the free volume.157,158 If dangling chains exist in the network, the Tg tends to decrease because chains can be actively moved within a certain part of the free space.159 The effect of change in the architecture of the epoxy–amine network on the free volume content has been studied using positron annihilation lifetime spectroscopy (PALS).160,161 PALS is an analytical technique that can quantify nanoholes on the order of 0.2–2.0 nm in diameter, which are consistent with the interchain dimensions of most polymers.160 Pujari et al. reported that the free volume fraction increased with increasing cross-linking density although the size decreased.162 This is in good accordance with the result shown in Figure 15. Besides, it should be noted that the effect of the free volume content associated with the local chain motion on the moisture uptake of the epoxy–amine network has been elegantly described by Soles et al.163 and more recently by other researchers.164,165 Such an approach provides useful knowledge regarding hydrothermal aging, which is one of the degradation phenomena due to the presence of moisture at an elevated temperature.

Thermal molecular motion in epoxy resins was also dependent on the cross-linking density, as evidenced by dielectric relaxation spectroscopy (DRS)166,167 and dynamic mechanical analysis (DMA).168,169Figure 16 shows temperature dependence of tensile storage (E′) and loss moduli (E″) for an epoxy resin composed of DGEBA and 1,2-diaminoethane (DAE) by DMA. Two E″ peaks were observed at around 420 and 250 K. The first peak at 420 K was accompanied by a clear decrease in E′. The peak was referred to as α-relaxation which was generally assigned to the segmental motion in the network.170 The latter peak at 250 K was named β-relaxation, which mainly corresponded to the motion of diphenylpropane groups and glyceryl units, −CH2–CH(−OH)–CH2–O–.171,172 Since the shape and intensity of the β-relaxation peak was affected by the cross-linking density, it was noted that the relaxation also contained a local motion coupled with the motion of the main chain in the network. The amplitude of the β-relaxation has been discussed in conjunction with the stiffness of the glassy epoxy resins.173

Figure 16.

Figure 16

Temperature dependence of E′ and E″ for an epoxy resin composed of DGEBA and DAE at various frequencies ranging from 3.5 to 110 Hz. These are unpublished data.

4.2. Dynamic Heterogeneity

Since epoxy resins are often in a glassy state at room temperature, the network structure is frozen in terms of its mobility. Thus, understanding the glass transition dynamics associated with cross-linking density is needed to regulate the mechanical properties such as yielding and fracture behaviors.174,175 The glassy dynamics can be characterized by the fragility index (m), which is defined as the apparent activation energy for the α-relaxation process near the glass transition.176Figure 17 shows a semilogarithmic plot of relaxation time (τ) for the α-relaxation process against the inverse temperature (T–1) for epoxy resins with different cross-linking densities. The abscissa is normalized by each Tg value, provided as an Angell plot.176 The m value corresponds to the steepness of the plot at Tg. Combining with an incoherent elastic neutron scattering technique, the magnitude of m is claimed to be the extent of the dynamic heterogeneity, which is a transient spatial fluctuation in the cooperative segmental dynamics near the glass transition.177 Based on the m value, the extent of the dynamic heterogeneity can be discussed.177 The glass transition dynamics can be also characterized by characteristic length scale of the cooperative rearranging region (CRR) (ξCRR).178180 The CRR is defined as the smallest subsystem in which one segment is necessarily involved in coordinated motions of other segments at a temperature near Tg.178180 The ξCRR value can be experimentally estimated by various techniques such as low-frequency Raman spectroscopy,181,182 Brillouin light scattering,183,184 four-dimensional NMR spectroscopy,185,186 and TMDSC.187,188 Some studies have reported that with increasing cross-linking density in the epoxy resin, the m and ξCRR values increased, while others decreased.162,189

Figure 17.

Figure 17

Angell plots for epoxy resins with various cross-linking densities. Symbols and solid lines denote experimental data and best-fit curves using the Vogel–Fulcher–Tammann equation, respectively. Data are taken from ref (155). Copyright 2021 American Chemical Society.

Recently, we studied the glass transition dynamics in epoxy resins in which the cross-linking density was systematically altered by chain length of n-alkyl diamines used as curing agent.155 As the cross-linking density increased, the Tg increased, accompanied by a reduction in ξCRR and an increase in the dynamic heterogeneity. Notably, the analysis of the self-part of the space-time correlation function by the MD simulation revealed that the thermal motion of nitrogen atoms, which acted as a cross-linking point, was suppressed in comparison with that of other constituent atoms. The motional difference between nitrogen and other atoms, which corresponded to the dynamic heterogeneity, became more significant as cross-linking density increased. In addition, by applying a time–temperature superposition (TTS) principle to the dynamic viscoelastic functions, we found that as the cross-linking density increased, the thermal expansion of the free volume was suppressed and the entropic elasticity became less remarkable in the temperature region above the Tg.190 With the aid of MD simulation, the entropy change was confirmed by isobaric molar heat capacity calculated from the ensemble variation of enthalpy. Here it should be noted that the TTS principle is one of the promising methods to predict long-term properties from short-term tests.191193 Actually, the creep measurements at various temperatures, which require a time of 10 h, enable us to access a time scale of up to 106 h.194 Since the long-term properties are closely related to their durability, the their prediction is of importance from a practical application perspective.194196

4.3. Fracture Toughness

Epoxy resins are generally brittle. Since this feature is one of the greatest drawbacks for usage as a structural material and adhesive, it is desired to overcoming this problem. So far, many researchers have studied the fracture toughness for polymer glasses without any chemical cross-links.197199 Through a series of works, it is known that the toughness of polymer glasses depends on the molecular weight, or the apparent entanglement density, of chains.197,199 This is explained in terms of the slippage of chains with others, induced by the deformation and/or craze formation.200,201 Thus, once the chains are chemically cross-linked with one another, the chain slippage is expected to be suppressed, resulting in an improvement of fracture toughness. This strategy should work for epoxy resins but do not necessarily. In fact, it has been reported that as the cross-linking density increases, the toughness increases and then begins to decrease.202,203 Hence, further study to obtain a better understanding of the mechanism of toughness manifestation should be conducted.

We recently reported on how curing temperature affected the fracture behavior of the resultant epoxy resins.88 Epoxy resins were prepared by precuring at four different temperatures and then postcured to eventually reach the same cross-linking density. However, as the precuring temperature increased, the m value decreased. That is, the dynamic heterogeneity became more apparent. Figure 18 shows photographs of the four epoxy resins immersed in tetrahydrofuran (THF), which is a good solvent for them. It is known that once a glassy material is exposed to a good solvent, or vapor, macroscopic fractures occur due to an enhancement of residual stress.204,205 Since the epoxy resins are often utilized in contact with a solvent, the resistance to the solvent-induced stress is required in various applications.206 Interestingly, it was found that the immersion time required to reach fracture became shorter as the extent of the dynamic heterogeneity increased as a result of the stress concentration.88

Figure 18.

Figure 18

Photographs of four epoxy resins with different m values of (a) 130, (b) 110, (c) 99, and (d) 85 in THF after various immersion times. Reproduced with permission from ref (88). Copyright 2020 Royal Society of Chemistry.

In practical applications, one of the common approaches to toughening epoxy resins is to disperse soft particles as a filler into the matrix.207210 For example, incorporating rubbery polymer particles such as carboxy-terminated butadiene acrylonitrile can remarkably enhance the fracture toughness.207209 The toughening mechanism proposed is the cavitation of the rubbery particles themselves followed by void growth, which leads to energy dissipation.208,209 Several types of nanomaterials such as carbon nanotubes, graphene, clay, and silica have also been tested.211214 Of these, silica particles have attracted attention because of their high specific surface area, high surface energy, low toxicity, and ease of manufacturability. Furthermore, the compatibility of silica particles into an epoxy matrix can be tuned by surface modification with silane coupling reagents.215218 Using silica particles as a filler, improvements in the toughness have been achieved.219222 Since silica particles can be regarded as a hard material, the toughening mechanism should differ from that based on rubbery polymer particles. After many works dealing with the effect of the size and the volume fraction of silica particles, the process related to the toughening mechanism is considered to be mainly categorized into two. One is an in-plane process such as crack tip pinning or bowing223 and crack path deflection,224,225 while the other is an out-plane process such as debonding and plastic void growth.226,227

Recently, Yamada, Kobayashi and co-workers reported in situ transmission electron microscopic (TEM) observation of the deformation and fracture processes for an epoxy resin film containing silica nanoparticles under the tensile process.228 Dispersed silica nanoparticles in the composite arrested the progress of the crack tip and prevented crack propagation. Concomitantly, the generation and growth of nanovoids at the epoxy matrix/nanoparticle interfaces were clearly observed, particularly in the region near the crack tip. Also, using a digital image correlation method, the presence of particles in the growing crack suppressed the generation of strain, potentially contributing to hindering crack growth, as shown in Figure 19.

Figure 19.

Figure 19

True strain distribution in the vicinity of the crack tip for epoxy resin composed of HDGEBA and CBMA. The tensile displacement is applied continuously in the y direction. The crack is expected to grow in the x direction. Reproduced with permission from ref (228). Copyright 2022 Royal Society of Chemistry.

4.4. Physical Properties by Simulations

Thermal and mechanical properties have also been studied using the network structure modeled by simulations. Evaluating the change in the specific volume with respect to the temperature, the coefficient of thermal expansion (CTE) and the Tg can be obtained. The Tg value is defined as the temperature at which the slope of the specific volume as a function of temperature changes upon the cooling process. However, the Tg value is generally higher in an MD simulation than in an experiment because the cooling rate used in the simulation is several orders of magnitude higher than in the experiment. The difference can be corrected using the Williams–Landel–Ferry (WLF) equation. Soni et al. compared the simulated CTE and Tg values of various epoxy resins with experimental ones.229 They also discussed the chain length effect of cross-linkers, shown in Figure 20, and pointed out that an increase in the chain length of the cross-linker led to a larger difference between the predicted and experimental values of Tg.

Figure 20.

Figure 20

Example of Tg determination from the temperature dependence of specific volume for the four epoxy resins composed of DGEBA and poly(oxypropylene) diamines with the different chain lengths (n). Data are taken from ref (229). Copyright 2012 Elsevier.

Mijović and Zhang examined the local relaxation dynamics of a cured epoxy resin via DRS and discussed the molecular interaction arisen from hydroxy and ether groups based on an MD simulation.230 Shenogina et al. estimated the elastic constants for highly cured epoxy resins and claimed that the values so obtained were higher than those by experimentation.231 They tried to explain the discrepancy on the basis of both finite-size effect and limitation of the static deformation approach to account for the dynamic effects. Okabe et al. evaluated Young’s modulus for cured products of several combinations of epoxy and amine and showed that the experimental values could be successfully reproduced after adjusting the van der Waals radius to fit the density in the experiment.232 They also claimed that electrostatic interaction plays an essential role in the mechanical properties. Odegard et al. proposed a simulation procedure using a reactive force field, which can handle the recombination of chemical bonds, and discussed the mechanical properties for epoxy systems comparing the results with experimental ones.233 In general, the strain rate in an MD simulation is several orders of magnitude higher than that of an experiment due to the limitation of computing time. They stated that the calculated values matched the straight line extrapolated by the experimental values, as shown in Figure 21.

Figure 21.

Figure 21

Relationship between strain rate and Young’s modulus obtained in experiments and simulations. EPON 862: diglycidyl ether of bisphenol F (DGEBF); EPON828: DGEBA; DETDA: diethyltoluenediamine; PEA: polyetheramine. Data are taken from ref (233). Copyright 2021 American Chemical Society.

4.5. Interfacial Properties

Epoxy resins have been widely used as an adhesive in various industrial applications. Two adherend surfaces adhered by the epoxy resin can be debonded either by “cohesive” or “adhesive” failure. Cohesive failure occurs in the bulk region of an epoxy resin or in the bulk of an adherend material. Conversely, adhesive failure occurs at the interface between an epoxy resin and the adherend. Actually, the failure occurs due to a combination of both cohesive and adhesive modes.234 In general, the cohesive mode is preferable to achieve a relatively high adhesive strength. If the strength of the interaction between the epoxy resin and the adherent surface is not high enough, adhesive failure takes place, leading to facile delamination.234 One way to avoid this is to evaluate the strength of the chemical interaction at the interface. Experimentally, the surface free energy may be measured to estimate the adhesive force at the interface but it is not easy to measure the epoxy interface during the reaction. As an alternative method, MD and density functional theory (DFT) techniques have been used to estimate the relationship between the surface chemical states of the adherend material and the interfacial interaction. Bahlakeh and Ramezanzadeh studied the adhesion mechanism for untreated/treated steel substrates under dry and wet conditions and showed the role of electrostatic and van der Waals interactions along with the order of the surface states.235

It is known that the physical properties of epoxy resins near the solid interface differ greatly from those in the internal bulk region.236 Since such an interfacial region especially contacted with a metal substrate is on (sub)micrometer scale in its thickness, it is often referred to as an “interphase”.237,238 The interface and/or interphase are/is considered to be an important factor for the material performance. Such can be seen in a flip-chip microelectronic packaging, which uses an epoxy resin as an electronical insulating adhesive. In this case, the interphase between the epoxy resin and the metal layer plays an important role in the long term durability.239 Thus, the structure and physical properties of the interphase have been extensively studied.240,241 For example, Carriere et al. examined the Tg value for an epoxy resin as a function of film thickness, suggesting an elevation of the Tg in the interphase contacted with a silicon substrate with a native oxide layer.242 Chung et al. reported that using scanning force microscopy-based force modulation microcopy (SFM-FMM), the interphase consisted of a high-stiffness region near the interface with copper which was adjacent to a relatively low-stiffness region along the direction normal to the interface.243

A possible explanation for the interphase formation is that the chemical composition of an epoxy resin is not uniform along the direction normal to the solid interface.244 That is, the epoxy or amine component is preferentially segregated to the interface. Other explanation includes the change in the reaction kinetics for the epoxy-amine mixture near the solid substrate due to the imbalance of the reactants,245 the catalytic effect of metallic oxide substrate,246 the suppressed diffusion of the reactants,247 and so forth.248 Using FT-IR with an attenuated total reflection (ATR) mode, we also found the initial reaction kinetics for the epoxy and amine compounds are slower near the solid interface than in the bulk region.249

So far, many researchers have discussed a possible formation mechanism for the interphase near the metal substrate.250,251 If the interphase formation is a result of the preferential segregation of the amine (or epoxy) component due only to the difference in surface energy between two components, the thickness would on the order of the size comparable to that of the epoxy or amine residues. However, the thickness of the interphase has been found to be much greater than expected, although it depends on the kind of metal.251 Previously, it has been pointed out that metal ions diffused out from the metal into the mixture of epoxy and amine and then coordinated with amine(s), resulting in the complex formation.250 In fact, energy dispersive X-ray spectrometry (EDX) and electron energy loss spectroscopy (EELS) for a cross-section of an epoxy resin contacted with aluminum and copper substrates revealed that metal species deeply migrated into the epoxy resin.252,253 Recently, we confirmed that an amine component was preferentially segregated near the copper interface by a nondestructive method using angular-dependent X-ray photoelectron spectroscopy (ADXPS) in which an incident X-ray was guided from the copper surface, as shown in Figure 22.254

Figure 22.

Figure 22

Schematic illustration of a nondestructive method to examine the depth profile of the chemical composition along the direction perpendicular to the copper interface. Illustration is reproduced with permission from ref (254). Copyright 2018 Springer Nature.

There are several reports on the aggregation states and the reactivity of molecules in an epoxy resin near the interface based on MD simulations. We found that in a mixture of epoxy and amine at the interface with copper amine with a smaller molecular size was selectively concentrated due to the packing entropy.255 We further demonstrated that epoxy segregated at the interface when smaller epoxy molecules were used.256 When larger and smaller epoxy and amine were mixed, each smaller molecule selectively segregated at the interface as shown in Figure 23. Consequently, the progress of the reaction was suppressed at the interface by the depletion of the reaction partner as well as the decrease in mobility.

Figure 23.

Figure 23

Representative snapshot near the interface. Molecules of DGEBA (Ep-L), 2,2-di(4-(3-aminopropyl)phenyl)propane (Am-L), ethylene glycol diglycidyl ether (Ep-S), and 1,8-diaminooctane (Am-S) are colored red, blue, green, and yellow, respectively. Here, Ep and Am denote epoxy and amine, respectively, and the letters L and S mean larger and smaller. Copper atoms are colored brown. Models are reproduced with permission from ref (256). Copyright 2021 Royal Society of Chemistry.

5. Applications

Epoxy resins have been widely used in various applications for industrial products and home appliances to take advantage of their excellent properties. In this section, the recent trends with pioneering applications are briefly summarized.

5.1. Structural Materials and Adhesives

As mentioned in section 4.3, the toughening of epoxy resins has been strongly desired for their application as structural materials and adhesives. This has been attempted by various approaches, which are classified into (i) elastomer modification, (ii) particulate modification, (iii) thermoplastic modification, and (iv) miscellaneous methods.210,257 Incorporating the polyrotaxane (PR) structure into epoxy resins is also a promising candidate method.258,259 Hanafusa et al. studied the molecular dynamics of PR in which poly(ε-caprolactone) (PCL) grafted onto α-cyclodextrin (CD) crossed the poly(ethylene glycol) (PEG) axis, uniformly dispersed in a cross-linked epoxy resin.260 As the temperature rose, PEG in PR underwent a glass-to-rubber transition that fluctuated within the glassy PCL-grafted CD confined in the matrix, causing the viscoelastic relaxation. This improved the deformability and toughness of the epoxy resin containing PR under uniaxial stretching.

The degradation of epoxy resins after use is a critical issue for material recycling, reducing the environmental load as well as the development of dismantlable adhesives. Tano and Sato reported an epoxy resin, composed of DGEBA with a photodimer of 9-anthracene carboxylic (9-AC) acid,261 which was successfully decross-linked to be solubilized in organic solvents upon heating. The solubilized products reformed a network structure by photodimerization of 9-AC units. Another example is the one with disulfide linkages attached to the main epoxy chains.262 The facile degradation was possible via disulfide exchange reactions thanks to the ability of disulfide bridges to be fragmented and detached from the main epoxy chains.

In the past two decades, many efforts have been made to produce recyclable, reprocessable, and healable epoxies by introducing reversible bonds into the network structure including reversible covalent bonds.263,264 From such a background, a new class of polymers, known as vitrimers, was introduced by Leibler et al.265 Vitrimers are materials containing a cross-linked network with dynamic covalent bonds, where cross-linking density remains unchanged when an exchangeable reaction happens. At a service temperature, vitrimers behave like a traditional thermoset. Once they are heated up to a temperature above the topology freezing transition temperature, an exchangeable reaction occurs rapidly, resulting in a fluid behavior.265,266 Such a feature makes it possible for the vitrimers to be reprocessed, reshaped, remolded, and recycled.267

In most cases, pure epoxy vitrimers seem to not satisfy increasing various industrial demands. Recently, to overcome this, the incorporation of fillers into the vitrimers was proposed. This approach often can provide vitrimer composites with various functions including mechanical reinforcement, stress relaxation, welding, self-healing (repairing) and shape memory.268 For example, the modulus, yield stress, fracture strain of the epoxy vitrimers could be enhanced by embedding graphene in it. In addition, the shape of the vitrimer could be controlled by near-infrared light due to the photothermal effect of graphene.269 Also, the photothermal effect of carbon nanotubes (CNT) dispersed in an epoxy vitrimer made it possible to control the welding behavior.270 Here, it should be noted that for the most vitrimer composites, there exists an inevitable drawback that the stress relaxation is suppressed at the filler interface due to hindered exchangeable reactions because of the less chain mobility.268 This issue would become more important for the future practical applications.

5.2. Thermal Conductive Materials

Epoxy resins have also been used as electrical insulating materials in electronic components. In recent years, power electronics products have improved greatly in performance and compactness, though the heat generated from the inside has increased along with the improvements. For this reason, how efficiently heat is dissipated to the outside of a device is an important issue that determines the performance and the life of the device, and heat dissipation technology is an extremely important factor. Silica (SiO2) and alumina (Al2O3) particles have been blended as a heat conductive filler to improve the thermal conductivity of an epoxy resin.271,272 Although the higher filling rate of particles increases the thermal conductivity, the material properties can deteriorate markedly due to the generation of voids. In order to further improve the thermal conductivity, high thermal conductivity fillers such as boron nitride (BN), aluminum nitride (AlN), and silicon carbide (SiC) whisker and their composite systems have been intensively studied.273275 However, when the filler content becomes high, it is inevitable that the material properties including processability are negatively affected. Thus, improvements in the thermal conductivity of the epoxy resin itself are keenly sought.

A strategy for increasing the thermal conductivity of an epoxy resin itself is to use the phonon transport mechanism, namely, to realize a highly oriented structure such as a crystalline or liquid crystal polymer. Considering the minimum thermal conductivity model (MTCM)276 in which the crystalline phase has a higher thermal conductivity, polymers with crystalline domains can be a candidate for a higher thermal conductive material.277 Similarly, much attention has been focused on emerging highly oriented structures and thereby increasing thermal conductivity by introducing mesogenic groups into the epoxy resins.278 Lv et al. succeeded in achieving thermal conductivity 2.5 times higher than that of the conventional epoxy resins by using a diamine with an anthraquinone backbone as a curing agent.279 They employed four isomers with different positions of the amino groups, and demonstrated that all products formed semicrystalline domains and their thermal conductivity had a positive correlation with the mass density. Mo et al. achieved high thermal conductivity by forming an oriented nanostructure and enhancing the chain rigidity when using epoxy mixed with 4,4′-dihydroxydiphenyl (DHDP) in DGEBF.280 Song et al. also reported high thermal conductivity of an epoxy resin obtained with mesogenic groups.281 The conductivity was associated with the presence of an agglomerated spherulite structure of highly ordered lamellae. In fact, the thermal conductivity was linearly proportional to the spherulite size, which was determined by the competition between the curing reaction and the spherulite formation. Instead of the self-organization of the highly ordered structure, methods using an external field have also been demonstrated. Harada et al. obtained high thermal conductivity by curing diglycidyl ether terephthalylidene-bis(4-amino-3-methylphenol) (DGETAM) and 4,4′-diaminodiphenylethane (DDE) under a magnetic field, which induces the orientation of mesogenic groups.282 Although solid strategies are being established for improving thermal conductivity of the epoxy resins, as mentioned above, the next challenge is to achieve a packaging structure for power modules that reduces stress and suppresses void formation.

5.3. Electrically Conductive Materials

Electrically conductive adhesives (ECAs) are promising materials in electronic applications thanks to their lower temperature processability, environmental friendliness (lead-free), and flexibility. Among them, epoxy resins in which silver (Ag) and gold (Au) fillers and carbon-based fillers such as CNT and carbon black (CB) are dispersed are widely used.211,214,283,284 The main mechanism of the electric conductivity for ECAs is the contact between fillers. Thus, it is necessary to disperse fillers at a high concentration above the percolation threshold.285 ECAs are classified into two types: isotropic conductive adhesives (ICAs) and anisotropic conductive adhesives (ACAs).286,287 In ICAs, the electric current flows in all directions, while in ACAs it flows in only one direction. This depends on whether the morphology formed by the filler is isotropic or anisotropic. ICAs are used as an alternative to solder in heat-sensitive electronic components. Among metal fillers, Ag is often used because of its high conductivity and corrosion resistance.288 Wu et al. reported on epoxy-based ICA filled with Ag nanowires.289 They claimed that their ICA exhibited lower bulk resistivity and higher shear strength with a lower filler content than conventional ICAs filled with micrometer- and nanometer-sized Ag particles. On the other hand, ACAs are widely used in flat panel display modules and flip-chip on glass, etc. In this case, the filler forms a percolated structure along only one direction. Jiang et al. reported high-performance electronic interconnection with CNTs.290 Also, Massoumi et al. proposed the fabrication procedure for electrically conductive nanocomposite adhesives based on an epoxy resin containing surface-modified multiwalled carbon nanotubes (MWCNTs).291 The main drawback of these electrically conductive materials can be the high filler loading amount to achieve the desired conductivity, resulting in reduced mechanical properties. To overcome this, the effect of conductive fillers on the curing process is necessary to be better understood.

5.4. Biobased Materials

The global trends toward the principles of sustainable development urge industry to produce renewable and recyclable products synthesized from biobased materials. Figure 24 shows examples of biobased epoxy compounds synthesized from biobased resources such as rosin, sugar, itaconic acid, cardanol, lignin, tannin, and vegetable oil.292,293 To achieve fully biobased epoxy resins, curing agents were also derived from renewable materials such as modified plant oil, biobased acid and anhydride, amidoamine from rosin and tung oil, lignin, biobased phenol, and rosin acid.294,295 Although various types of biobased epoxy compounds and curing agents have been hitherto proposed, most of them have not reached commercial products. The major reason for this is the additional cost due to the isolation and synthesis of the natual monomers as with other excellent materials.295 Thus, the biobased epoxy resins must provide added value to justify their cost. Recently, to this end, various attempts to improve their performance have been ongoing. A typical example for such trials is the preparation of biobased epoxy composites reinforced with nanocellulose (NC), which is a class of shape-anisotropic materials and is generally extracted from natural resources (e.g., wood pulp, cotton, etc.).126 Utilization of NC as a filler provided fully biobased epoxy composites with the enhanced mechanical and thermal properties.296,297

Figure 24.

Figure 24

Examples of biobased epoxy compounds synthesized from renewable resources.

6. Summary

Epoxy resins will continue to be in the forefront of many thermoset applications due to their versatile properties. To expand the future applications of epoxy resins, toughness and flexibility, rapid curing potential, self-healing ability, reprocessability, recyclability, high-temperature stability, and conductivity should be improved. A precise prediction of long-term durability is also of pivotal importance from a practical application perspective. This Perspective summarizes works associated with some of the oldest, newest, and most difficult problems. We believe this contributes to a better understanding of how the network formation accompanying the curing reaction affects the physical properties of the resultant epoxy resin and shows that the use of robust physical chemistry techniques will lead to radical advances in thermosetting polymers, with many practical applications.

Acknowledgments

The authors thank Dr. Mika Aoki for her fruitful discussion and are grateful for support from the JST-Mirai Program (JPMJMI18A2) (K.T.) and the JSPS KAKENHI for Scientific Research (B) (no. JP20H02790, K.T.; no. JP19H02780, A.S.).

The authors declare no competing financial interest.

References

  1. Baekeland L. H. The Synthesis, Constitution, and Uses of Bakelite. J. Ind. Eng. Chem. 1909, 1 (3), 149–161. 10.1021/ie50003a004. [DOI] [Google Scholar]
  2. Pascault J.-P.; Sautereau H.; Verdu J.; Williams R. J. J. In Thermosetting Polymers, 1st ed.; CRC Press, 2002; pp1–5. [Google Scholar]
  3. Kandelbauer A.Processing. In Handbook of Thermoset Plastics, 3rd ed.; Dodiuk H., Goodman S. H., Eds.; William Andrew, 2014; pp739–753. [Google Scholar]
  4. Prolongo S. G.; del Rosario G.; Urena A. Comparative Study on the Adhesive Properties of Different Epoxy Resins. Int. J. Adhes. Adhes. 2006, 26 (3), 125–132. 10.1016/j.ijadhadh.2005.02.004. [DOI] [Google Scholar]
  5. Liu Y. Trends of Power Semiconductor Wafer Level Packaging. Microelectron. Reliab. 2010, 50 (4), 514–521. 10.1016/j.microrel.2009.09.002. [DOI] [Google Scholar]
  6. Yousefi N.; Sun X. Y.; Lin X. Y.; Shen X.; Jia J. J.; Zhang B.; Tang B. Z.; Chan M. S.; Kim J. K. Highly Aligned Graphene/Polymer Nanocomposites with Excellent Dielectric Properties for High-performance Electromagnetic Interference Shielding. Adv. Mater. 2014, 26 (31), 5480–5487. 10.1002/adma.201305293. [DOI] [PubMed] [Google Scholar]
  7. Forsdyke K. l.; Starr T. F.. Thermoset Resins Market Report; Smithers Rapra Technology, 2002; pp53–57. [Google Scholar]
  8. Ratna D.Handbook of Thermoset Resins; Smithers Rapra Technology, 2009; pp155–181. [Google Scholar]
  9. Vidil T.; Tournilhac F.; Musso S.; Robisson A.; Leibler L. Control of Reactions and Network Structures of Epoxy Thermosets. Prog. Polym. Sci. 2016, 62, 126–179. 10.1016/j.progpolymsci.2016.06.003. [DOI] [Google Scholar]
  10. Xiang Q.; Xiao F. P. Applications of Epoxy Materials in Pavement Engineering. Constr. Build. Mater. 2020, 235, 117529. 10.1016/j.conbuildmat.2019.117529. [DOI] [Google Scholar]
  11. Dearborn E. C.; Fuoss R. M.; MacKenzie A. K.; Shepherd R. G. Epoxy Resins from Bis-, Tris-, and Tetrakisglycidyl Ethers. Ind. Eng. Chem. 1953, 45 (12), 2715–2721. 10.1021/ie50528a047. [DOI] [Google Scholar]
  12. Pearce P. J.; Davidson R. G.; Morris C. E. M. Hydrolytic Stability of Some Uncured Epoxy Resins. J. Appl. Polym. Sci. 1981, 26 (7), 2363–2372. 10.1002/app.1981.070260722. [DOI] [Google Scholar]
  13. Tanaka K.; Yamaguchi M. Dynamic-mechanical Properties of Multifunctional Epoxy Resin Cured with Diamine and Filled with Pitch-based Carbon Short Fibers Treated with Coupling Agents. Adv. Compos. Mater. 1995, 5 (1), 45–62. 10.1163/156855195X00113. [DOI] [Google Scholar]
  14. Mustata F.; Bicu I. Multifunctional Epoxy Resins: Synthesis and Characterization. J. Appl. Polym. Sci. 2000, 77 (11), 2430–2436. . [DOI] [Google Scholar]
  15. Fache M.; Monteremal C.; Boutevin B.; Caillol S. Amine Hardeners and Epoxy Cross-linker from Aromatic Renewable Resources. Eur. Polym. J. 2015, 73, 344–362. 10.1016/j.eurpolymj.2015.10.032. [DOI] [Google Scholar]
  16. Chiao L. Mechanistic Reaction-kinetics of 4,4’-Diaminodiphenyl Sulfone Cured Tetraglycidyl-4,4’-diaminodiphenylmethane Epoxy-resins. Macromolecules 1990, 23 (5), 1286–1290. 10.1021/ma00207a010. [DOI] [Google Scholar]
  17. Hsieh T. H.; Su A. C. Cure Kinetics of an Epoxy-novolac Molding Compound. J. Appl. Polym. Sci. 1990, 41 (5–6), 1271–1280. 10.1002/app.1990.070410535. [DOI] [Google Scholar]
  18. Rwei S. P.; Liu A. Y.; Liou G. S.; Cheng K. C.; Guo W. Curing and Pyrolysis of Cresol Novolac Epoxy Resins Containing 2- (6-Oxido-6H-dibenz (c,e) (1,2)oxaphosphorin-6-yl)-1,4-naphthalenediol. Polym. Eng. Sci. 2004, 44 (2), 376–387. 10.1002/pen.20034. [DOI] [Google Scholar]
  19. Wu S.; Soucek M. D. Crosslinking of Acrylic Latex Coatings with Cycloaliphatic Diepoxide. Polymer 2000, 41 (6), 2017–2028. 10.1016/S0032-3861(99)00370-5. [DOI] [Google Scholar]
  20. Lu M. P.; Liu Y. C.; Du X. X.; Zhang S. H.; Chen G. K.; Zhang Q.; Yao S.; Liang L. Y.; Lu M. G. Cure Kinetics and Properties of High Performance Cycloaliphatic Epoxy Resins Cured with Anhydride. Ind. Eng. Chem. Res. 2019, 58 (16), 6907–6918. 10.1021/acs.iecr.8b06442. [DOI] [Google Scholar]
  21. Morita Y. Cationic Polymerization of Hydrogenated Bisphenol-A Glycidyl Ether with Cycloaliphatic Epoxy Resin and Its Thermal Discoloration. J. Appl. Polym. Sci. 2005, 97 (3), 1395–1400. 10.1002/app.21896. [DOI] [Google Scholar]
  22. Spurr A. R. A Low-viscosity Epoxy Resin Embedding Medium for Electron Microscopy. J. Ultrastruct. Res. 1969, 26 (1–2), 31–43. 10.1016/S0022-5320(69)90033-1. [DOI] [PubMed] [Google Scholar]
  23. Liu W. S.; Wang Z. G.; Xiong L.; Zhao L. N. Phosphorus-containing Liquid Cycloaliphatic Epoxy Resins for Reworkable Environment-friendly Electronic Packaging Materials. Polymer 2010, 51 (21), 4776–4783. 10.1016/j.polymer.2010.08.039. [DOI] [Google Scholar]
  24. Katritzky A. R.; Ramsden. C A.; Joule J. A.; Zhdankin V. V.. Structure of Small and Large Rings. In Handbook of Heterocyclic Chemistry, 3rd ed; Elsevier, 2010; pp 210–237. [Google Scholar]
  25. Ellis B.Introduction to the Chemistry, Synthesis, Manufacture and Characterization of Epoxy Resins. In Chemistry and Technology of Epoxy Resins; Springer, 1993; pp 1–36. [Google Scholar]
  26. Jin F. L.; Li X.; Park S. J. Synthesis and Application of Epoxy Resins: A Review. J. Ind. Eng. Chem. 2015, 29, 1–11. 10.1016/j.jiec.2015.03.026. [DOI] [Google Scholar]
  27. Bouillon N.; Pascault J. P.; Tighzert L. Epoxy Prepolymers Cured with Boron Trifluoride-amine Complexes, 2. Polymerization Mechanisms. Makromol. Chem. 1990, 191 (6), 1417–1433. 10.1002/macp.1990.021910620. [DOI] [Google Scholar]
  28. Ghaemy M. The Polymerization Mechanism and Kinetics of DGEBA with BF3-EDA. Eur. Polym. J. 1998, 34 (8), 1151–1156. 10.1016/S0014-3057(97)00247-4. [DOI] [Google Scholar]
  29. Crivello J. V.; Lam J. H. W. Diaryliodonium Salts - New Class of Photo-initiators for Cationic Polymerization. Macromolecules 1977, 10 (6), 1307–1315. 10.1021/ma60060a028. [DOI] [Google Scholar]
  30. Crivello J. V.; Lam J. H. W. Photoinitiated Cationic Polymerization with Triarylsulfonium Salts. J. Polym. Sci. A Polym. Chem. 1979, 17 (4), 977–999. 10.1002/pol.1979.170170405. [DOI] [Google Scholar]
  31. Takuma K.; Takata T.; Endo T. Cationic Polymerization of Epoxide with Benzyl Phosphonium Salts as the Latent Thermal Initiator. Macromolecules 1993, 26 (4), 862–863. 10.1021/ma00056a047. [DOI] [Google Scholar]
  32. Yagci Y.; Reetz I. Externally Stimulated Initiator Systems for Cationic Polymerization. Prog. Polym. Sci. 1998, 23 (8), 1485–1538. 10.1016/S0079-6700(98)00010-0. [DOI] [Google Scholar]
  33. Heise M. S.; Martin G. C. Curing Mechanism and Thermal-properties of Epoxy Imidazole Systems. Macromolecules 1989, 22 (1), 99–104. 10.1021/ma00191a020. [DOI] [Google Scholar]
  34. Jisova V. Curing Mechanism of Epoxides by Imidazoles. J. Appl. Polym. Sci. 1987, 34 (7), 2547–2558. 10.1002/app.1987.070340718. [DOI] [Google Scholar]
  35. Maka H.; Spychaj T.; Pilawka R. Epoxy Resin/Ionic Liquid Systems: The Influence of Imidazolium Cation Size and Anion Type on Reactivity and Thermomechanical Properties. Ind. Eng. Chem. Res. 2012, 51 (14), 5197–5206. 10.1021/ie202321j. [DOI] [Google Scholar]
  36. Matejka L.; Pokorny S.; Dusek K. Acid Curing of Epoxy Resins - A Comparison between the Polymerization of Diepoxide-diacid and Monoepoxide-cyclic Anhydride Systems. Makromol. Chem. 1985, 186 (10), 2025–2036. 10.1002/macp.1985.021861006. [DOI] [Google Scholar]
  37. Rocks J.; Rintoul L.; Vohwinkel F.; George G. The Kinetics and Mechanism of Cure of an Amino-glycidyl Epoxy Resin by a Co-anhydride as Studied by FT-Raman Spectroscopy. Polymer 2004, 45 (20), 6799–6811. 10.1016/j.polymer.2004.07.066. [DOI] [Google Scholar]
  38. Fernandez-Francos X.; Ramis X.; Serra A. From Curing Kinetics to Network Structure: A Novel Approach to the Modeling of the Network Buildup of Epoxy-anhydride Thermosets. J. Polym. Sci. A Polym. Chem. 2014, 52 (1), 61–75. 10.1002/pola.26972. [DOI] [Google Scholar]
  39. Dusek K.; Ilavsky M. Formation, Structure, and Elasticity of Loosely Crosslinked Epoxy-amine Networks, 1. Statistics of Formation. J. Polym. Sci. B Polym. Phys. 1983, 21 (8), 1323–1339. 10.1002/pol.1983.180210805. [DOI] [Google Scholar]
  40. Galante M. J.; Williams R. J. J. Polymer Networks Based on the Diepoxide Diisocyanate Reaction Catalyzed by Tertiary-amines. J. Appl. Polym. Sci. 1995, 55 (1), 89–98. 10.1002/app.1995.070550109. [DOI] [Google Scholar]
  41. Yang J. L.; Keller M. W.; Moore J. S.; White S. R.; Sottos N. R. Microencapsulation of Isocyanates for Self-healing Polymers. Macromolecules 2008, 41 (24), 9650–9655. 10.1021/ma801718v. [DOI] [Google Scholar]
  42. Yuan Y. C.; Rong M. Z.; Zhang M. Q.; Yang G. C.; Zhao J. Q. Self-healing of Fatigue Crack in Epoxy Materials with Epoxy/mercaptan System. Express Polym. Lett. 2011, 5 (1), 47–59. 10.3144/expresspolymlett.2011.6. [DOI] [Google Scholar]
  43. Flores M.; Fernandez-Francos X.; Morancho J. M.; Serra A.; Ramis X. Ytterbium Triflate as a New Catalyst on the Curing of Epoxy-isocyanate Based Thermosets. Thermochim. Acta 2012, 543, 188–196. 10.1016/j.tca.2012.05.012. [DOI] [Google Scholar]
  44. Shechter L.; Wynstra J.; Kurkjy R. P. Glycidyl Ether Reactions with Amines. Ind. Eng. Chem. 1956, 48 (1), 94–97. 10.1021/ie50553a029. [DOI] [Google Scholar]
  45. Mora A. S.; Tayouo R.; Boutevin B.; David G.; Caillol S. A Perspective Approach on the Amine Reactivity and the Hydrogen Bonds Effect on Epoxy-amine Systems. Eur. Polym. J. 2020, 123, 109460. 10.1016/j.eurpolymj.2019.109460. [DOI] [Google Scholar]
  46. Thomas R.; Durix S.; Sinturel C.; Omonov T.; Goossens S.; Groeninckx G.; Moldenaers P.; Thomas S. Cure Kinetics, Morphology and Miscibility of Modified DGEBA-based Epoxy Resin - Effects of a Liquid Rubber Inclusion. Polymer 2007, 48 (6), 1695–1710. 10.1016/j.polymer.2007.01.018. [DOI] [Google Scholar]
  47. Duemichen E.; Javdanitehran M.; Erdmann M.; Trappe V.; Sturm H.; Braun U.; Ziegmann G. Analyzing the Network Formation and Curing Kinetics of Epoxy Resins by in Situ Near-infrared Measurements with Variable Heating Rates. Thermochim. Acta 2015, 616, 49–60. 10.1016/j.tca.2015.08.008. [DOI] [Google Scholar]
  48. Gilbert M. D.; Schneider N. S.; MacKnight W. J. Mechanism of the Dicyandiamide Epoxide Reaction. Macromolecules 1991, 24 (2), 360–369. 10.1021/ma00002a004. [DOI] [Google Scholar]
  49. Raetzke K.; Shaikh M. Q.; Faupel F.; Noeske P. L. M. Shelf Stability of Reactive Adhesive Formulations: A Case Study for Dicyandiamide-cured Epoxy Systems. Int. J. Adhes. Adhes. 2010, 30 (2), 105–110. 10.1016/j.ijadhadh.2009.09.004. [DOI] [Google Scholar]
  50. Lin Y. G.; Sautereau H.; Pascault J. P. Processing Property Relationship for Dicyandiamide-cured Epoxy Networks. J. Appl. Polym. Sci. 1987, 33 (2), 685–691. 10.1002/app.1987.070330229. [DOI] [Google Scholar]
  51. Tomuta A. M.; Ramis X.; Ferrando F.; Serra A. The Use of Dihydrazides as Latent Curing Agents in Diglycidyl Ether of Bisphenol A Coatings. Prog. Org. Coat. 2012, 74 (1), 59–66. 10.1016/j.porgcoat.2011.10.004. [DOI] [Google Scholar]
  52. Spindler R.; Fréchet J. M. J. Synthesis and Characterization of Hyperbranched Polyurethanes Prepared from Blocked Isocyanate Monomers by Step-growth Polymerization. Macromolecules 1993, 26 (18), 4809–4813. 10.1021/ma00070a013. [DOI] [Google Scholar]
  53. Wicks D. A.; Wicks Z. W. Blocked Isocyanates III Part B: Uses and Applications of Blocked Isocyanates. Prog. Org. Coat. 2001, 41 (1–3), 1–83. 10.1016/S0300-9440(00)00164-8. [DOI] [Google Scholar]
  54. Suzuki K.; Nobuki M.; Horii H.; Sugita Y.; Sanda F.; Endo T. One-pot Curing System of Epoxy Resin Imines Initiated with Water. J. Appl. Polym. Sci. 2003, 88 (4), 878–882. 10.1002/app.11675. [DOI] [Google Scholar]
  55. Suzuki K.; Matsu-ura N.; Horii H.; Sugita Y.; Sanda F.; Endo T. Diethyl Ketone-based Imine as Efficient Latent Hardener for Epoxy Resin. J. Appl. Polym. Sci. 2002, 83 (8), 1744–1749. 10.1002/app.10101. [DOI] [Google Scholar]
  56. Okuhira H.; Kii T.; Ochi M.; Takeyama H. Novel Moisture-curable Epoxy Resins and Their Characterization. J. Appl. Polym. Sci. 2003, 89 (1), 91–95. 10.1002/app.12151. [DOI] [Google Scholar]
  57. Cameron J. F.; Fréchet J. M. J. Photogeneration of Organic-bases from Ortho-nitrobenzyl-derived Carbamates. J. Am. Chem. Soc. 1991, 113 (11), 4303–4313. 10.1021/ja00011a038. [DOI] [Google Scholar]
  58. Arimitsu K.; Miyamoto M.; Ichimura K. Applications of A Nonlinear Organic Reaction of Carbamates to Proliferate Aliphatic Amines. Angew. Chem., Int. Ed. 2000, 39 (19), 3425–3428. . [DOI] [PubMed] [Google Scholar]
  59. Ito K. I.; Nishimura M.; Sashio M.; Tsunooka M. Thermal Cross-linking of Poly(glycidyl methacrylate) films and Epoxy Resin Films Using Amines Formed by Photolysis of O-Acyloximes. J. Polym. Sci. A Polym. Chem. 1994, 32 (9), 1793–1796. 10.1002/pola.1994.080320922. [DOI] [Google Scholar]
  60. Liu X. D.; Kimura M.; Sudo A.; Endo T. Accelerating Effects of N-Aryl-N ’, N ’-dialkyl Ureas on Epoxy-dicyandiamide Curing System. J. Polym. Sci. A Polym. Chem. 2010, 48 (23), 5298–5305. 10.1002/pola.24329. [DOI] [Google Scholar]
  61. Unruh D. A.; Pastine S. J.; Moreton J. C.; Frechet J. M. J. Thermally Activated, Single Component Epoxy Systems. Macromolecules 2011, 44 (16), 6318–6325. 10.1021/ma200673v. [DOI] [Google Scholar]
  62. Gupta A. M.; Macosko C. W. Modeling Strategy for Systems with Both Stepwise and Chainwise Chemistry - Amine-epoxy Networks with Etherification. J. Polym. Sci. B Polym. Phys. 1990, 28 (13), 2585–2606. 10.1002/polb.1990.090281309. [DOI] [Google Scholar]
  63. Cizmecioglu M.; Gupta A.; Fedors R. F. Influence of Cure Conditions on Glass-transition Temperature and Density of an Epoxy Resin. J. Appl. Polym. Sci. 1986, 32 (8), 6177–6190. 10.1002/app.1986.070320804. [DOI] [Google Scholar]
  64. Montserrat S. Vitrification and Further Structural Relaxation in the Isothermal Curing of an Epoxy Resin. J. Appl. Polym. Sci. 1992, 44 (3), 545–554. 10.1002/app.1992.070440319. [DOI] [Google Scholar]
  65. Rabinowitch E. Collision, Co-ordination, Diffusion and Reaction Velocity in Condensed Systems. Trans. Faraday Soc. 1937, 33 (2), 1225–1232. 10.1039/tf9373301225. [DOI] [Google Scholar]
  66. Wise C. W.; Cook W. D.; Goodwin A. A. Chemico-diffusion Kinetics of Model Epoxy-amine Resins. Polymer 1997, 38 (13), 3251–3261. 10.1016/S0032-3861(96)00882-8. [DOI] [Google Scholar]
  67. Mak H. D.; Rogers M. G. Determination of Chain Branching in Epoxy Resins by Nuclear Magnetic Resonance Spectrometry. Anal. Chem. 1972, 44 (4), 837–839. 10.1021/ac60312a055. [DOI] [PubMed] [Google Scholar]
  68. Cocker R. P.; Chadwick D. L.; Dare D. J.; Challis R. E. A Low Resolution Pulsed NMR and Ultrasound Study to Monitor the Cure of an Epoxy Resin Adhesive. Int. J. Adhes. Adhes. 1998, 18 (5), 319–331. 10.1016/S0143-7496(98)00013-X. [DOI] [Google Scholar]
  69. Podzimek S.; Kastanek A. Characterization of Bisphenol A-based Epoxy Resins by HPLC, GPC, GPC-MALLS, VPO, Viscometry, and End Group Analysis: On the Identification of 2,3-Dihydroxypropyl-containing Compounds, Determination of Molar Mass Distribution, and Branching. J. Appl. Polym. Sci. 1999, 74 (10), 2432–2438. . [DOI] [Google Scholar]
  70. Chian W.; Timm D. C. Kinetic Reaction Analysis of an Anhydride-cured Thermoplastic Epoxy: PGE/NMA/BDMA. Macromolecules 2004, 37 (21), 8091–8097. 10.1021/ma049293x. [DOI] [Google Scholar]
  71. Riccardi C. C.; Adabbo H. E.; Williams R. J. J. Curing Reaction of Epoxy Resins with Diamines. J. Appl. Polym. Sci. 1984, 29 (8), 2481–2492. 10.1002/app.1984.070290805. [DOI] [Google Scholar]
  72. Javdanitehran M.; Berg D. C.; Duemichen E.; Ziegmann G. An Iterative Approach for Isothermal Curing Kinetics Modelling of an Epoxy Resin System. Thermochim. Acta 2016, 623, 72–79. 10.1016/j.tca.2015.11.014. [DOI] [Google Scholar]
  73. Zvetkov V. L.; Calado V. Comparative DSC Kinetics of the Reaction of DGEBA with Aromatic Diamines. III. Formal Kinetic Study of the Reaction of DGEBA with Diamino Diphenyl Methane. Thermochim. Acta 2013, 560, 95–103. 10.1016/j.tca.2013.02.017. [DOI] [Google Scholar]
  74. Moroni A.; Mijovic J.; Pearce E. M.; Foun C. C. Cure Kinetics of Epoxy Resins and Aromatic Diamines. J. Appl. Polym. Sci. 1986, 32 (2), 3761–3773. 10.1002/app.1986.070320231. [DOI] [Google Scholar]
  75. Chabert B.; Lachenal G.; Tung C. V. Epoxy Resins and Epoxy Blends Studied by Near-infrared Spectroscopy. Macromolecul. Symp. 1995, 94 (1), 145–158. 10.1002/masy.19950940113. [DOI] [Google Scholar]
  76. Gonzalez M. G.; Cabanelas J. C.; Baselga J.. Applications of FTIR on Epoxy Resins - Identification, Monitoring the Curing Process, Phase Separation and Water Uptake. In Infrared Spectroscopy - Materials Science, Engineering and Technology; Theophanides T., Eds.; InTech, 2012; pp 261–284. [Google Scholar]
  77. Huang C. Y.; Sun X. T.; Yuan H. F.; Song C. F.; Meng Y.; Li X. Y. Study on the Reactivity and Kinetics of Primary and Secondary Amines During Epoxy Curing by NIR Spectroscopy Combined with Multivariate Analysis. Vib. Spectrosc. 2020, 106, 102993. 10.1016/j.vibspec.2019.102993. [DOI] [Google Scholar]
  78. Pandita S. D.; Wang L. W.; Mahendran R. S.; Machavaram V. R.; Irfan M. S.; Harris D.; Fernando G. F. Simultaneous DSC-FTIR Spectroscopy: Comparison of Cross-linking Kinetics of an Epoxy/amine Resin System. Thermochim. Acta 2012, 543 (10), 9–17. 10.1016/j.tca.2012.04.024. [DOI] [Google Scholar]
  79. Sbirrazzuoli N.; Mititelu-Mija A.; Vincent L.; Alzina C. Isoconversional Kinetic Analysis of Stoichiometric and Off-stoichiometric Epoxy-amine Cures. Thermochim. Acta 2006, 447 (2), 167–177. 10.1016/j.tca.2006.06.005. [DOI] [Google Scholar]
  80. Sourour S.; Kamal M. R. Differential Scanning Calorimetry of Epoxy Cure - Isothermal Cure Kinetics. Thermochim. Acta 1976, 14 (1–2), 41–59. 10.1016/0040-6031(76)80056-1. [DOI] [Google Scholar]
  81. Zvetkov V. L.; Krastev R. K.; Paz-Abuin S. Is the Kamal’s Model Appropriate in the Study of the Epoxy-amine Addition Kinetics?. Thermochim. Acta 2010, 505 (1–2), 47–52. 10.1016/j.tca.2010.03.024. [DOI] [Google Scholar]
  82. Dušek K.; Havlíc̆ek I. Diffusion-controlled Kinetics of Cross-linking. Prog. Org. Coat. 1993, 22 (1–4), 145–159. 10.1016/0033-0655(93)80020-B. [DOI] [Google Scholar]
  83. Chern C. S.; Poehlein G. W. A Kinetic-model for Curing Reactions of Epoxides with Amines. Polym. Eng. Sci. 1987, 27 (11), 788–795. 10.1002/pen.760271104. [DOI] [Google Scholar]
  84. Wisanrakkit G.; Gillham J. K. The Glass-transition Temperature (Tg) as an Index of Chemical Conversion for a High-Tg Amine Epoxy System - Chemical and Diffusion-controlled Reaction-kinetics. J. Appl. Polym. Sci. 1990, 41 (11–12), 2885–2929. 10.1002/app.1990.070411129. [DOI] [Google Scholar]
  85. Charlesworth J. Analysis of the Substitution Effects Involved in Diepoxide-diamine Copolymerization Reactions. J. Polym. Sci. A Polym. Chem. 1980, 18 (2), 621–628. 10.1002/pol.1980.170180221. [DOI] [Google Scholar]
  86. Mijovic J.; Fishbain A.; Wijaya J. Mechanistic Modeling of Epoxy Amine Kinetics, 1. Model-compound Study. Macromolecules 1992, 25 (2), 979–985. 10.1021/ma00028a072. [DOI] [Google Scholar]
  87. Stjohn N. A.; George G. A. Cure Kinetics and Mechanisms of A Tetraglycidyl-4,4’-diaminodiphenylmethane Diaminodiphenylsulphone Epoxy Resin Using Near IR Spectroscopy. Polymer 1992, 33 (13), 2679–2688. 10.1016/0032-3861(92)90438-3. [DOI] [Google Scholar]
  88. Aoki M.; Shundo A.; Yamamoto S.; Tanaka K. Effect of A Heterogeneous Network on Glass Transition Dynamics and Solvent Crack Behavior of Epoxy Resins. Soft Matter 2020, 16 (32), 7470–7478. 10.1039/D0SM00625D. [DOI] [PubMed] [Google Scholar]
  89. PazAbuin S.; Pellin M. P.; PazPazos M.; LopezQuintela A. Influence of the Reactivity of Amine Hydrogens and the Evaporation of Monomers on the Cure Kinetics of Epoxy-amine: Kinetic Questions. Polymer 1997, 38 (15), 3795–3804. 10.1016/S0032-3861(96)00957-3. [DOI] [Google Scholar]
  90. Sahagun C. M.; Morgan S. E. Thermal Control of Nanostructure and Molecular Network Development in Epoxy-amine Thermosets. ACS Appl. Mater. Interfaces 2012, 4 (2), 564–572. 10.1021/am201515y. [DOI] [PubMed] [Google Scholar]
  91. Dušek K.; Pleštil J.; Lednický F.; Luňák S. Are Cured Epoxy Resins Inhomogeneous. Polymer 1978, 19 (4), 393–397. 10.1016/0032-3861(78)90243-4. [DOI] [Google Scholar]
  92. Mijovic J.; Koutsky J. A. Correlation between Nodular Morphology and Fracture Properties of Cured Epoxy Resins. Polymer 1979, 20 (9), 1095–1107. 10.1016/0032-3861(79)90301-X. [DOI] [Google Scholar]
  93. Takahama T.; Geil P. H. Structural Inhomogeneities of Cured Epoxy Resins. Makromol. Chem. Rapid Commun. 1982, 3 (6), 389–394. 10.1002/marc.1982.030030603. [DOI] [Google Scholar]
  94. Gupta V. B.; Drzal L. T.; Adams W. W.; Omlor R. An Electron-microscopic Study of the Morphology of Cured Epoxy resin. J. Mater. Sci. 1985, 20 (10), 3439–3452. 10.1007/BF01113751. [DOI] [Google Scholar]
  95. Sahagun C. M.; Knauer K. M.; Morgan S. E. Molecular Network Development and Evolution of Nanoscale Morphology in an Epoxy-amine Thermoset Polymer. J. Appl. Polym. Sci. 2012, 126 (4), 1394–1405. 10.1002/app.36763. [DOI] [Google Scholar]
  96. Haba D.; Kaufmann J.; Brunner A. J.; Resch K.; Teichert C. Observation of Elastic Modulus Inhomogeneities in Thermosetting Epoxies Using AFM - Discerning Facts and Artifacts. Polymer 2014, 55 (16), 4032–4040. 10.1016/j.polymer.2014.06.030. [DOI] [Google Scholar]
  97. Dazzi A.; Prater C. B. AFM-IR: Technology and Applications in Nanoscale Infrared Spectroscopy and Chemical Imaging. Chem. Rev. 2017, 117 (7), 5146–5173. 10.1021/acs.chemrev.6b00448. [DOI] [PubMed] [Google Scholar]
  98. Morsch S.; Liu Y. W.; Lyon S. B.; Gibbon S. R. Insights into Epoxy Network Nanostructural Heterogeneity Using AFM-IR. ACS Appl. Mater. Interfaces 2016, 8 (1), 959–966. 10.1021/acsami.5b10767. [DOI] [PubMed] [Google Scholar]
  99. Morsch S.; Liu Y. W.; Greensmith P.; Lyon S. B.; Gibbon S. R. Molecularly Controlled Epoxy Network Nanostructures. Polymer 2017, 108 (1), 146–153. 10.1016/j.polymer.2016.11.050. [DOI] [Google Scholar]
  100. Izumi A.; Nakao T.; Shibayama M. Gelation and Cross-link Inhomogeneity of Phenolic Resins Studied by 13C-NMR Spectroscopy and Small-angle X-ray Scattering. Soft Matter 2013, 9 (16), 4188–4197. 10.1039/c3sm27438a. [DOI] [Google Scholar]
  101. Shudo Y.; Izumi A.; Takeuchi T.; Nakao T.; Shibayama M. Dynamic Light Scattering Study of the Curing Mechanisms of Novolac-type Phenolic Resins. Polym. J. 2015, 47 (6), 428–433. 10.1038/pj.2015.15. [DOI] [Google Scholar]
  102. Izumi A.; Shudo Y.; Nakao T.; Shibayama M. Cross-link Inhomogeneity in Phenolic Resins at the Initial Stage of Curing Studied by 1H-Pulse NMR Spectroscopy and Complementary SAXS/WAXS and SANS/WANS with a Solvent-swelling Technique. Polymer 2016, 103 (10), 152–162. 10.1016/j.polymer.2016.09.067. [DOI] [Google Scholar]
  103. Izumi A.; Nakao T.; Shibayama M. Gelation and Cross-link Inhomogeneity of Phenolic Resins Studied by Small- and Wide-angle X-ray Scattering and 1H-Pulse NMR Spectroscopy. Polymer 2015, 59 (2), 226–233. 10.1016/j.polymer.2015.01.018. [DOI] [Google Scholar]
  104. Aoki M.; Shundo A.; Kuwahara R.; Yamamoto S.; Tanaka K. Mesoscopic Heterogeneity in the Curing Process of an Epoxy-amine System. Macromolecules 2019, 52 (5), 2075–2082. 10.1021/acs.macromol.8b02416. [DOI] [Google Scholar]
  105. Enns J. B.; Gillham J. K. Time Temperature Transformation (TTT) Cure Diagram - Modeling the Cure Behavior of Thermosets. J. Appl. Polym. Sci. 1983, 28 (8), 2567–2591. 10.1002/app.1983.070280810. [DOI] [Google Scholar]
  106. Zhang C.; Wang Y. C.; Liu Y. H. Construction of Improved Isothermal TTT Cure Diagram Based on an Epoxy-amine Thermoset. J. Appl. Polym. Sci. 2019, 136 (13), 47279. 10.1002/app.47279. [DOI] [Google Scholar]
  107. Mangialetto J.; Verhelle R.; Van Assche G.; Van den Brande N.; Van Mele B. Time-temperature-transformation, Temperature-conversion-transformation, and Continuous-heating-transformation Diagrams of Reversible Covalent Polymer Networks. Macromolecules 2021, 54 (1), 412–425. 10.1021/acs.macromol.0c02491. [DOI] [Google Scholar]
  108. Chen Y. C.; Chiu W. Y. The Structural Properties of Imidazole Cured Epoxy-phenol Resins. Polymer 2001, 42 (12), 5439–5448. 10.1016/S0032-3861(01)00024-6. [DOI] [Google Scholar]
  109. Chiou P. L.; Letton A. Modeling the Chemorheology of an Epoxy Resin System Exhibiting Complex Curing Behavior. Polymer 1992, 33 (18), 3925–3931. 10.1016/0032-3861(92)90384-9. [DOI] [Google Scholar]
  110. Winter H. H. Can the Gel Point of a Cross-linking Polymer Be Detected by the G’-G’’ Crossover. Polym. Eng. Sci. 1987, 27 (22), 1698–1702. 10.1002/pen.760272209. [DOI] [Google Scholar]
  111. Halley P. J.; Mackay M. E. Chemorheology of Thermosets - An Overview. Polym. Eng. Sci. 1996, 36 (5), 593–609. 10.1002/pen.10447. [DOI] [Google Scholar]
  112. Mortimer S.; Ryan A. J.; Stanford J. L. Rheological Behavior and Gel-point Determination for A Model Lewis Acid-initiated Chain Growth Epoxy Resin. Macromolecules 2001, 34 (9), 2973–2980. 10.1021/ma001835x. [DOI] [Google Scholar]
  113. ASTM International . Standard Test Method for Gel Time and Peak Exothermic Temperature of Reacting Thermosetting Resins; ASTM D2471-99; West Conshohocken, PA, 1999. [Google Scholar]
  114. Tung C. Y. M.; Dynes P. J. Relationship between Viscoelastic Properties and Gelation in Thermosetting Systems. J. Appl. Polym. Sci. 1982, 27 (2), 569–574. 10.1002/app.1982.070270220. [DOI] [Google Scholar]
  115. Matejka L. Rheology of Epoxy Networks Near the Gel Point. Polym. Bull. 1991, 26 (1), 109–116. 10.1007/BF00299355. [DOI] [Google Scholar]
  116. Lairez D.; Adam M.; Emery J. R.; Durand D. Rheological Behavior of an Epoxy Amine System Near the Gel Point. Macromolecules 1992, 25 (1), 286–289. 10.1021/ma00027a046. [DOI] [Google Scholar]
  117. Eloundou J. P.; Gerard J. F.; Harran D.; Pascault J. P. Temperature Dependence of the Behavior of a Reactive Epoxy-amine System by Means of Dynamic Rheology, 2. High-Tg Epoxy-amine System. Macromolecules 1996, 29 (21), 6917–6927. 10.1021/ma9602886. [DOI] [Google Scholar]
  118. Lange J.; Altmann N.; Kelly C. T.; Halley P. J. Understanding Vitrification during Cure of Epoxy Resins Using Dynamic Scanning Calorimetry and Rheological Techniques. Polymer 2000, 41 (15), 5949–5955. 10.1016/S0032-3861(99)00758-2. [DOI] [Google Scholar]
  119. Meng Y.; Simon S. L. Relation between Mobility Factor and Diffusion Factor for Thermoset Cure. Thermochim. Acta 2005, 437 (1–2), 179–189. 10.1016/j.tca.2005.06.032. [DOI] [Google Scholar]
  120. Valentine M. T.; Kaplan P. D.; Thota D.; Crocker J. C.; Gisler T.; Prud’homme R. K.; Beck M.; Weitz D. A. Investigating the Microenvironments of Inhomogeneous Soft Materials with Multiple Particle Tracking. Phys. Rev. E 2001, 64 (6), 061506. 10.1103/PhysRevE.64.061506. [DOI] [PubMed] [Google Scholar]
  121. Waigh T. A. Microrheology of Complex Fluids. Rep. Prog. Phys. 2005, 68 (3), 685–742. 10.1088/0034-4885/68/3/R04. [DOI] [PubMed] [Google Scholar]
  122. Penaloza D. P.; Hori K.; Shundo A.; Tanaka K. Spatial Heterogeneity in A Lyotropic Liquid Crystal with Hexagonal Phase. Phys. Chem. Chem. Phys. 2012, 14 (15), 5247–5250. 10.1039/c2cp40284j. [DOI] [PubMed] [Google Scholar]
  123. Penaloza D. P.; Shundo A.; Matsumoto K.; Ohno M.; Miyaji K.; Goto M.; Tanaka K. Spatial Heterogeneity in the Sol-gel Transition of a Supramolecular System. Soft Matter 2013, 9 (21), 5166–5172. 10.1039/c3sm50225b. [DOI] [Google Scholar]
  124. Shundo A.; Hori K.; Penaloza D. P.; Matsumoto Y.; Okumura Y.; Kikuchi H.; Lee K. E.; Kim S. O.; Tanaka K. Hierarchical Spatial Heterogeneity in Liquid Crystals Composed of Graphene Oxides. Phys. Chem. Chem. Phys. 2016, 18 (32), 22399–22406. 10.1039/C6CP03614G. [DOI] [PubMed] [Google Scholar]
  125. Matsumoto Y.; Shundo A.; Ohno M.; Tsuruzoe N.; Goto M.; Tanaka K. Mesoscopic Heterogeneity in Pore Size of Supramolecular Networks. Langmuir 2018, 34 (25), 7503–7508. 10.1021/acs.langmuir.8b00641. [DOI] [PubMed] [Google Scholar]
  126. Matsumoto Y.; Shundo A.; Hayashi H.; Tsuruzoe N.; Tanaka K. Effect of the Heterogeneous Structure on Mechanical Properties for a Nanocellulose-reinforced Polymer Composite. Macromolecules 2019, 52 (21), 8266–8274. 10.1021/acs.macromol.9b01866. [DOI] [Google Scholar]
  127. Shundo A.; Matsumoto Y.; Hayashi H.; Tsuruzoe N.; Matsuno H.; Tanaka K. Mesoscopic Heterogeneity in a Nanocellulose-containing Cell Storage Medium. J. Mater. Chem. B 2020, 8 (21), 4570–4574. 10.1039/D0TB00219D. [DOI] [PubMed] [Google Scholar]
  128. Kogo T.; Shundo A.; Wang C.; Tanaka K. Spatial Heterogeneity Accompanying Gel Formation of Poly(N-isopropylacrylamide) Aqueous Solution at a Temperature below Cloud Point. Macromolecules 2020, 53 (24), 10964–10971. 10.1021/acs.macromol.0c02292. [DOI] [Google Scholar]
  129. Nguyen H. K.; Aoki M.; Liang X.; Yamamoto S.; Tanaka K.; Nakajima K. Local Mechanical Properties of Heterogeneous Nanostructures Developed in a Cured Epoxy Network: Implications for Innovative Adhesion Technology. ACS Applied Nano Materials 2021, 4 (11), 12188–12196. 10.1021/acsanm.1c02692. [DOI] [Google Scholar]
  130. Li C. Y.; Strachan A. Molecular Scale Simulations on Thermoset Polymers: A Review. J. Polym. Sci. B Polym. Phys. 2015, 53 (2), 103–122. 10.1002/polb.23489. [DOI] [Google Scholar]
  131. Lyulin S. V.; Larin S. V.; Nazarychev V. M.; Fal’kovich S. G.; Kenny J. M. Multiscale Computer Simulation of Polymer Nanocomposites Based on Thermoplastics. Polym. Sci. Ser. C 2016, 58 (1), 2–15. 10.1134/S1811238216010082. [DOI] [Google Scholar]
  132. Liu Z.; Li J. H.; Zhou C.; Zhu W. H. A Molecular Dynamics Study on Thermal and Rheological Properties of BNNS-epoxy Nanocomposites. Int. J. Heat Mass Transfer 2018, 126, 353–362. 10.1016/j.ijheatmasstransfer.2018.05.149. [DOI] [Google Scholar]
  133. Hsissou R. Review on Epoxy Polymers and Its Composites as A Potential Anticorrosive Coatings for Carbon Steel in 3.5% NaCl Solution: Computational Approaches. J. Mol. Liq. 2021, 336, 116307. 10.1016/j.molliq.2021.116307. [DOI] [Google Scholar]
  134. Yarovsky I.; Evans E. Computer Simulation of Structure and Properties of Crosslinked Polymers: Application to Epoxy Resins. Polymer 2002, 43 (3), 963–969. 10.1016/S0032-3861(01)00634-6. [DOI] [Google Scholar]
  135. Wu C. F.; Xu W. J. Atomistic Molecular Modelling of Crosslinked Epoxy Resin. Polymer 2006, 47 (16), 6004–6009. 10.1016/j.polymer.2006.06.025. [DOI] [Google Scholar]
  136. Komarov P. V.; Chiu Y. T.; Chen S. M.; Khalatur P. G.; Reineker P. Highly Cross-linked Epoxy Resins: An Atomistic Molecular Dynamics Simulation Combined with A Mapping/reverse Mapping Procedure. Macromolecules 2007, 40 (22), 8104–8113. 10.1021/ma070702+. [DOI] [Google Scholar]
  137. Varshney V.; Patnaik S. S.; Roy A. K.; Farmer B. L. A Molecular Dynamics Study of Epoxy-based Networks: Cross-linking Procedure and Prediction of Molecular and Material Properties. Macromolecules 2008, 41 (18), 6837–6842. 10.1021/ma801153e. [DOI] [Google Scholar]
  138. Bandyopadhyay A.; Valavala P. K.; Clancy T. C.; Wise K. E.; Odegard G. M. Molecular Modeling of Crosslinked Epoxy Polymers: The Effect of Crosslink Density on Thermomechanical Properties. Polymer 2011, 52 (11), 2445–2452. 10.1016/j.polymer.2011.03.052. [DOI] [Google Scholar]
  139. Heine D. R.; Grest G. S.; Lorenz C. D.; Tsige M.; Stevens M. J. Atomistic Simulations of End-linked Poly(dimethylsiloxane) Networks: Structure and Relaxation. Macromolecules 2004, 37 (10), 3857–3864. 10.1021/ma035760j. [DOI] [Google Scholar]
  140. Khare R.; Paulaitis M. E.; Lustig S. R. Generation of Glass Structures for Molecular Simulations of Polymers Containing Large Monomer Units - Application to Polystyrene. Macromolecules 1993, 26 (26), 7203–7209. 10.1021/ma00078a014. [DOI] [Google Scholar]
  141. Lin P. H.; Khare R. Molecular Simulation of Cross-linked Epoxy and Epoxy-POSS Nanocomposite. Macromolecules 2009, 42 (12), 4319–4327. 10.1021/ma9004007. [DOI] [Google Scholar]
  142. Rappe A. K.; Goddard W. A. Charge Equilibration for Molecular-dynamics Simulations. J. Phys. Chem. 1991, 95 (8), 3358–3363. 10.1021/j100161a070. [DOI] [Google Scholar]
  143. Li C. Y.; Strachan A. Molecular Simulations of Crosslinking Process of Thermosetting Polymers. Polymer 2010, 51 (25), 6058–6070. 10.1016/j.polymer.2010.10.033. [DOI] [Google Scholar]
  144. Okabe T.; Takehara T.; Inose K.; Hirano N.; Nishikawa M.; Uehara T. Curing Reaction of Epoxy Resin Composed of Mixed Base Resin and Curing Agent: Experiments and Molecular Simulation. Polymer 2013, 54 (17), 4660–4668. 10.1016/j.polymer.2013.06.026. [DOI] [Google Scholar]
  145. Oya Y.; Nakazawa M.; Shirasu K.; Hino Y.; Inuyama K.; Kikugawa G.; Li J.; Kuwahara R.; Kishimoto N.; Waizumi H.; Nishikawa M.; Waas A.; Odagiri N.; Koyanagi A.; Salviato M.; Okabe T. Molecular Dynamics Simulation of Cross-linking Processes and Material Properties for Epoxy Resins Using First-principle Calculation Combined with Global Reaction Route Mapping Algorithms. Chem. Phys. Lett. 2021, 762, 138104. 10.1016/j.cplett.2020.138104. [DOI] [Google Scholar]
  146. Yamamoto S.; Tanaka K. Molecular Size Effect on Curing Process for Epoxy and Amine Mixture. Nihon Reoroji Gakkaishi 2021, 49 (2), 55–60. 10.1678/rheology.49.55. [DOI] [Google Scholar]
  147. Yamamoto S.; Kuwahara R.; Tanaka K. Dynamic Behaviour of Water Molecules in Heterogeneous Free Space Formed in an Epoxy Resin. Soft Matter 2021, 17 (25), 6073–6080. 10.1039/D1SM00529D. [DOI] [PubMed] [Google Scholar]
  148. Levita G.; Depetris S.; Marchetti A.; Lazzeri A. Cross-link Density and Fracture-toughness of Epoxy-resins. J. Mater. Sci. 1991, 26 (9), 2348–2352. 10.1007/BF01130180. [DOI] [Google Scholar]
  149. Rahul R.; Kitey R. Effect of Cross-linking on Dynamic Mechanical and Fracture Behavior of Epoxy Variants. Compos. B Eng. 2016, 85, 336–342. 10.1016/j.compositesb.2015.09.017. [DOI] [Google Scholar]
  150. Pandini S.; Bignotti F.; Baldi F.; Sartore L.; Consolati G.; Panzarasa G. Thermomechanical and Large Deformation Behaviors of Antiplasticized Epoxy Resins: Effect of Material Formulation and Network Architecture. Polym. Eng. Sci. 2017, 57 (6), 553–565. 10.1002/pen.24555. [DOI] [Google Scholar]
  151. Cook W. D.; Mehrabi M.; Edward G. H. Ageing and Yielding in Model Epoxy Thermosets. Polymer 1999, 40 (5), 1209–1218. 10.1016/S0032-3861(98)00343-7. [DOI] [Google Scholar]
  152. Blanco M.; Corcuera M. A.; Riccardi C. C.; Mondragon I. Mechanistic Kinetic Model of an Epoxy Resin Cured with a Mixture of Amines of Different Functionalities. Polymer 2005, 46 (19), 7989–8000. 10.1016/j.polymer.2005.06.117. [DOI] [Google Scholar]
  153. Koike T. Determination of Glass Transition Temperature from Viscosity and Conductivity Measurements for an Epoxy-amine System During Curing. J. Appl. Polym. Sci. 1993, 50 (11), 1943–1950. 10.1002/app.1993.070501111. [DOI] [Google Scholar]
  154. Lesser A. J.; Crawford E. The Role of Network Architecture on the Glass Transition Temperature of Epoxy Resins. J. Appl. Polym. Sci. 1997, 66 (2), 387–395. . [DOI] [Google Scholar]
  155. Shundo A.; Aoki M.; Yamamoto S.; Tanaka K. Cross-linking Effect on Segmental Dynamics of Well-defined Epoxy Resins. Macromolecules 2021, 54 (13), 5950–5956. 10.1021/acs.macromol.1c00513. [DOI] [Google Scholar]
  156. Stutz H.; Illers K. H.; Mertes J. A Generalized Theory for the Glass-transition Temperature of Cross-linked and Uncrosslinked Polymers. J. Polym. Sci., Part B: Polym. Phys. 1990, 28 (9), 1483–1498. 10.1002/polb.1990.090280906. [DOI] [Google Scholar]
  157. Jeffrey K.; Pethrick R. A. Influence of Chemical Structure on Free Volume in Epoxy Resins: A Positron Annihilation Study. Eur. Polym. J.. 1994, 30 (2), 153–158. 10.1016/0014-3057(94)90153-8. [DOI] [Google Scholar]
  158. Goyanes S.; Salguleiro W.; Somoza A.; Ramos J. A.; Mondragon I. Direct Relationships between Volume Variations at Macro and Nanoscale in Epoxy Systems. PALS/PVT Measurements. Polymer 2004, 45 (19), 6691–6697. 10.1016/j.polymer.2004.07.057. [DOI] [Google Scholar]
  159. Xie Q.; Liang S. D.; Liu B. W.; Fu K. X.; Zhan Z. Y.; Lu L.; Yang X. M.; Lu F. C.; Huang Z. Y. Structure, Microparameters and Properties of Crosslinked DGEBA/MTHPA: A Molecular Dynamics Simulation. AIP Adv. 2018, 8, 075332. 10.1063/1.5041283. [DOI] [Google Scholar]
  160. Uedono A.; Aiko W.; Yamamoto T.; Nakamichi T.; Tanigawa S. Open Spaces and Molecular Motions of Acrylic Epoxy-based Network Polymers Probed by Positron Annihilation. J. Polym. Sci., Part B: Polym. Phys. 1999, 37 (20), 2875–2880. . [DOI] [Google Scholar]
  161. Salgueiro W.; Ramos J.; Somoza A.; Goyanes S.; Mondragon I. Nanohole Volume Dependence on the Cure Schedule in Epoxy Thermosetting Networks: A PALS Study. Polymer 2006, 47 (14), 5066–5070. 10.1016/j.polymer.2006.05.032. [DOI] [Google Scholar]
  162. Patil P. N.; Rath S. K.; Sharma S. K.; Sudarshan K.; Maheshwari P.; Patri M.; Praveen S.; Khandelwal P.; Pujari P. K. Free Volumes and Structural Relaxations in Diglycidyl Ether of Bisphenol-A Based Epoxy-polyether Amine Networks. Soft Matter 2013, 9 (13), 3589–3599. 10.1039/c3sm27525f. [DOI] [Google Scholar]
  163. 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 (5), 776–791. . [DOI] [Google Scholar]
  164. Odegard G. M.; Bandyopadhyay A. Physical Aging of Epoxy Polymers and Their Composites. J. Polym. Sci., Part B: Polym. Phys. 2011, 49 (24), 1695–1716. 10.1002/polb.22384. [DOI] [Google Scholar]
  165. 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 (20), 4528–4535. 10.1016/j.polymer.2011.07.042. [DOI] [Google Scholar]
  166. Jilani W.; Mzabi N.; Fourati N.; Zerrouki C.; Gallot-Lavallee O.; Zerrouki R.; Guermazi H. Effects of Curing Agent on Conductivity, Structural and Dielectric Properties of an Epoxy Polymer. Polymer 2015, 79 (11), 73–81. 10.1016/j.polymer.2015.09.078. [DOI] [Google Scholar]
  167. Garden L.; Pethrick R. A. A Dielectric Study of Water Uptake in Epoxy Resin Systems. J. Appl. Polym. Sci. 2017, 134 (16), 44717. 10.1002/app.44717. [DOI] [Google Scholar]
  168. Williams J. G. Beta-relaxation in Epoxy Resin-based Networks. J. Appl. Polym. Sci. 1979, 23 (12), 3433–3444. 10.1002/app.1979.070231201. [DOI] [Google Scholar]
  169. Wang Y. M.; Chen S. F.; Chen X. C.; Lu Y. F.; Miao M. H.; Zhang D. H. Controllability of Epoxy Equivalent Weight and Performance of Hyperbranched Epoxy Resins. Composites, Part B 2019, 160, 615–625. 10.1016/j.compositesb.2018.12.103. [DOI] [Google Scholar]
  170. Feng J. M.; Guo Z. S. Temperature-frequency-dependent Mechanical Properties Model of Epoxy Resin and Its Composites. Composites, Part B 2016, 85, 161–169. 10.1016/j.compositesb.2015.09.040. [DOI] [Google Scholar]
  171. Heux L.; Halary J. L.; Laupretre F.; Monnerie L. Dynamic Mechanical and 13C NMR Investigations of Molecular Motions Involved in the Beta Relaxation of Epoxy Networks Based on DGEBA and Aliphatic Amines. Polymer 1997, 38 (8), 1767–1778. 10.1016/S0032-3861(96)00694-5. [DOI] [Google Scholar]
  172. Bershtein V. A.; Peschanskaya N. N.; Halary J. L.; Monnerie L. The Sub-Tg Relaxations in Pure and Antiplasticized Model Epoxy Networks as Studied by High Resolution Creep Rate Spectroscopy. Polymer 1999, 40 (24), 6687–6698. 10.1016/S0032-3861(98)00873-8. [DOI] [Google Scholar]
  173. Garcia F. G.; Soares B. G.; Pita V.; Sanchez R.; Rieumont J. Mechanical Properties of Epoxy Networks Based on DGEBA and Aliphatic Amines. J. Appl. Polym. Sci. 2007, 106 (3), 2047–2055. 10.1002/app.24895. [DOI] [Google Scholar]
  174. Cook W. D.; Mayr A. E.; Edward G. H. Yielding Behaviour in Model Epoxy Thermosets - II. Temperature Dependence. Polymer 1998, 39 (16), 3725–3733. 10.1016/S0032-3861(97)10335-4. [DOI] [Google Scholar]
  175. Tcharkhtchi A.; Trotignon J. P.; Verdu J. Yielding and fracture in crosslinked epoxies. Makromol. Chem. Macromol. Symp. 1999, 147, 221–234. 10.1002/masy.19991470122. [DOI] [Google Scholar]
  176. Angell C. A. Formation of Glasses from Liquids and Biopolymers. Science 1995, 267 (5206), 1924–1935. 10.1126/science.267.5206.1924. [DOI] [PubMed] [Google Scholar]
  177. Kanaya T.; Tsukushi I.; Kaji K. Non-Gaussian Parameter and Heterogeneity of Amorphous Polymers. Prog. Theor. Phys. Suppl. 1997, 126 (126), 133–140. 10.1143/PTPS.126.133. [DOI] [Google Scholar]
  178. Adam G.; Gibbs J. H. On Temperature Dependence of Cooperative Relaxation Properties in Glass-forming Liquids. J. Chem. Phys. 1965, 43 (1), 139–146. 10.1063/1.1696442. [DOI] [Google Scholar]
  179. Donth E. The Size of Cooperatively Rearranging Regions at the Glass Transition. J. Non-Cryst. Solids 1982, 53 (3), 325–330. 10.1016/0022-3093(82)90089-8. [DOI] [Google Scholar]
  180. Matsuoka S.; Quan X. A Model for Intermolecular Cooperativity in Conformational Relaxations Near the Glass-transition. Macromolecules 1991, 24 (10), 2770–2779. 10.1021/ma00010a021. [DOI] [Google Scholar]
  181. Achibat T.; Boukenter A.; Duval E.; Lorentz G.; Etienne S. Low-frequency Raman Scattering and Structure of Amorphous Polymers Stretching Effect. J. Chem. Phys. 1991, 95 (4), 2949–2954. 10.1063/1.460896. [DOI] [Google Scholar]
  182. Hong L.; Novikov V. N.; Sokolov A. P. Is There a Connection between Fragility of Glass Forming Systems and Dynamic Heterogeneity/Cooperativity?. J. Non-Cryst. Solids 2011, 357 (2), 351–356. 10.1016/j.jnoncrysol.2010.06.071. [DOI] [Google Scholar]
  183. Fischer E. W.; Donth E.; Steffen W. Temperature Dependence of Characteristic Length for Glass Transition. Phys. Rev. Lett. 1992, 68 (15), 2344–2346. 10.1103/PhysRevLett.68.2344. [DOI] [PubMed] [Google Scholar]
  184. Hong L.; Gujrati P. D.; Novikov V. N.; Sokolov A. P. Molecular Cooperativity in the Dynamics of Glass-forming Systems: A New Insight. J. Chem. Phys. 2009, 131 (19), 194511. 10.1063/1.3266508. [DOI] [PubMed] [Google Scholar]
  185. Tracht U.; Wilhelm M.; Heuer A.; Feng H.; Schmidt-Rohr K.; Spiess H. W. Length Scale of Dynamic Heterogeneities at the Glass Transition Determined by Multidimensional Nuclear Magnetic Resonance. Phys. Rev. Lett. 1998, 81 (13), 2727–2730. 10.1103/PhysRevLett.81.2727. [DOI] [Google Scholar]
  186. Qiu X. H.; Ediger M. D. Length Scale of Dynamic Heterogeneity in Supercooled D-Sorbitol: Comparison to Model Predictions. J. Phys. Chem. B 2003, 107 (2), 459–464. 10.1021/jp021888b. [DOI] [Google Scholar]
  187. Hempel E.; Hempel G.; Hensel A.; Schick C.; Donth E. Characteristic Length of Dynamic Glass Transition Near Tg For a Wide Assortment of Glass-forming Substances. J. Phys. Chem. B 2000, 104 (11), 2460–2466. 10.1021/jp991153f. [DOI] [Google Scholar]
  188. Rijal B.; Delbreilh L.; Saiter A. Dynamic Heterogeneity and Cooperative Length Scale at Dynamic Glass Transition in Glass Forming Liquids. Macromolecules 2015, 48 (22), 8219–8231. 10.1021/acs.macromol.5b01152. [DOI] [Google Scholar]
  189. Jilani W.; Mzabi N.; Fourati N.; Zerrouki C.; Gallot-Lavallee O.; Zerrouki R.; Guermazi H. A Comparative Study of Structural and Dielectric Properties of Diglycidyl Ether of Bisphenol A (DGEBA) Cured with Aromatic or Aliphatic Hardeners. J. Mater. Sci. 2016, 51 (17), 7874–7886. 10.1007/s10853-016-0043-0. [DOI] [Google Scholar]
  190. Shundo A.; Aoki M.; Yamamoto S.; Tanaka K. Effect of Cross-linking Density on Horizontal and Vertical Shift Factors in Linear Viscoelastic Functions of Epoxy Resins. Macromolecules 2021, 54 (20), 9618–9624. 10.1021/acs.macromol.1c01293. [DOI] [Google Scholar]
  191. Huang Y.; Kinloch A. J. The Use of Time-temperature Superpositioning in Studying the Fracture Properties of Rubber-toughened Epoxy Polymers. J. Adhes. 1993, 41 (1–4), 5–22. 10.1080/00218469308026551. [DOI] [Google Scholar]
  192. Ghosh S. K.; Rajesh P.; Srikavya B.; Rathore D. K.; Prusty R. K.; Ray B. C. Creep Behaviour Prediction of Multi-layer Graphene Embedded Glass Fiber/Epoxy Composites Using Time-temperature Superposition Principle. Composites, Part A 2018, 107, 507–518. 10.1016/j.compositesa.2018.01.030. [DOI] [Google Scholar]
  193. Anand A.; Banerjee P.; Prusty R. K.; Ray B. C. Lifetime Prediction of Nano-silica Based Glass Fibre/Epoxy Composite by Time Temperature Superposition Principle. IOP Conf. Ser.: Mater. Sci. Eng. 2018, 338, 012020. 10.1088/1757-899X/338/1/012020. [DOI] [Google Scholar]
  194. Pohl M.; Kupfer R.; Koch I.; Modler N.; Hufenbach W. A. Determination of the Long-term Properties in Laminate-thickness Direction of Textile-reinforced Thermoplastic Composites under Compression Using Time-temperature Superposition. Adv. Eng. Mater. 2016, 18 (3), 369–375. 10.1002/adem.201500475. [DOI] [Google Scholar]
  195. Schwarzl F.; Staverman A. J. Time-temperature Dependence of Linear Viscoelastic Behavior. J. Appl. Phys. 1952, 23 (8), 838–843. 10.1063/1.1702316. [DOI] [Google Scholar]
  196. Tobolsky A. V. Stress Relaxation Studies of the Viscoelastic Properties of Polymers. J. Appl. Phys. 1956, 27 (7), 673–685. 10.1063/1.1722465. [DOI] [Google Scholar]
  197. Kramer E. J. Molecular Theory of the Fracture-toughness of Low-molecular Weight Polymers. J. Mater. Sci. 1979, 14 (6), 1381–1388. 10.1007/BF00549312. [DOI] [Google Scholar]
  198. Kogan L.; Hui C. Y.; Ruina A. Theory of Chain Pull-out and Stability of Weak Polymer Interfaces.1. Macromolecules 1996, 29 (11), 4090–4100. 10.1021/ma950907v. [DOI] [Google Scholar]
  199. Sanchez-Valencia A.; Smerdova O.; Hutchings L. R.; De Focatiis D. S. A. The Roles of Blending and of Molecular Weight Distribution on Craze Initiation. Macromolecules 2017, 50 (23), 9507–9514. 10.1021/acs.macromol.7b01289. [DOI] [Google Scholar]
  200. Donald A. M.; Kramer E. J. Effect of Molecular Entanglements on Craze Microstructure in Glassy-polymers. J. Polym. Sci., Part B: Polym. Phys. 1982, 20 (5), 899–909. [Google Scholar]
  201. Liu J. W.; Yee A. F. Effect of Local Conformational Transition on Craze Initiation in Polyestercarbonates Containing Cyclohexylene Linkages. Macromolecules 2000, 33 (4), 1338–1344. 10.1021/ma981034i. [DOI] [Google Scholar]
  202. Bell J. P. Mechanical Properties of a Glassy Epoxide Polymer - Effect of Molecular Weight between Crosslinks. J. Appl. Polym. Sci. 1970, 14 (7), 1901–1906. 10.1002/app.1970.070140722. [DOI] [Google Scholar]
  203. Cho K.; Lee D.; Park C. E. Effect of Molecular Weight between Crosslinks on Fracture Behaviour of Diallylterephthalate Resins. Polymer 1996, 37 (5), 813–817. 10.1016/0032-3861(96)87258-2. [DOI] [Google Scholar]
  204. Yamini S.; Young R. J. The Mechanical Properties of Epoxy Resins 2. Effect of Plastic Deformation Upon Crack Propagation. J. Mater. Sci. 1980, 15 (7), 1823–1831. 10.1007/BF00550603. [DOI] [Google Scholar]
  205. Chuang Y. F.; Wu H. C.; Yang F. Q.; Yang T. J.; Lee S. Cracking and Healing in Poly (methyl methacrylate): Effect of Solvent. J. Polym. Res. 2016, 24 (2), 1–11. [Google Scholar]
  206. 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. Stab. 2016, 133, 255–263. 10.1016/j.polymdegradstab.2016.09.004. [DOI] [Google Scholar]
  207. Yee A. F.; Pearson R. A. Toughening Mechanisms in Elastomer-modified Epoxies.1. Mechanical Studies. J. Mater. Sci. 1986, 21 (7), 2462–2474. 10.1007/BF01114293. [DOI] [Google Scholar]
  208. Huang Y.; Kinloch A. J. The Role of Plastic Void Growth in the Fracture of Rubber-toughened Epoxy Polymers. J. Mater. Sci. Lett. 1992, 11 (8), 484–487. 10.1007/BF00731112. [DOI] [Google Scholar]
  209. Azimi H. R.; Pearson R. A.; Hertzberg R. W. Fatigue of Rubber-modified Epoxies: Effect of Particle Size and Volume Fraction. J. Mater. Sci. 1996, 31 (14), 3777–3789. 10.1007/BF00352793. [DOI] [Google Scholar]
  210. Hodgkin J. H.; Simon G. P.; Varley R. J. Thermoplastic Toughening of Epoxy Resins: A Critical Review. Polym. Adv. Technol. 1998, 9 (1), 3–10. . [DOI] [Google Scholar]
  211. Chandrasekaran V. C. S.; Advani S. G.; Santare M. H. Role Of Processing on Interlaminar Shear Strength Enhancement of Epoxy/Glass Fiber/Multi-walled Carbon Nanotube Hybrid Composites. Carbon 2010, 48 (13), 3692–3699. 10.1016/j.carbon.2010.06.010. [DOI] [Google Scholar]
  212. Zabihi O.; Ahmadi M.; Nikafshar S.; Preyeswary K. C.; Naebe M. A Technical Review on Epoxy-clay Nanocomposites: Structure, Properties, and Their Applications in Fiber Reinforced Composites. Composites, Part B 2018, 135, 1–24. 10.1016/j.compositesb.2017.09.066. [DOI] [Google Scholar]
  213. El-Fattah M. A.; El Saeed A. M.; El-Ghazawy R. A. Chemical Interaction of Different Sized Fumed Silica with Epoxy Via Ultrasonication for Improved Coating. Prog. Org. Coat. 2019, 129, 1–9. 10.1016/j.porgcoat.2018.12.023. [DOI] [Google Scholar]
  214. Sasidharan S.; Anand A. Epoxy-based Hybrid Structural Composites with Nanofillers: A Review. Ind. Eng. Chem. Res. 2020, 59 (28), 12617–12631. 10.1021/acs.iecr.0c01711. [DOI] [Google Scholar]
  215. Shundo A.; Sakurai T.; Takafuji M.; Nagaoka S.; Ihara H. Molecular-length and Chiral Discriminations by β-Structural Poly (L-alanine) on Silica. J. Chromatogr. A 2005, 1073 (1–2), 169–174. 10.1016/j.chroma.2004.08.062. [DOI] [PubMed] [Google Scholar]
  216. Bagwe R. P.; Hilliard L. R.; Tan W. H. Surface Modification of Silica Nanoparticles to Reduce Aggregation and Nonspecific Binding. Langmuir 2006, 22 (9), 4357–4362. 10.1021/la052797j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Shundo A.; Mallik A. K.; Nakashima H.; Takafuji M.; Nagaoka S.; Chowdhury M. A. J.; Ihara H. Enhancement of Discrimination Ability for Cis- and Trans-decalins through Side-chain Ordering in Comb-shaped Polymer. Chem. Lett. 2010, 39 (8), 844–845. 10.1246/cl.2010.844. [DOI] [Google Scholar]
  218. Sprenger S. Epoxy Resin Composites with Surface-modified Silicon Dioxide Nanoparticles: A Review. J. Appl. Polym. Sci. 2013, 130 (3), 1421–1428. 10.1002/app.39208. [DOI] [Google Scholar]
  219. Nakamura Y.; Yamaguchi M.; Okubo M.; Matsumoto T. Effect of Particle Size on the Fracture Toughness of Epoxy Resin Filled with Spherical Silica. Polymer 1992, 33 (16), 3415–3426. 10.1016/0032-3861(92)91099-N. [DOI] [Google Scholar]
  220. Johnsen B. B.; Kinloch A. J.; Mohammed R. D.; Taylor A. C.; Sprenger S. Toughening Mechanisms of Nanoparticle-modified Epoxy Polymers. Polymer 2007, 48 (2), 530–541. 10.1016/j.polymer.2006.11.038. [DOI] [Google Scholar]
  221. Liu H. Y.; Wang G. T.; Mai Y. W.; Zeng Y. on Fracture Toughness of Nano-particle Modified Epoxy. Composites, Part B 2011, 42 (8), 2170–2175. 10.1016/j.compositesb.2011.05.014. [DOI] [Google Scholar]
  222. Dittanet P.; Pearson R. A. Effect of Silica Nanoparticle Size on Toughening Mechanisms of Filled Epoxy. Polymer 2012, 53 (9), 1890–1905. 10.1016/j.polymer.2012.02.052. [DOI] [Google Scholar]
  223. Lin B.; Zhao L. G.; Tong J. A Crystal Plasticity Study of Cyclic Constitutive Behaviour, Crack-tip Deformation and Crack-growth Path for a Polycrystalline Nickel-based Superalloy. Eng. Fract. Mech. 2011, 78 (10), 2174–2192. 10.1016/j.engfracmech.2011.04.006. [DOI] [Google Scholar]
  224. Liao Y. S.; Li Y. B.; Huang M. H.; Wang B. S.; Yang Y. C.; Pei S. K. Effect of Hole Relative Size and Position on Crack Deflection Angle of Repaired Structure. Theor. Appl. Fract. Mech. 2019, 101, 92–102. 10.1016/j.tafmec.2019.02.010. [DOI] [Google Scholar]
  225. Wang M.; Zhao L.; Fourmeau M.; Nelias D. Crack Plane Deflection and Shear Wave Effects in the Dynamic Fracture of Silicon Single Crystal. J. Mech. Phys. Solids. 2019, 122, 472–488. 10.1016/j.jmps.2018.09.031. [DOI] [Google Scholar]
  226. Sassoni E.; Andreotti S.; Bellini A.; Mazzanti B.; Bignozzi M. C.; Mazzotti C.; Franzoni E. Influence of Mechanical Properties, Anisotropy, Surface Roughness and Porosity of Brick on FRP Debonding Force. Composites, Part B 2017, 108, 257–269. 10.1016/j.compositesb.2016.10.020. [DOI] [Google Scholar]
  227. Kabel J.; Yang Y.; Balooch M.; Howard C.; Koyanagi T.; Terrani K. A.; Katoh Y.; Hosemann P. Micro-mechanical Evaluation of SiC-SiC Composite Interphase Properties and Debond Mechanisms. Composites, Part B 2017, 131, 173–183. 10.1016/j.compositesb.2017.07.035. [DOI] [Google Scholar]
  228. Wang P. P.; Maeda R.; Aoki M.; Kubozono T.; Yoshihara D.; Shundo A.; Kobayashi T.; Yamamoto S.; Tanaka K.; Yamada S. In Situ Transmission Electron Microscopy Observation of the Deformation and Fracture Processes of an Epoxy/Silica Nanocomposite. Soft Matter 2022, 18 (6), 1149–1153. 10.1039/D1SM01452H. [DOI] [PubMed] [Google Scholar]
  229. Soni N. J.; Lin P. H.; Khare R. Effect of Cross-linker Length on the Thermal and Volumetric Properties of Cross-linked Epoxy Networks: A Molecular Simulation Study. Polymer 2012, 53 (4), 1015–1019. 10.1016/j.polymer.2011.12.051. [DOI] [Google Scholar]
  230. Mijovic J.; Zhang H. Local Dynamics and Molecular Origin of Polymer Network-water Interactions as Studied by Broadband Dielectric Relaxation Spectroscopy, FTIR, and Molecular Simulations. Macromolecules 2003, 36 (4), 1279–1288. 10.1021/ma021568q. [DOI] [Google Scholar]
  231. Shenogina N. B.; Tsige M.; Patnaik S. S.; Mukhopadhyay S. M. Molecular Modeling Approach to Prediction of Thermo-mechanical Behavior of Thermoset Polymer Networks. Macromolecules 2012, 45 (12), 5307–5315. 10.1021/ma3007587. [DOI] [Google Scholar]
  232. Okabe T.; Oya Y.; Tanabe K.; Kikugawa G.; Yoshioka K. Molecular Dynamics Simulation of Crosslinked Epoxy Resins: Curing and Mechanical Properties. Eur. Polym. J.. 2016, 80, 78–88. 10.1016/j.eurpolymj.2016.04.019. [DOI] [Google Scholar]
  233. Odegard G. M.; Patil S. U.; Deshpande P. P.; Kanhaiya K.; Winetrout J. J.; Heinz H.; Shah S. P.; Maiaru M. Molecular Dynamics Modeling of Epoxy Resins Using the Reactive Interface Force Field. Macromolecules 2021, 54 (21), 9815–9824. 10.1021/acs.macromol.1c01813. [DOI] [Google Scholar]
  234. Andrews E. H.; Kinloch A. J. Mechanics of Adhesive Failure. II. Proc. R. Soc. London A 1973, 332 (1590), 401–414. 10.1098/rspa.1973.0033. [DOI] [Google Scholar]
  235. Bahlakeh G.; Ramezanzadeh B. A Detailed Molecular Dynamics Simulation and Experimental Investigation on the Interfacial Bonding Mechanism of an Epoxy Adhesive on Carbon Steel Sheets Decorated with a Novel Cerium-lanthanum Nanofilm. ACS Appl. Mater. Interfaces 2017, 9 (20), 17536–17551. 10.1021/acsami.7b00644. [DOI] [PubMed] [Google Scholar]
  236. Okamoto K.; Ganbe T.; Sekine N.; Aoki M.; Inutsuka M.; Shundo A.; Kawaguchi D.; Tanaka K. Nanoscale Characterization of Epoxy Interface on Silica. IEEE Int. Conf. Dielectrics 2016, 84–87. [Google Scholar]
  237. Munz M.; Sturm H.; Schulz E. Interphase Characterization via the Stiffness Contrast of a Scanning Force Microscope in Displacement Modulation Mode. Surf. Interface Anal. 2000, 30 (1), 410–414. . [DOI] [Google Scholar]
  238. Munz M.; Sturm H.; Stark W. Mechanical Gradient Interphase by Interdiffusion and Antiplasticisation Effect - Study of an Epoxy/Thermoplastic System. Polymer 2005, 46 (21), 9097–9112. 10.1016/j.polymer.2005.06.098. [DOI] [Google Scholar]
  239. Turunen M. P. K.; Marjamaki P.; Paajanen M.; Lahtinen J.; Kivilahti J. K. Pull-off Test in the Assessment of Adhesion at Printed Wiring Board Metallisation/Epoxy Interface. Microelectron. Reliab. 2004, 44 (6), 993–1007. 10.1016/j.microrel.2004.01.001. [DOI] [Google Scholar]
  240. Watts J. F.; Castle J. E.; Ludlam S. J. The Orientation of Molecules at the Locus of Failure of Polymer-coatings on Steel. J. Mater. Sci. 1986, 21 (8), 2965–2971. 10.1007/BF00551517. [DOI] [Google Scholar]
  241. Bockenheimer C.; Valeske B.; Possart W. Network Structure in Epoxy Aluminium Bonds after Mechanical Treatment. Int. J. Adhes. Adhes. 2002, 22 (5), 349–356. 10.1016/S0143-7496(02)00014-3. [DOI] [Google Scholar]
  242. Onard S.; Martin I.; Chailan J. F.; Crespy A.; Carriere P. Nanostructuration in Thin Epoxy-amine Films Inducing Controlled Specific Phase Etherification: Effect on the Glass Transition Temperatures. Macromolecules 2011, 44 (9), 3485–3493. 10.1021/ma102002j. [DOI] [Google Scholar]
  243. Chung J.; Munz M.; Sturm H. Stiffness Variation in the Interphase of Amine-cured Epoxy Adjacent to Copper Microstructures. Surf. Interface Anal. 2007, 39 (7), 624–633. 10.1002/sia.2571. [DOI] [Google Scholar]
  244. Aufray M.; Roche A.. Properties of the Interphase Epoxy-amine/Metal: Influences from the Nature of the Amine and the Metal. In Adhesion-Current Research and Application; Possart W., Ed.; Wiley-VCH, 2005; pp 89–102. [Google Scholar]
  245. Marsh J.; Minel L.; Barthes-Labrousse M. G.; Gorse D. Interaction of Epoxy Model Molecules with Aluminium, Anodised Titanium and Copper Surfaces: An XPS Study. Appl. Surf. Sci. 1998, 133 (4), 270–286. 10.1016/S0169-4332(97)00789-7. [DOI] [Google Scholar]
  246. Roche A. A.; Bouchet J.; Bentadjine S. Formation of Epoxy-diamine/Metal Interphases. Int. J. Adhes. Adhes. 2002, 22 (6), 431–441. 10.1016/S0143-7496(02)00021-0. [DOI] [Google Scholar]
  247. Possart W.; Kruger J. K.; Wehlack C.; Muller U.; Petersen C.; Bactavatchalou R.; Meiser A. Formation and Structure of Epoxy Network Interphases at the Contact to Native Metal Surfaces. CR Chim. 2006, 9 (1), 60–79. 10.1016/j.crci.2005.04.009. [DOI] [Google Scholar]
  248. Deng S. Q.; Djukic L.; Paton R.; Ye L. Thermoplastic-epoxy Interactions and Their Potential Applications in Joining Composite Structures - A Review. Composites, Part A 2015, 68, 121–132. 10.1016/j.compositesa.2014.09.027. [DOI] [Google Scholar]
  249. Hirai T.; Kawasaki K.; Tanaka K. Interfacial Kinetics of a Model Epoxy-amine Addition Reaction. Phys. Chem. Chem. Phys. 2012, 14 (39), 13532–13534. 10.1039/c2cp42673k. [DOI] [PubMed] [Google Scholar]
  250. Bentadjine S.; Petiaud R.; Roche A. A.; Massardier V. Organo-metallic Complex Characterization Formed When Liquid Epoxy-diamine Mixtures Are Applied onto Metallic Substrates. Polymer 2001, 42 (14), 6271–6282. 10.1016/S0032-3861(01)00034-9. [DOI] [Google Scholar]
  251. Karasinski E. N.; Da Luz M. G.; Lepienski C. M.; Coelho L. A. F. Nanostructured Coating Based on Epoxy/Metal Oxides: Kinetic Curing and Mechanical Properties. Thermochim. Acta 2013, 569, 167–176. 10.1016/j.tca.2013.07.015. [DOI] [Google Scholar]
  252. Meiser A.; Kubel C.; Schafer H.; Possart W. Electron Microscopic Studies on the Diffusion of Metal Ions in Epoxy-metal Interphases. Int. J. Adhes. Adhes. 2010, 30 (3), 170–177. 10.1016/j.ijadhadh.2009.12.001. [DOI] [Google Scholar]
  253. Liu Y.; Hamon A. L.; Haghi-Ashtiani P.; Reiss T.; Fan B. H.; He D. L.; Baio J. B. Quantitative Study of Interface/Interphase in Epoxy/Graphene-based Nanocomposites by Combining STEM and EELS. ACS Appl. Mater. Interfaces 2016, 8 (49), 34151–34158. 10.1021/acsami.6b12915. [DOI] [PubMed] [Google Scholar]
  254. Aoki M.; Shundo A.; Okamoto K.; Ganbe T.; Tanaka K. Segregation of an Amine Component in a Model Epoxy Resin at a Copper Interface. Polym. J. 2019, 51 (3), 359–363. 10.1038/s41428-018-0129-4. [DOI] [Google Scholar]
  255. Yamamoto S.; Kuwahara R.; Aoki M.; Shundo A.; Tanaka K. Molecular Events for an Epoxy-amine System at a Copper Interface. ACS Appl. Polym. Mater. 2020, 2 (4), 1474–1481. 10.1021/acsapm.9b01154. [DOI] [Google Scholar]
  256. Yamamoto S.; Tanaka K. Entropy-driven Segregation in Epoxy-amine Systems at a Copper Interface. Soft Matter 2021, 17 (5), 1359–1367. 10.1039/D0SM01600D. [DOI] [PubMed] [Google Scholar]
  257. Unnikrishnan K. P.; Thachil E. T. Toughening of Epoxy Resins. Des. Monomers Polym. 2006, 9 (2), 129–152. 10.1163/156855506776382664. [DOI] [Google Scholar]
  258. Wang X. S.; Kim H. K.; Fujita Y.; Sudo A.; Nishida H.; Endo T. Relaxation and Reinforcing Effects of Polyrotaxane in an Epoxy Resin Matrix. Macromolecules 2006, 39 (3), 1046–1052. 10.1021/ma051888k. [DOI] [Google Scholar]
  259. Pruksawan S.; Samitsu S.; Yokoyama H.; Naito M. Homogeneously Dispersed Polyrotaxane in Epoxy Adhesive and Its Improvement in the Fracture Toughness. Macromolecules 2019, 52 (6), 2464–2475. 10.1021/acs.macromol.8b02450. [DOI] [Google Scholar]
  260. Hanafusa A.; Ando S.; Ozawa S.; Ito M.; Hasegawa R.; Mayumi K.; Ito K. Viscoelastic Relaxation Attributed to the Molecular Dynamics of Polyrotaxane Confined iin an Epoxy Resin Network. Polymer J. 2020, 52 (10), 1211–1221. 10.1038/s41428-020-0373-2. [DOI] [Google Scholar]
  261. Tano K.; Sato E. Synthesis and Dissociation Behavior of Degradable Network Polymers Consisting of Epoxides and 9-Anthracene Carboxylic Acid Dimer. Chem. Lett. 2021, 50 (10), 1787–1790. 10.1246/cl.210332. [DOI] [Google Scholar]
  262. Takahashi A.; Ohishi T.; Goseki R.; Otsuka H. Degradable Epoxy Resins Prepared from Diepoxide Monomer with Dynamic Covalent Disulfide Linkage. Polymer 2016, 82 (1), 319–326. 10.1016/j.polymer.2015.11.057. [DOI] [Google Scholar]
  263. Ruiz de Luzuriaga A.; Matxain J. M.; Ruiperez F.; Martin R.; Asua J. M.; Cabanero G.; Odriozola I. Transient Mechanochromism in Epoxy Vitrimer Composites Containing Aromatic Disulfide Crosslinks. J. Mater. Chem. C 2016, 4 (26), 6220–6223. 10.1039/C6TC02383E. [DOI] [Google Scholar]
  264. Mo R. B.; Hu J.; Huang H. W.; Sheng X. X.; Zhang X. Y. Tunable, Self-healing and Corrosion Inhibiting Dynamic Epoxy-polyimine Network Built by Post-crosslinking. J. Mater. Chem. A 2019, 7 (7), 3031–3038. 10.1039/C8TA11546J. [DOI] [Google Scholar]
  265. Montarnal D.; Capelot M.; Tournilhac F.; Leibler L. Silica-like Malleable Materials from Permanent Organic Networks. Science 2011, 334 (6058), 965–968. 10.1126/science.1212648. [DOI] [PubMed] [Google Scholar]
  266. Capelot M.; Unterlass M. M.; Tournilhac F.; Leibler L. Catalytic Control of the Vitrimer Glass Transition. ACS Macro Lett. 2012, 1 (7), 789–792. 10.1021/mz300239f. [DOI] [PubMed] [Google Scholar]
  267. Chen M.; Zhou L.; Wu Y. P.; Zhao X. L.; Zhang Y. J. Rapid Stress Relaxation and Moderate Temperature of Malleability Enabled by the Synergy of Disulfide Metathesis and Carboxylate Transesterification in Epoxy Vitrimers. ACS Macro Lett. 2019, 8 (3), 255–260. 10.1021/acsmacrolett.9b00015. [DOI] [PubMed] [Google Scholar]
  268. Yang Y.; Xu Y.; Ji Y.; Wei Y. Functional Epoxy Vitrimers and Composites. Prog. Mater. Sci. 2021, 120 (6), 100710. 10.1016/j.pmatsci.2020.100710. [DOI] [Google Scholar]
  269. Yang Z.; Wang Q.; Wang T. Dual-triggered and Thermally Reconfigurable Shape Memory Graphene-vitrimer Composites. ACS Appl. Mater. Interfaces 2016, 8 (33), 21691–21699. 10.1021/acsami.6b07403. [DOI] [PubMed] [Google Scholar]
  270. Yang Y.; Pei Z.; Zhang X.; Tao L.; Wei Y.; Ji Y. Carbon Nanotube-vitrimer Composite for Facile and Efficient Photo-welding of Epoxy. Chem. Sci. 2014, 5 (9), 3486–3492. 10.1039/C4SC00543K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Sun Lee W.; Yu J. Comparative Study of Thermally Conductive Fillers in Underfill for the Electronic Components. Diamond Relat. Mater. 2005, 14 (10), 1647–1653. 10.1016/j.diamond.2005.05.008. [DOI] [Google Scholar]
  272. Choi S.; Kim J. Thermal Conductivity of Epoxy Composites with a Binary-particle System of Aluminum Oxide and Aluminum Nitride Fillers. Composites, Part B 2013, 51, 140–147. 10.1016/j.compositesb.2013.03.002. [DOI] [Google Scholar]
  273. Shi Z. Q.; Radwan M.; Kirihara S.; Miyamoto Y.; Jin Z. H. Enhanced Thermal Conductivity of Polymer Composites Filled with Three-dimensional Brushlike ALN Nanowhiskers. Appl. Phys. Lett. 2009, 95 (22), 224104. 10.1063/1.3271028. [DOI] [Google Scholar]
  274. Li T. L.; Hsu S. L. C. Enhanced Thermal Conductivity of Polyimide Films via a Hybrid of Micro- and Nano-sized Boron Nitride. J. Phys. Chem. B 2010, 114 (20), 6825–6829. 10.1021/jp101857w. [DOI] [PubMed] [Google Scholar]
  275. Wang Q.; Chen X.; Huang X.; Muhammad A.; Shi Y.; Dai C.; Ren N.; Paramane A. Tailoring Electric Field Distortion in High-voltage Power Modules Utilizing Epoxy Resin/Silicon Carbide Whisker Composites with Field-dependent Conductivity. ACS Appl. Electron. Mater. 2022, 4 (1), 478–493. 10.1021/acsaelm.1c01099. [DOI] [Google Scholar]
  276. Cahill D. G.; Watson S. K.; Pohl R. O. Lower Limit to the Thermal Conductivity of Disordered Crystals. Phys. Rev. B 1992, 46 (10), 6131–6140. 10.1103/PhysRevB.46.6131. [DOI] [PubMed] [Google Scholar]
  277. Xu Y.; Wang X.; Zhou J.; Song B.; Jiang Z.; Lee E. M. Y.; Huberman S.; Gleason K. K.; Chen G. Molecular Engineered Conjugated Polymer with High Thermal Conductivity. Sci. Adv. 2018, 4 (3), eaar3031 10.1126/sciadv.aar3031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  278. Maeda R.; Okuhara K.; Nakamura A.; Hayakawa T.; Uehara Y.; Motoya T.; Nobutoki H. Higher-order-structure Formation in Liquid Crystal Epoxy Thermosets Investigated by Synchrotron Radiation Wide-angle X-ray Diffraction. Chem. Lett. 2016, 45 (7), 795–797. 10.1246/cl.160249. [DOI] [Google Scholar]
  279. Lv G.; Jensen E.; Evans C. M.; Cahill D. G. High Thermal Conductivity Semicrystalline Epoxy Resins with Anthraquinone-based Hardners. ACS Appl. Polym. Mater. 2021, 3 (9), 4430–4435. 10.1021/acsapm.1c00737. [DOI] [Google Scholar]
  280. Mo H. L.; Huang X. Y.; Liu F.; Yang K.; Li S. T.; Jiang P. K. Nanostructured Electrical Insulating Epoxy Thermosets with High Thermal Conductivity, High Thermal Stability, High Glass Transition Temperatures and Excellent Dielectric Properties. IEEE Trans. Dielectr. Electr. Insul. 2015, 22 (2), 906–915. 10.1109/TDEI.2015.7076791. [DOI] [Google Scholar]
  281. Song S. H.; Katagi H.; Takezawa Y. Study on High Thermal Conductivity of Mesogenic Epoxy Resin with Spherulite Structure. Polymer 2012, 53 (20), 4489–4492. 10.1016/j.polymer.2012.07.065. [DOI] [Google Scholar]
  282. Harada M.; Ochi M.; Tobita M.; Kimura T.; Ishigaki T.; Shimoyama N.; Aoki H. Thermal-conductivity Properties of Liquid-crystalline Epoxy Resin Cured under a Magnetic Field. J. Polym. Sci., Part B: Polym. Phys. 2003, 41 (14), 1739–1743. 10.1002/polb.10531. [DOI] [Google Scholar]
  283. Cheng W. T.; Chih Y. W.; Yeh W. T. In Situ Fabrication of Photocurable Conductive Adhesives with Silver Nano-particles in the Absence of Capping Agent. Int. J. Adhes. Adhes. 2007, 27 (3), 236–243. 10.1016/j.ijadhadh.2006.05.001. [DOI] [Google Scholar]
  284. Martin-Gallego M.; Lopez-Manchado M. A.; Calza P.; Roppolo I.; Sangermano M. Gold-functionalized Graphene as Conductive Filler in UV-curable Epoxy Resin. J. Mater. Sci. 2015, 50 (2), 605–610. 10.1007/s10853-014-8619-z. [DOI] [Google Scholar]
  285. Lee H. H.; Chou K. S.; Shih Z. W. Effect of Nano-sized Silver Particles on the Resistivity of Polymeric Conductive Adhesives. Int. J. Adhes. Adhes. 2005, 25 (5), 437–441. 10.1016/j.ijadhadh.2004.11.008. [DOI] [Google Scholar]
  286. Yim M. J.; Paik K. W. Recent Advances on Anisotropic Conductive Adhesives (ACAs) for Flat Panel Displays and Semiconductor Packaging Applications. Int. J. Adhes. Adhes. 2006, 26 (5), 304–313. 10.1016/j.ijadhadh.2005.04.004. [DOI] [Google Scholar]
  287. Mir I.; Kumar D. Recent Advances in Isotropic Conductive Adhesives for Electronics Packaging Applications. Int. J. Adhes. Adhes. 2008, 28 (7), 362–371. 10.1016/j.ijadhadh.2007.10.004. [DOI] [Google Scholar]
  288. Li Y.; Wong C. P. Recent Advances of Conductive Adhesives as a Lead-free Alternative in Electronic Packaging: Materials, Processing, Reliability and Applications. Mater. Sci. Eng. R Rep. 2006, 51 (1–3), 1–35. 10.1016/j.mser.2006.01.001. [DOI] [Google Scholar]
  289. Wu H. P.; Liu J. F.; Wu X. J.; Ge M. Y.; Wang Y. W.; Zhang G. Q.; Jiang J. Z. High Conductivity of Isotropic Conductive Adhesives Filled with Silver Nanowires. Int. J. Adhes. Adhes. 2006, 26 (8), 617–621. 10.1016/j.ijadhadh.2005.10.001. [DOI] [Google Scholar]
  290. Jiang H. J.; Yim M. J.; Lin W.; Wong C. P. Novel Nonconductive Adhesives/Films with Carbon Nanotubes for High-performance Interconnects. IEEE Trans. Compon. Packag. Technol. 2009, 32 (4), 754–758. 10.1109/TCAPT.2009.2014742. [DOI] [Google Scholar]
  291. Massoumi B.; Hosseinzadeh M.; Jaymand M. Electrically Conductive Nanocomposite Adhesives Based on Epoxy or Chloroprene Containing Polyaniline, and Carbon Nanotubes. J. Mater. Sci.: Mater. Electron. 2015, 26 (8), 6057–6067. [Google Scholar]
  292. Ng F. F.; Couture G.; Philippe C.; Boutevin B.; Caillol S. Bio-based Aromatic Epoxy Monomers for Thermoset Materials. Molecules 2017, 22 (1), 149. 10.3390/molecules22010149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  293. Wang X.; Guo W. W.; Song L.; Hu Y. Intrinsically Flame Retardant Bio-based Epoxy Thermosets: A Review. Composites, Part B 2019, 179, 107487. 10.1016/j.compositesb.2019.107487. [DOI] [Google Scholar]
  294. Auvergne R.; Caillol S.; David G.; Boutevin B.; Pascault J. P. Biobased Thermosetting Epoxy: Present and Future. Chem. Rev. 2014, 114 (2), 1082–1115. 10.1021/cr3001274. [DOI] [PubMed] [Google Scholar]
  295. Baroncini E. A.; Yadav S. K.; Palmese G. R.; Stanzione J. F. Recent Advances in Bio-based Epoxy Resins and Bio-based Epoxy Curing Agents. J. Appl. Polym. Sci. 2016, 133 (45), 44103. 10.1002/app.44103. [DOI] [Google Scholar]
  296. Yue L.; Liu F.; Mekala S.; Patel A.; Gross R. A.; Manas-Zloczower I. High Performance Biobased Epoxy Nanocomposite Reinforced with a Bacterial Cellulose Nanofiber Network. ACS Sustainable Chem. Eng. 2019, 7 (6), 5986–5992. 10.1021/acssuschemeng.8b06073. [DOI] [Google Scholar]
  297. Wang Z.; Gnanasekar P.; Sudhakaran Nair S.; Farnood R.; Yi S.; Yan N. Biobased Epoxy Synthesized from a Vanillin Derivative and Its Reinforcement Using Lignin-containing Cellulose Nanofibrils. ACS Sustainable Chem. Eng. 2020, 8 (30), 11215–11223. 10.1021/acssuschemeng.0c02559. [DOI] [Google Scholar]

Articles from JACS Au are provided here courtesy of American Chemical Society

RESOURCES