Next Article in Journal
Effects of Organic Acidic Products from Discharge-Induced Decomposition of the FRP Matrix on ECR Glass Fibers in Composite Insulators
Previous Article in Journal
An Electrospun DFO-Loaded Microsphere/SAIB System Orchestrates Angiogenesis–Osteogenesis Coupling via HIF-1α Activation for Vascularized Bone Regeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Study of the Dielectric Relaxation of Nitrile–Butadiene Rubber, Ethylene–Propylene–Diene Monomer, and Fluoroelastomer Polymers with a Self-Developed Deconvolution Analysis Program

1
Department of Electrical Engineering, Pohang University of Science and Technology, Pohang 37673, Gyeonsangbuk-do, Republic of Korea
2
Advanced Institute of Convergence Technology, Suwon 16229, Gyeonggi-do, Republic of Korea
3
Hydrogen Energy Materials Research Center, Korea Research Institute of Standards and Science, Daejeon 34113, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Polymers 2025, 17(11), 1539; https://doi.org/10.3390/polym17111539
Submission received: 19 March 2025 / Revised: 9 May 2025 / Accepted: 22 May 2025 / Published: 31 May 2025
(This article belongs to the Section Polymer Physics and Theory)

Abstract

:
This study presents an integrated analysis of the dielectric characteristics of nitrile–butadiene rubber (NBR), ethylene–propylene–diene monomer (EPDM), and fluoroelastomer (FKM) polymers. Dispersion spectra were obtained over a wide range of frequencies and temperatures, and, via our self-developed “Dispersion Analysis” program, the obtained dielectric spectra were precisely deconvoluted. Notably, α, α’, β, and γ relaxation phenomena, including the DC conduction process, were identified in NBR, whereas three relaxation processes, namely, α, β, and the Maxwell‒Wagner‒Sillars (MWS) process, as well as DC conduction, were observed in EPDM and FKM copolymers. The activation energies ( E a ) for secondary relaxation—namely, β, γ, and MWS—and the DC conduction process, which are observed in NBR, EPDM, and FKM, were determined via the Arrhenius temperature dependence model, and these values were compared with previously published results. Furthermore, the glass transition temperature ( T g ), extrapolated from the relaxation rate of the α process, was estimated via the Vogel–Fulcher–Tamman–Hesse (VFTH) law. The values of T g obtained using dielectric spectroscopy for NBR, EPDM, and FKM agreed well with the differential scanning calorimetry (DSC) measurements. This study provides foundational insights into the dielectric properties of widely used rubber polymers, offering a comprehensive reference for future research.

1. Introduction

Owing to their outstanding physical, chemical, and mechanical features, polymeric rubber compounds have been widely utilized in various industrial applications, such as O-ring gaskets, pipelines and valve sealants, electrical insulation tubes, and automotive and aerospace components [1,2,3,4,5]. Recently, rubber compounds have played a crucial role in hydrogen electric vehicle (HEV) applications by increasing the load-bearing capacity, reducing rolling resistance, and improving durability [6,7,8,9]. Furthermore, low chemical reactivity and minimal outgassing properties are essential for hydrogen fuel cells and high-vacuum applications (down to approximately 1−9 Torr) [10,11,12,13,14,15,16,17,18,19,20]. Thus, extensive research has been conducted to develop advanced rubber composites, and various experimental techniques have been employed to evaluate their physical and mechanical characteristics [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67]. Representative methods—such as tensile testing, dynamic mechanical analysis (DMA), transport analysis, thermogravimetric analysis (TGA), Fourier-transform infrared spectroscopy (FTIR), nuclear magnetic resonance (NMR), and energy-dispersive X-ray spectroscopy (EDS)—have provided valuable insights into the structural and functional attributes of these materials [68,69,70,71,72,73,74,75,76,77,78,79]. Among these techniques, dielectric spectroscopy stands out as a nondestructive and effective method for evaluating the electrical characteristics of rubber composites as it enables precise measurements to be made of polarization dynamics and charge carrier transport phenomena across a broad frequency range of alternating electric fields.
Debye, who first illuminated the molecular aspects of dielectric behavior, established the foundational concepts and models underlying dielectric spectroscopy [80,81,82]. This technique offers unique advantages, including the ability to probe molecular dynamics, charge carrier distributions, and determine a material’s response to external electric fields ranging from direct current (DC) to approximately 1012 Hz [83,84,85,86,87,88,89,90,91,92,93,94,95]. It is particularly useful for understanding the behavior of polarized molecular structures under various environmental conditions, including temperature and humidity variations [96,97,98,99,100,101,102,103,104,105,106,107,108]. Consequently, numerous studies have employed dielectric spectroscopy to investigate the dielectric properties of various polymeric rubbers, such as styrene−butadiene rubber (SBR), nitrile–butadiene rubber (NBR), ethylene–propylene–diene monomer (EPDM), and fluoroelastomer (FKM) [74,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167].
In our previous work, we reported comprehensive studies of NBR, EPDM, and FKM rubbers via dielectric spectroscopy over a wide range of frequencies and temperatures [33,168,169,170,171,172,173]. In the present study, we review the dielectric analysis results of these rubber composites via a self-developed deconvolution analysis program. By deconvoluting the obtained dielectric spectra with a combination of the Havriliak–Negami model and an exponential power law for DC conduction, we identified various relaxation phenomena, including the DC conduction process. The origins of each relaxation were determined on the basis of their temperature-dependent relaxation times and the corresponding dipole orientation dynamics of specific molecular structures. Additionally, the activation energies for the relaxation processes and DC conduction, both of which follow the Arrhenius temperature dependence law, were obtained and compared with previously reported values. Moreover, the glass transition temperature ( T g ) was calculated from the α relaxation, which follows the Vogel–Fulcher–Tammann–Hesse (VFTH) temperature dependence. The obtained T g value was then compared with the value obtained via differential scanning calorimetry (DSC). Furthermore, the basic algorithm and outlook of the self-developed dielectric deconvolution program based on C# were introduced. The entire code of this developed deconvolution software is available in the Supplementary Materials. This study provides fundamental insights into the dielectric properties of widely used rubber polymers and offers a comprehensive reference for future research.

2. Relaxation Process and Model Analysis

Although numerous relaxation and transition dynamics have been studied across various types of materials [93,174,175,176,177,178,179,180,181,182,183,184,185,186,187], we briefly discuss the relaxation processes limited to synthetic rubber polymers. The internal structure of these polymers primarily comprises repeated monomer chains, filling, and vulcanizing agents [188,189,190]. In the millimeter to microwave frequency range, the dielectric response of rubber polymers is mainly attributed to the dipole orientation of the backbone and side chains as well as interfacial polarization processes [191,192,193,194,195,196,197,198,199,200]. Each of these relaxation processes can be characterized by evaluating the shape of the dispersion spectra and their temperature dependence. The underlying mechanisms of these dielectric relaxation processes, along with the corresponding analytical functions, are given below.

2.1. α Relaxation

The α relaxation process represents the dominant relaxation behavior in amorphous polymers, originating from the reorientation of segmental (backbone) chain motions [199,200,201]. It is directly associated with the polymer’s glass transition dynamics and is thus often referred to as the dynamic glass transition. The relaxation rate ( f p , α ) of α relaxation obeys the non-Arrhenius temperature dependence described by the VFTH temperature dependence model:
ln f p , α = ln f p , A T T 0
where f p , and A are constants; T 0 is the ideal glass transition temperature or Vogel temperature; and the static T g , determined using dielectric spectroscopy, corresponds to the temperature at which the curve of the VFTH temperature behavior is attained at approximately 1.6 mHz [199]. This frequency can vary depending on the thermal protocol. Meanwhile, the dynamic T g is identified as the temperature showing the peak in dielectric loss at a specified measurement frequency.

2.2. α′ Relaxation (Normal Mode Relaxation)

The α′ relaxation process is associated with fluctuations in the end-to-end vector of polymer chains and is characteristic of type A polymers. According to Stockmayer [202] and Block [203], the vector sum of the total dipole moment in polymer chains comprises the parallel direction with the segmental chain (backbone chain) called polymer type A. α′ relaxation, called normal mode relaxation, can be observed when it is proportional to the fluctuation of the end-to-end vector of the polymer chain. Like α relaxation, α′ relaxation has a central frequency with a nonlinear temperature dependence. The relaxation rate of α′ can also be described by VFTH temperature dependence.

2.3. β and γ Relaxations (Secondary Relaxation)

β and γ relaxations are a type of secondary relaxation resulting from the fluctuating motion of the side groups attached to polymer chains [199,200,201]. In addition, secondary relaxations may also originate from isolated molecules or impurities, which exhibit high relaxation rates. The activation energy ( E a ) for these relaxations is determined by assuming an Arrhenius temperature dependence:
f p = f exp E a k B T
where f p is the position of the center peak of the secondary relaxation rate, f is the pre-exponential factor, and E a is related to the slope of log( f p ) versus 1/T and depends on the internal rotation barriers and the environment of a moving unit.

2.4. Interfacial Relaxation (MWS Relaxation)

Interfacial relaxation, called Maxwell‒Wagner–Sillars (MWS) relaxation, originates from polarization processes between polymer chains and filler contents or other elements in a heterogeneous polymeric system [199,200,201]. These relaxation phenomena can only be detected via dielectric spectroscopy. Thus, dielectric measurements can provide valuable information on the morphology of a phase-separated polymer system. Similarly to secondary relaxation, interfacial relaxation has a relaxation rate on interfacial polarization with an Arrhenius temperature dependence determined via Equation (2).

2.5. DC Conduction Process

The contribution of the DC conduction process to dielectric spectra typically appears in the low-frequency (a few millihertz to kilohertz) and high-temperature (above 280~300 K) regimes of the dielectric loss spectrum. The origin of this process in polymers can be regarded as a transport phenomenon by considering the movement of charge carriers such as ionic impurities [204] and thermally excited electrons between the valence and conduction bands [205]. The complex dielectric constant ε * ω is written as [199,200,201]
ε * ω = ε ω ε ( ω ) = ε ω i σ ( ω ) ε 0 ω
where σ ( ω ) is the total conductivity, and ε 0 is the vacuum permittivity of 8.854 × 10−12 F/m. The measured dielectric loss σ ( ω ) ε 0 ω is composed of two components, namely, one due to dielectric loss ε″(ω) and one due to DC electrical conductivity σ d c , which is the low-frequency limit of σ(ω). Thus, the general expression of Equation (3) is written as
ε * ω = ε ω i ε ω + σ d c ε 0 ω
The DC conductivity loss, σ d c ε 0 ω , increases with decreasing frequency. The temperature dependence of σ d c can be described by the VFTH model:
σ d c = σ 0 exp E a k B ( T T 0 )
where σ 0 is the maximum conductivity at the temperature of infinity and is proportional to the number of charge carriers. k B is Boltzmann’s constant. T 0 is the Vogel temperature, which defines the low-temperature limit of the conductivity curves and is the temperature at which the ion mobility approaches zero. E a is the activation energy of the conduction process. In some systems, especially those without strong glassy dynamics, the temperature dependence of σ d c is better described by the Arrhenius law. This behavior reflects thermally activated hopping transport through a more uniform energy landscape [199].

2.6. Models for Analyzing the Relaxation Process

The permittivity of dielectric materials varies with the frequency ω of an alternating electric field. The permittivity is normally expressed by the complex permittivity ε* of a medium as a function of the angular frequency, ω = 2πf. The simplest model is the Debye dielectric model [80], which is written as follows:
ε * ω = ε + ε s ε 1 + i ω τ 0
where ε is the real permittivity at the high-frequency limit (ω = ∞). ε s is the real permittivity under a static electric field (ω = 0). τ 0 is the characteristic relaxation time for the polarization to reach an equilibrium state after a disturbance. ε s ε is the dielectric relaxation strength of the corresponding relaxation loss peak.
In the case of relaxation with respect to time distribution, the Debye dielectric relaxation model was expanded to the Cole‒Cole model [206] by introducing the empirical parameter α:
ε * ω = ε + ε s ε 1 + i ω τ 0 1 α
This Cole‒Cole equation is used when the dielectric loss peak shows symmetric broadening. α lies in the range of 0 ≤ α ≤ 1. For α = 0, the Cole‒Cole model becomes the Debye model. Furthermore, the Cole‒Davidson model [207] with a different empirical parameter β was modified as follows:
ε * ω = ε + ε s ε 1 + i ω τ 0 β
The two expanded models were integrated into the Havriliak‒Negami (HN) model [208] by introducing the two empirical parameters, α and β, as follows:
ε * ω = ε + ε s ε 1 + i ω τ 0 1 α β
This equation considers symmetric and asymmetric broadening. α and β are shape parameters with values between 0 and 1 that depend on the relaxation time distribution. α and β are related to the linewidth and asymmetric relaxation loss peak, respectively. As α and β approach zero, the relaxation spectrum exhibits increased broadening and a more asymmetrical shape. To understand the dynamics of the relaxation processes in the polymer, the data for the experimental complex dielectric permittivity, ε * ω , were fitted with the HN function. The HN function formalism can be phenomenologically described as a combination of the conductivity term with the HN functional form as follows [199,200,201]:
ε * ω = i σ d c ε 0 ω N + ε + k = 1,2 ε k 1 + i ω τ H N k 1 α k β k
where the first term accounts for conductivity. N is an exponent that characterizes the nature of the conduction process as well as the electrode polarization effect on the dielectric dispersion spectrum. This contribution becomes most significant at high temperatures and low frequencies. The temperature dependence of the exponent N for EPDM and FKM is presented in Figure S1. The last term is the HN function. The summation symbol indicates the occurrence of more than one relaxation process. τ H N k is the average characteristic relaxation time of the corresponding process k. The exponents α k and β k (0 ≤ α k ≤ 1, β k ≤ 1) are the HN fitting parameters (shape parameters) describing the distributions of the relaxation times. The dielectric relaxation strength, ε s ε , is represented by ε k .

3. Methods

3.1. Dielectric Spectroscopy

The dielectric spectroscopy system (Figure 1) used to measure the complex permittivity comprises a temperature-controlled chamber, a digital multimeter (Fluke 8842A) for temperature monitoring, and an impedance analyzer (VSP-300, Bio Logics) with a GPIB interface to a PC. The temperature is controlled and maintained by proportional–integral–derivative (PID) control with a heater and a circulating refrigerating medium, varying at intervals of 5 K with a temperature stability better than 0.1 K at temperatures ranging from 233 to 404 K. Cylindrical samples (50 mm in diameter and 2.4 mm in thickness) are placed between two circle-shaped copper electrodes. An AC voltage is applied to the electrodes, the resulting current is measured, and the impedance is obtained. The measured complex impedance values ( Z and Z ) are then converted to complex permittivity values via the following relationships:
ε = d ω ε 0 A Z Z 2 + Z 2 ,   ε = d ω ε 0 A Z Z 2 + Z 2 ,

3.2. Differential Scanning Calorimetry

Differential scanning calorimetry (DSC) was employed to determine the dynamic glass transition temperature using a Q-600 instrument (TA Instruments, New Castle, DE, USA). Approximately 10 mg of granulated rubber samples was analyzed at a controlled heating rate of 1 °C/min over the temperature range of −80 °C to 80 °C.

4. Development of Self-Developed Deconvolution Analysis Program

In numerous previous studies, deconvolution analysis via a combination of Debye-based fitting models was performed with the help of numerical fitting software or algorithms developed in C#, Python, and MATLAB. Additionally, commercialized and open-access or publicly available graphical user interface (GUI) software for data analysis also exists, as shown in Figure 2a,b. These commercial software packages offer various useful features, including support for nonlinear curve fitting via the Havriliak‒Negami model. For example, the “EIS-smart-tool” (Figure 2a) provides Nyquist plots for equivalent circuit analysis and an interactive GUI that allows for the easy rotation of 3D plots, along with a convenient fitting setup for Arrhenius plots. Furthermore, the “WinFit” (Figure 2b) software supports a diverse range of domain analysis functions, including frequency, temperature, and time. Notably, the temperature and time domain analysis features, which are rarely found in other impedance analysis software, enable the easy determination of the activation energies and glass transition temperatures.
However, in practical experimental settings, impedance data are acquired across numerous temperature points to track the evolution of dielectric relaxations with thermal variation. As a result, these tools require repetitive manual fitting procedures for each dataset, significantly increasing the analysis time and reducing overall efficiency.
To address this challenge, we developed a custom data analysis software named “Dispersion Analyzer” to facilitate comprehensive and consistent deconvolution across wide temperature ranges. The software incorporates a smoothness constraint function that enables the simultaneous fitting of multiple datasets while maintaining physical continuity between temperature-dependent parameters. Unlike conventional tools that fit each temperature individually, our software supports the following:
  • Weighted fitting between the real and imaginary parts of permittivity spectra to emphasize dominant spectral contributions;
  • Global fitting across temperature series under physically meaningful constraints (e.g., shared activation energy or Arrhenius/VFT behavior);
  • User-controllable fitting accuracy achieved by adjusting convergence thresholds, iteration limits, and optimization tolerances.
These capabilities allow for both interpretability and efficiency in complex dielectric analysis, which—based on our assessment—are not commonly supported in existing commercial packages.
The developed software and additional information are provided in the Supplementary Materials.
The self-developed deconvolution analysis program, which was written in C# using Microsoft Visual Studio Net 2018, has the GUI shown in Figure 3. It can be executed on a Microsoft Windows system. The dimensional information of the sample can be inserted into the material information at the bottom left of the main panel (presented in Figure 3) to extract the complex permittivity value from the raw impedance data. The software provides two-dimensional plots on the right window and three-dimensional plots on the left window to visualize the complex permittivity spectra as a function of frequency and temperature. Moreover, users can easily configure the fitting information, such as the number of fitting functions, constraints on fitting parameters, and weighting for the variance of fitting parameters across closely related temperatures. In the three-dimensional plot on the left window, the red and blue surfaces represent the real and imaginary parts of the complex permittivity, respectively, and the solid line represents the dielectric spectra at the specific temperature, which is displayed in the two-dimensional plot in the right panel of the main window. Furthermore, the red and blue cross symbols in two-dimensional plots indicate the measured real and imaginary permittivity data values, respectively, whereas the solid lines represent the fitted results obtained via a combination of two HN functions. The ten parameters obtained via the two HN model functions in model information are listed in the lower right section of the main window.

5. Results and Discussion

5.1. NBR

NBR is a synthetic copolymer composed of butadiene (CH2CH=CHCH2) and acrylonitrile (CH2CHCN). Its representative molecular structure is illustrated in Figure 4 [169]. The detailed chemical compositions of the NBR samples used in this study are summarized in Table 1.
The three-dimensional frequency and temperature dependences of the imaginary parts of the permittivity of NBR are shown in Figure 5a. All four relaxation processes, called α, α’, β, and γ relaxations, and the contribution of the DC conduction process were observed in the experimental temperature range (233–404 K) and frequency range (0.01–3 MHz). The yellow arrows in Figure 5a indicate the evolution of the relaxation peaks versus the temperature variation. The temperature ranges observed for each relaxation process are presented (Figure 5b) as follows:
(i).
The temperature range for β relaxation: 233–269 K.
(ii).
The temperature range for γ relaxation: 233–269 K.
(iii).
The temperature range for α relaxation: 233–341 K.
(iv).
The temperature range for α’ relaxation: 254–404 K.
(v).
The temperature range for DC conduction contribution: 324–404 K.
The simulation results of the relaxation processes for seven representative temperatures (404 K, 359 K, 326 K, 293 K, 272 K, 251 K, and 233 K) in the frequency domain after deconvolution via the self-developed deconvolution analysis program are shown in Figure 6a–g. The α′ and α relaxation processes were well described by the HN function, as given in Equation (10). The β and γ relaxations, exhibiting symmetrical loss profiles, were fitted using the Cole–Cole equation, which corresponds to the case of β = 1 in the HN model. As the temperature decreased, all relaxation loss spectra shifted toward lower frequencies.
The α’ relaxation, known as normal mode relaxation, reflects fluctuations in the end-to-end dipole vector of a polymer chain. This phenomenon typically occurs in Stockmayer type A polymers, where the dipole moment aligns along the polymer backbone [202,209,210,211]. Among the studied materials, this α’ relaxation was identified exclusively in NBR due to its dipole vector being oriented parallel to the chain backbone. In contrast, the α relaxation is linked to the cooperative segmental dynamics of the polymer backbone itself [212,213,214,215]. The β and γ relaxations, on the other hand, originated from the orientational movements of side groups—vinyl (–C₂H₃) and cyano (–CN), respectively. Notably, the γ relaxation, which involves the motion of the smaller cyano groups, typically appears at higher frequencies or lower temperatures than the β relaxation in isochronal measurements [201,216,217,218]. Figure 7 provides a schematic of the NBR molecular structure along with its characteristic relaxation processes.
Figure 8 presents the center frequency (f₀) of the relaxation peak for NBR, plotted as a function of the reciprocal temperature. As previously observed in Figure 6, the relaxation peak shifts to lower frequencies as the temperature decreases. The activation energies for the β, γ, and DC relaxation processes were extracted using the Arrhenius relationship, as expressed in Equation (2). This analysis showed that the β and γ processes exhibited activation energies of 32.0 kJ/mol and 24.3 kJ/mol, respectively. Additionally, as the contribution of the DC conduction process increased with an increasing temperature, its activation energy was 38.7 kJ/mol. Pietrasik et al. [148] reported an activation energy of 32 kJ/mol for β relaxation; this value is identical to our experimental result. In contrast, Khalaf et al. [145] reported an activation energy of 22.41 kJ/mol for the DC conduction process, which is considerably different from our results. This discrepancy may be attributed to the differences in the type of carbon black (CB) filler used. In our study, we utilized medium thermal (MT) CB as a filling agent, whereas Khalaf et al. [145] used high-abrasion furnace (HAF) CB. HAF CB has a smaller primary particle size (30~40 nm) than MT CB (250~550 nm). Thus, the aggregation structure and percolation path in HAF-filled rubber composites are better formed than those in MT-filled composites. Consequently, the electrical conductivity of HAF-filled rubber is relatively high, resulting in relatively low activation energy. Similar results were reported in our previous study [173].
In Figure 6, the center positions of the loss peak for α show nonlinear temperature dependency; therefore, the fitting of α is conducted with the VFTH function. As mentioned in Figure 5, the α relaxation phenomenon results from the polarization of the backbone chain and is related to the glass transition mechanism. The T g value can be calculated by fitting it with the VFTH equation. The T g determined using dielectric relaxation spectroscopy (DRS) was identified as the temperature at which the extrapolated VFTH fit of the central frequency ( f 0 ) versus 1000/T reaches 1.6 mHz, following an empirical criterion established in previous studies [199,200,201] (Figure 9a). Based on this method, the T g was determined to be 242.1 K (Figure 9a). In comparison, the T g obtained from DSC (Figure 9b) corresponds to the midpoint of the heat capacity change (1/2 ΔCp) often referred to as the “temperature of half-unfreezing” [201], yielding a value of 242.4 K. These two T g estimates show good agreement.

5.2. EPDM

EPDM is a synthetic rubber composed of ethylene (CH2CH2), propylene (CH2CHCH3), and ethylidene norbornene (ENB) units. Its representative molecular structure is illustrated in Figure 10, and the corresponding chemical composition is summarized in Table 2.
Figure 11a presents the three-dimensional frequency and temperature dependences of the imaginary part of the complex permittivity for EPDM. Within the experimental temperature range (233–404 K) and frequency range (0.01 Hz to 3 MHz), three distinct relaxation processes—α, β, and MWS relaxations—along with a conductivity contribution, were observed. Notably, a merged αβ relaxation behavior appeared in the temperature range of 305–404 K. The yellow arrows in Figure 11a indicate the evolution of the relaxation peaks with increasing temperature.
(i).
The temperature range for α relaxation: 233 K to 302 K.
(ii).
The temperature range for β relaxation: 233 K to 305 K.
(iii).
The temperature range for αβ merged relaxation: 305 K to 404 K.
(iv).
The temperature range for the DC conduction contribution: 308 K to 404 K.
(v).
The temperature range for MWS relaxation: 233 K to 404 K.
The deconvoluted dielectric loss spectra of EPDM at seven representative temperatures (404 K, 359 K, 326 K, 293 K, 272 K, 251 K, and 233 K) are shown in Figure 12a–g, which show the evolution of three relaxation peaks and the DC conduction process with varying temperature. The black solid lines represent the total fitted curves, composed of individual components: MWS (orange dashed), β (blue dashed), and DC conductivity (navy dashed) fitted using two HN functions and a conductivity model for the spectra at 404 K, 359 K, 326 K, and 293 K or the fitted sums of the MWS (orange dashed line), α (red dashed line), and β (blue dashed line) relaxations by three HN fits (272 K, 251 K, and 233 K). The αβ merged relaxation process from 305 K to 404 K originates from the β process. All relaxation peaks shift toward higher frequencies with an increasing temperature.
The α relaxation in EPDM is observed at temperatures below room temperature. In the dielectric loss spectra, each frequency-dependent plot exhibits a broad and symmetrical loss peak, attributed to the local segmental motion of the polymer backbone. In contrast, the β relaxation, which appears across the entire temperature range studied, presents a narrower and symmetrical loss peak. This process is associated with the rotational dynamics of the side groups. Specifically, the α relaxation corresponds to the rotational motion around the main chain C–C bonds in the ethylene, propylene, and ENB units, as illustrated in Figure 13. Meanwhile, the β relaxation arises from the reorientation of the methyl (CH₃) groups located on the side chains of the propylene units, also depicted in Figure 13.
The position (f₀) of the central relaxation peak as a function of the reciprocal temperature for EPDM is presented in Figure 14. As the temperature decreases, the entire dielectric loss spectra shift toward lower frequencies. The inset of Figure 14 shows that the dielectric strength (Δε) of the α relaxation increases with decreasing temperature, while that of the β relaxation decreases. The temperature behavior of Δε is a general trend for α and β relaxation.
The activation energies for the β and MWS relaxation processes were estimated using the Arrhenius temperature dependence described by Equation (2), yielding values of 17.7 kJ/mol and 113 kJ/mol, respectively. Due to the limited availability of prior studies on the activation energies of EPDM rubber compounds, direct comparisons with the existing literature were not conducted in this work. As shown in Figure 15, the α and β relaxation peaks are clearly distinguishable in the temperature range of 233 K to 290 K, but gradually merge between 290 K and 373 K. The merging point at 290 K corresponds to the temperature where both processes exhibit the same relaxation time, as evidenced by the intersection in both the Δε (inset) and f₀ plots. This is the first time this has been observed in polymers with the help of the deconvolution analysis program. Additionally, because EPDM contains flexibly attached dipolar side groups (methyl groups (-CH3)), the value of Δε for β relaxation was smaller than that for α relaxation at low temperatures [219,220].
The temperature dependence of the α relaxation was analyzed using the VFTH model (Equation (1)). The Tg of EPDM was determined and compared with DSC measurements. The results obtained from both measurement methods are presented in Figure 15a,b. As shown in Figure 15a,b, the Tg value obtained from dielectric measurements was 227.3 K, while the DSC measurement yielded a value of 221.0 K. The relatively large difference between the two methods may be attributed to differences in the heating rates used in each technique. However, the exact cause of this discrepancy is still under investigation.

5.3. FKM

The FKMs shown in Figure 16 are fluorocarbon-based synthetic rubbers synthesized by copolymerizing vinylidene fluoride (CH2=CF2), hexafluoropropylene (CF2=CFCF3), chlorotrifluoroethylene (CF2=CFCL), and tetrafluoroethylene (C2=F4). Due to their excellent chemical stability, thermal resistance, low electrical conductivity, and cost-effectiveness, FKMs are widely employed in various industrial applications [221,222,223,224,225]. However, various studies have reported on the dielectric properties of NBR, SBR, and EPDM copolymers. In contrast, the dielectric dispersion analysis of FKM copolymers remains rare [129,130,131,132,133,134,168,172]. The chemical compositions of the FKM used in this study are summarized in Table 3.
Figure 17a shows three-dimensional dielectric loss spectra over a wide range of temperatures (233 K–373 K) and frequencies (0.01 Hz–3 MHz). Within these temperature and frequency ranges, MWS relaxation, α relaxation, β relaxation, and the DC conduction process were observed. The arrows indicate the evolution of the corresponding relaxation type during temperature variation. Each temperature range used to observe these relaxation phenomena is presented (Figure 17b) as follows:
(i).
The temperature range for β relaxation: 233 K to 308 K.
(ii).
The temperature range for α relaxation: 268 K to 308 K.
(iii).
The temperature range for MWS relaxation: 268 K to 373 K.
(iv).
The temperature range for the DC conduction contribution: 283 K to 373 K.
To extract valuable physical parameters from the dispersion spectra, the dielectric frequency response data for each temperature point were deconvoluted, as illustrated in Figure 18. The data were fitted via a combination of HN functions and the power law to account for DC conductivity contributions. Four main relaxation peaks were identified, and DC contributions emerged in the high-temperature and low-frequency ranges. Each β and MWS relaxation displays a symmetrical peak shape, and the peak for α is presented as an asymmetrical line shape. Furthermore, the linewidth of MWS relaxation was narrower than that of the α and β relaxations.
The α relaxation in FKM arises from the polarization of the polymer backbone (segmental chain) and is associated with the glass transition dynamics of the polymeric system. In contrast, the β relaxation originates from the rotational fluctuations and local motions of the trifluoromethyl groups (TFM:CF3) in the hexafluoropropylene (HFP) units. The molecular origins and corresponding motional modes of these dielectric relaxation processes are schematically illustrated in Figure 19.
The deconvolution of the dielectric loss spectra allowed for the extraction of the center frequencies (f₀) corresponding to the α, β, and MWS relaxation processes, as well as DC conductivity (σ_dc). These values were plotted against the reciprocal of temperature, as shown in Figure 20. In this plot, both the β and MWS relaxations exhibit a linear dependence on the reciprocal temperature and a logarithmic variation in f_0. Therefore, fitting for β and MWS relaxation was performed via the Arrhenius temperature dependence function (Eq. 2.). From these fitting results, the relaxation rates for the β and MWS relaxation processes provided activation energies of Ea = 50.88 kJ/mol and Ea = 88.57 kJ/mol, respectively. The logarithm of DC conductivity also decreased as the reciprocal of T increased; the Ea of DC conductivity was calculated via the VFTH model (Eq. 5.) as 3.83 eV. These activation energy values are consistent with those reported in four previous studies, as summarized in Table 4. Despite differences in sample preparation and composite conditions, the obtained activation energy values were comparable with previously reported values.
Similarly to a previous study on NBR and EPDM rubber compounds, we evaluated the rate of α relaxation for FKM via the VFTH temperature dependence law. Consequently, the Tg value of the FKM was obtained via dielectric relaxation spectroscopy (DRS) and compared with that obtained via DSC measurements. The results obtained using both measurement methods are presented in Figure 21a,b. The Tg value obtained via dielectric measurements was 252.6 K, and the Tg value obtained via DSC measurements was 253.9 K. Both results revealed comparable Tg values with an uncertainty of less than 1 K.

6. Conclusions

This study involved a comprehensive analysis of the dielectric responses of nitrile–butadiene rubber (NBR), ethylene–propylene–diene monomer (EPDM), and fluoroelastomer (FKM) copolymers across a wide range of frequencies and temperatures. A self-developed “Dispersion Analysis” program, implemented in C#, was utilized to deconvolute the dielectric relaxation spectra. This program utilized a smoothness function to constrain fitting parameters across multiple temperature ranges, enabling the automated tracking of relaxation peaks with temperature variations. Consequently, this developed program allows for precise and efficient deconvolution analysis via a combination of various fitting functions and related parameters.
The measured dielectric spectra revealed four main relaxation processes, including the DC conduction process in NBR, namely, α, α’, β, and γ relaxation. In contrast, three main α, β, and Maxwell‒Wagner‒Sillars (MWS) relaxations and DC conduction processes were identified in EPDM and FKM. The origins of these observed relaxation processes were identified entirely in relation to the molecular structure of the polymer chain. The activation energies for α’, β, γ, MWS, and DC conductivity were determined by applying the Arrhenius temperature dependence law to the position of the imaginary loss peak versus the reciprocal temperature. The activation energies for NBR and FKM were also compared with previously published results. Furthermore, the glass transition temperatures (Tg) for three rubber polymers was estimated via Vogel–Fulcher–Tammann–Hesse (VFTH) fitting to the rate of α relaxation. The Tg values obtained through dielectric spectroscopy agreed well with the differential scanning calorimetry (DSC) measurements.
Additionally, by analyzing the temperature dependence of the dielectric relaxation strength and the position of the relaxation peak in EPDM, the emergence and splitting of α and β relaxations in EPDM were observed for the first time in this study.
In conclusion, this study provides fundamental insights into the dielectric properties of NBR, EPDM, and FKM, which serve as valuable references for future research and the development of advanced rubber materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym17111539/s1. Figure S1. (a) The screenshot image of fitting status panel. Red cross symbols and solid red lines represent the measured and fitted real parts of the permittivity, respectively, while blue cross symbols and solid blue lines indicate the measured and fitted imaginary parts. The calculated Figure of Merit (FOM), representing the fitting accuracy, is shown in the lower-left corner of the iteration tab. (b) Screenshot of the detailed iteration setup panel. Users can define their fitting strategy by selecting the optimization method (e.g., Simplex, Hessian, or Parallel), the number of iterations, and the number of CPU cores to be used for parallel computation. Figure S2. (a) All of the fitted parameters as imposing smoothness requirements are plotted as a function of temperature; (b) Algorithm of the dispersion analysis program, which employs parallel Nelder‒Mead optimization. The entire code of this program is available in the attached supplementary zip file. Figure S3. The real and imaginary parts of the complex permittivity for NBR (a,b), EPDM (c,d), and FKM (e,f), respectively. Figure S4. Temperature dependence of the power-law exponent N for EPDM (a) and FKM (b). References [228,229] are cited in the Supplementary Materials.

Author Contributions

Y.M.: investigation, data curation, formal analysis, writing—original draft, and editing. G.K.: investigation, data curation, formal analysis, writing—original draft, and editing. J.J.: conceptualization, methodology, writing—review, supervision, and project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Evaluation Institute of Industrial Technology (KEIT) and was granted financial resources from the Ministry of Trade, Industry & Energy, Republic of Korea (No. RS-2024-00439368). Further support was provided by the Development of Reliability Measurement Technology for Hydrogen Refueling Station funded by the Korea Research Institute of Standards and Science (KRISS-2025-GP2025-0014).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data used to support the findings of this study are available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rodgers, B. Rubber Compounding: Chemistry and Applications; CRC Press: New York, NY, USA, 2015. [Google Scholar]
  2. Vijayaram, T.R. A technical review on rubber. Int. J. Des. Manuf. Technol. 2009, 3, 25–37. [Google Scholar] [CrossRef]
  3. Azahar, N.M.; Hassan, N.A.; Jaya, R.P.; Kadir, M.A.A.; Yunus, N.Z.M.; Mahmud, M.Z.H. An overview on natural rubber application for asphalt modification. Int. J. Agric. For. Plant. 2016, 2, 212–218. [Google Scholar]
  4. Myhre, M.; Saiwari, S.; Dierkes, W.; Noordermeer, J. Rubber recycling: Chemistry, processing, and applications. Rubber Chem. Technol. 2012, 85, 408–449. [Google Scholar] [CrossRef]
  5. Yoon, B.; Kim, J.Y.; Hong, U.; Oh, M.K.; Kim, M.; Han, S.B.; Nam, J.-D.; Suhr, J. Dynamic viscoelasticity of silica-filled styrene-butadiene rubber/polybutadiene rubber (SBR/BR) elastomer composites. Compos. Part B Eng. 2020, 187, 107865. [Google Scholar] [CrossRef]
  6. Liu, Z. Rational design of high-performance nanomaterials for electric vehicle tires. Highlights Sci. Eng. Technol. 2023, 62, 110–118. [Google Scholar] [CrossRef]
  7. Liao, L.; Zuo, Y.; Meng, H.; Liao, X. Research on the technology of noise reduction in hybrid electric vehicle with composite materials. Adv. Mech. Eng. 2018, 10, 1687814018766916. [Google Scholar] [CrossRef]
  8. Sancho-Turrado, A.; Lopez, J.A.R.; Calleja, A.J. Improved composite for tires of urban electric vehicles. In Proceedings of the 2013 International Conference on New Concepts in Smart Cities: Fostering Public and Private Alliances (SmartMILE), Gijon, Spain, 11–13 December 2013; pp. 1–6. [Google Scholar]
  9. Rahnavard, M.; Motlagh, G.H.; Heidarian, J.; Mobarake, M.D. Improving heat ageing, mechanical properties, and thermal conductivity of silicon rubber by surface modification of iron oxide and field testing of silicon rubber O-rings. J. Pet. Sci. Technol. 2023, 13, 17–27. [Google Scholar]
  10. Yamabe, J.; Nishimura, S. Influence of fillers on hydrogen penetration properties and blister fracture of rubber composites for O-ring exposed to high-pressure hydrogen gas. Int. J. Hydrogen Energy 2009, 34, 1977–1989. [Google Scholar] [CrossRef]
  11. Mao, H.; Zou, H.; Liu, W.; Zhuang, X.; Xing, B. Preparation and application research of composites with low vacuum outgassing and excellent electromagnetic sealing performance. e-Polymers 2024, 24, 170–177. [Google Scholar] [CrossRef]
  12. Zhou, C.; Zheng, J.; Gu, C.; Zhao, Y.; Liu, P. Sealing performance analysis of rubber O-ring in high-pressure gaseous hydrogen based on finite element method. Int. J. Hydrogen Energy 2017, 42, 11996–12004. [Google Scholar] [CrossRef]
  13. Jung, J.-K.; Kim, K.-T.; Chung, N.-K.; Baek, U.-B.; Nahm, S.-H. Characterizing the diffusion property of hydrogen sorption and desorption processes in several spherical-shaped polymers. Polymers 2022, 14, 1468. [Google Scholar] [CrossRef] [PubMed]
  14. Vinogradov, M.L.; Barchenko, V.T.; Lisenkov, A.A.; Kostrin, D.K.; Babinov, N.A. Gas permeation through vacuum materials: Mass-spectrometry measurement system. Vak. Forsch. Prax. 2015, 27, 26–29. [Google Scholar] [CrossRef]
  15. Chea, S.; Luengchavanon, M.; Anancharoenwong, E.; Techato, K.-a.; Jutidamrongphan, W.; Chaiprapat, S.; Niyomwas, S.; Marthosa, S. Development of an O-ring from NR/EPDM filled silica/CB hybrid filler for use in a solid oxide fuel cell testing system. Polym. Test. 2020, 88, 106568. [Google Scholar] [CrossRef]
  16. Rivera, M. Elastomers in space and in other high vacuum environments. Rubber Chem. Technol. 1966, 39, 1127–1140. [Google Scholar] [CrossRef]
  17. Koga, A.; Uchida, K.; Yamabe, J.; Nishimura, S. Evaluation on high-pressure hydrogen decompression failure of rubber O-ring using design of experiments. Int. J. Automot. Eng. 2011, 2, 123–129. [Google Scholar] [CrossRef]
  18. Jin, L.; Che, L. Experimental study on stress relaxation of rubber O-ring for mechanical seal. J. Phys. Conf. Ser. 2024, 2691, 012015. [Google Scholar] [CrossRef]
  19. Yang, H.; Yao, X.; Zheng, Z.; Liu, Y. Graphene rubber composites integrated sealing rings for monitoring contact pressure and the aging process. Compos. Part A Appl. Sci. Manuf. 2019, 118, 171–178. [Google Scholar] [CrossRef]
  20. Jung, J.K.; Lee, J.H.; Kim, Y.W.; Chung, N.K. Development of portable gas sensing system for measuring gas emission concentration and diffusivity using commercial manometric sensors in gas exposed polymers: Application to pure gases, H2, He, N2, O2 and Ar. Sens. Actuators B Chem. 2024, 418, 136240. [Google Scholar] [CrossRef]
  21. Jung, J.K.; Lee, J.H. High-performance hydrogen gas sensor system based on transparent coaxial cylinder capacitive electrodes and a volumetric analysis technique. Sci. Rep. 2024, 14, 1967. [Google Scholar] [CrossRef]
  22. Jung, J.K.; Kim, K.-T.; Lee, J.H.; Baek, U.B. Effective and low-cost gas sensor based on a light intensity analysis of a webcam image: Gas enriched polymers under high pressure. Sens. Actuators B Chem. 2023, 393, 134258. [Google Scholar] [CrossRef]
  23. Jung, J.K.; Kim, K.-T.; Baek, U.B. Simultaneous three-channel measurements of hydrogen diffusion with light intensity analysis of images by employing webcam. Curr. Appl. Phys. 2022, 37, 19–26. [Google Scholar] [CrossRef]
  24. Jung, J.; Kim, G.; Gim, G.; Park, C.; Lee, J. Determination of gas permeation properties in polymer using capacitive electrode sensors. Sensors 2022, 22, 1141. [Google Scholar] [CrossRef] [PubMed]
  25. Lee, J.H.; Kim, Y.W.; Kim, D.J.; Chung, N.K.; Jung, J.K. Comparison of two methods for measuring the temperature dependence of H2 permeation parameters in nitrile butadiene rubber polymer composites blended with fillers: The volumetric analysis method and the differential pressure method. Polymers 2024, 16, 280. [Google Scholar] [CrossRef]
  26. Lee, J.-H.; Kim, Y.-W.; Jung, J.-K. Investigation of the gas permeation properties using the volumetric analysis technique for polyethylene materials enriched with pure gases under high pressure: H2, He, N2, O2 and Ar. Polymers 2023, 15, 4019. [Google Scholar] [CrossRef] [PubMed]
  27. Jung, J.K.; Lee, J.H.; Jang, J.S.; Chung, N.K.; Park, C.Y.; Baek, U.B.; Nahm, S.H. Characterization technique of gases permeation properties in polymers: H2, He, N2 and Ar gas. Sci. Rep. 2022, 12, 3328. [Google Scholar] [CrossRef]
  28. Jung, J.K.; Kim, K.-T.; Chung, K.S. Two volumetric techniques for determining the transport properties of hydrogen gas in polymer. Mater. Chem. Phys. 2022, 276, 125364. [Google Scholar] [CrossRef]
  29. Jung, J.K.; Kim, I.G.; Chung, K.S.; Baek, U.B. Analyses of permeation characteristics of hydrogen in nitrile butadiene rubber using gas chromatography. Mater. Chem. Phys. 2021, 267, 124653. [Google Scholar] [CrossRef]
  30. Jung, J.K.; Kim, I.G.; Jeon, S.K.; Kim, K.-T.; Baek, U.B.; Nahm, S.H. Volumetric analysis technique for analyzing the transport properties of hydrogen gas in cylindrical-shaped rubbery polymers. Polym. Test. 2021, 99, 107147. [Google Scholar] [CrossRef]
  31. Jung, J.K.; Kim, I.G.; Kim, K.T.; Ryu, K.S.; Chung, K.S. Evaluation techniques of hydrogen permeation in sealing rubber materials. Polym. Test. 2021, 93, 107016. [Google Scholar] [CrossRef]
  32. Jung, J.K.; Kim, I.G.; Chung, K.S.; Kim, Y.-I.; Kim, D.H. Determination of permeation properties of hydrogen gas in sealing rubbers using thermal desorption analysis gas chromatography. Sci. Rep. 2021, 11, 17092. [Google Scholar] [CrossRef]
  33. Jung, J.K.; Jeon, S.K.; Kim, K.-T.; Lee, C.H.; Baek, U.B.; Chung, K.S. Impedance spectroscopy for in situ and real-time observations of the effects of hydrogen on nitrile butadiene rubber polymer under high pressure. Sci. Rep. 2019, 9, 13035. [Google Scholar] [CrossRef] [PubMed]
  34. Lee, J.H.; Kim, Y.W.; Chung, N.K.; Kang, H.M.; Moon, W.J.; Chan Choi, M.; Jung, J.K. Multiphase modeling of pressure-dependent hydrogen diffusivity in fractal porous structures of acrylonitrile butadiene rubber-carbon black composites with different fillers. Polymer 2024, 311, 127552. [Google Scholar] [CrossRef]
  35. Lee, C.H.; Jung, J.K.; Kim, K.S.; Kim, C.J. Hierarchical channel morphology in O-rings after two cycling exposures to 70 MPa hydrogen gas: A case study of sealing failure. Sci. Rep. 2024, 14, 5319. [Google Scholar] [CrossRef] [PubMed]
  36. Jung, J.K.; Lee, J.H.; Jeon, S.K.; Tak, N.H.; Chung, N.K.; Baek, U.B.; Lee, S.H.; Lee, C.H.; Choi, M.C.; Kang, H.M.; et al. Correlations between H2 permeation and physical/mechanical properties in ethylene propylene diene monomer polymers blended with carbon black and silica fillers. Int. J. Mol. Sci. 2023, 24, 2865. [Google Scholar] [CrossRef]
  37. Moon, Y.I.; Lee, H.E.; Jung, J.K.; Han, H.W. Direct visualization of carbon black aggregates in nitrile butadiene rubber by THz near-field microscope. Sci. Rep. 2023, 13, 7846. [Google Scholar] [CrossRef]
  38. Jung, J.K.; Baek, U.B.; Lee, S.H.; Choi, M.C.; Bae, J.W. Hydrogen gas permeation in peroxide-crosslinked ethylene propylene diene monomer polymer composites with carbon black and silica fillers. J. Polym. Sci. 2023, 61, 460–471. [Google Scholar] [CrossRef]
  39. Jung, J.K.; Lee, C.H.; Baek, U.B.; Choi, M.C.; Bae, J.W. Filler influence on H2 permeation properties in sulfur-crosslinked ethylene propylene diene monomer polymers blended with different concentrations of carbon black and silica fillers. Polymers 2022, 14, 592. [Google Scholar] [CrossRef]
  40. Jung, J.K.; Lee, C.H.; Son, M.S.; Lee, J.H.; Baek, U.B.; Chung, K.S.; Choi, M.C.; Bae, J.W. Filler effects on H2 diffusion behavior in nitrile butadiene rubber blended with carbon black and silica fillers of different concentrations. Polymers 2022, 14, 700. [Google Scholar] [CrossRef]
  41. Jung, J.K.; Baek, U.B.; Nahm, S.H.; Chung, K.S. Hydrogen sorption and desorption properties in rubbery polymer. Mater. Chem. Phys. 2022, 279, 125745. [Google Scholar] [CrossRef]
  42. Kang, H.; Bae, J.; Lee, J.; Yun, Y.; Jeon, S.; Chung, N.; Jung, J.; Baek, U.; Lee, J.; Kim, Y.; et al. The synergistic effect of carbon black/carbon nanotube hybrid fillers on the physical and mechanical properties of EPDM composites after exposure to high-pressure hydrogen gas. Polymers 2024, 16, 1065. [Google Scholar] [CrossRef]
  43. Chai, A.B.; Andriyana, A.; Verron, E.; Johan, M.R.; Haseeb, A.S.M.A. Development of a compression test device for investigating interaction between diffusion of biodiesel and large deformation in rubber. Polym. Test. 2011, 30, 867–875. [Google Scholar] [CrossRef]
  44. Rowland, S.M.; Robertson, J.; Xiong, Y.; Day, R.J. Electrical and material characterization of field-aged 400 kV silicone rubber composite insulators. IEEE Trans. Dielectr. Electr. Insul. 2010, 17, 375–383. [Google Scholar] [CrossRef]
  45. Dick, T.A.; dos Santos, L.A. In situ synthesis and characterization of hydroxyapatite/natural rubber composites for biomedical applications. Mater. Sci. Eng. C 2017, 77, 874–882. [Google Scholar] [CrossRef] [PubMed]
  46. Song, K. Interphase characterization in rubber nanocomposites. In Progress in Rubber Nanocomposites; Thomas, S., Maria, H.J., Eds.; Elsevier Inc.: Cambridge, MA, USA, 2017; pp. 115–152. [Google Scholar]
  47. Bandyopadhyay, A.; Bhowmick, A.K.; De Sarkar, M. Synthesis and characterization of acrylic rubber/silica hybrid composites prepared by sol-gel technique. J. Appl. Polym. Sci. 2004, 93, 2579–2589. [Google Scholar] [CrossRef]
  48. Iqbal, N.; Khan, M.B.; Sagar, S.; Maqsood, A. Fabrication and characterization of multiwalled carbon nanotubes/silicone rubber composites. J. Appl. Polym. Sci. 2012, 128, 2439–2446. [Google Scholar] [CrossRef]
  49. Chaikumpollert, O.; Yamamoto, Y.; Suchiva, K.; Kawahara, S. Mechanical properties and cross-linking structure of cross-linked natural rubber. Polym. J. 2012, 44, 772–777. [Google Scholar] [CrossRef]
  50. Jung, J.K.; Lee, J.H.; Park, J.Y.; Jeon, S.K. Modeling of the time-dependent H2 emission and equilibrium time in H2-enriched polymers with cylindrical, spherical and sheet shapes and comparisons with experimental investigations. Polymers 2024, 16, 2158. [Google Scholar] [CrossRef]
  51. Zedler, Ł.; Colom, X.; Saeb, M.R.; Formela, K. Preparation and characterization of natural rubber composites highly filled with brewers’ spent grain/ground tire rubber hybrid reinforcement. Compos. Part B Eng. 2018, 145, 182–188. [Google Scholar] [CrossRef]
  52. Qiu, X.; Yin, H.; Xing, Q. Research progress on fatigue life of rubber materials. Polymers 2022, 14, 4592. [Google Scholar] [CrossRef]
  53. Genovese, A.; D’Angelo, G.A.; Sakhnevych, A.; Farroni, F. Review on friction and wear test rigs: An overview on the state of the art in tyre tread friction evaluation. Lubricants 2020, 8, 91. [Google Scholar] [CrossRef]
  54. Jeon, S.K.; Jung, J.K.; Chung, N.K.; Baek, U.B.; Nahm, S.H. Investigation of physical and mechanical characteristics of rubber materials exposed to high-pressure hydrogen. Polymers 2022, 14, 2233. [Google Scholar] [CrossRef] [PubMed]
  55. Valentín, J.L.; Carretero-González, J.; Mora-Barrantes, I.; Chassé, W.; Saalwächter, K. Uncertainties in the determination of cross-link density by equilibrium swelling experiments in natural rubber. Macromolecules 2008, 41, 4717–4729. [Google Scholar] [CrossRef]
  56. Karabork, F.; Pehlivan, E.; Akdemir, A. Characterization of styrene butadiene rubber and microwave devulcanized ground tire rubber composites. J. Polym. Eng. 2014, 34, 543–554. [Google Scholar] [CrossRef]
  57. Soma, P.; Tada, N.; Uchida, M.; Nakahara, K.; Taga, Y. A fracture mechanics approach for evaluating the effects of heat aging on fatigue crack growth of vulcanized natural rubber. J. Solid Mech. Mater. Eng. 2010, 4, 727–737. [Google Scholar] [CrossRef]
  58. Rattanasom, N.; Prasertsri, S.; Ruangritnumchai, T. Comparison of the mechanical properties at similar hardness level of natural rubber filled with various reinforcing-fillers. Polym. Test. 2009, 28, 8–12. [Google Scholar] [CrossRef]
  59. Jung, J.K. Review of developed methods for measuring gas uptake and diffusivity in polymers enriched by pure gas under high pressure. Polymers 2024, 16, 723. [Google Scholar] [CrossRef]
  60. Jung, J.K.; Kim, I.G.; Kim, K.T. Evaluation of hydrogen permeation characteristics in rubbery polymers. Curr. Appl. Phys. 2021, 21, 43–49. [Google Scholar] [CrossRef]
  61. Jung, J.K.; Kim, I.G.; Chung, K.S.; Baek, U.B. Gas chromatography techniques to evaluate the hydrogen permeation characteristics in rubber: Ethylene propylene diene monomer. Sci. Rep. 2021, 11, 4859. [Google Scholar] [CrossRef]
  62. Choi, B.-L.; Jung, J.K.; Baek, U.B.; Choi, B.-H. Effect of functional fillers on tribological characteristics of acrylonitrile butadiene rubber after high-pressure hydrogen exposures. Polymers 2022, 14, 861. [Google Scholar] [CrossRef]
  63. Jung, J.K.; Kim, I.G.; Kim, K.-T.; Baek, U.B.; Nahm, S.H. Novel volumetric analysis technique for characterizing the solubility and diffusivity of hydrogen in rubbers. Curr. Appl. Phys. 2021, 26, 9–15. [Google Scholar] [CrossRef]
  64. Lee, J.H.; Jung, J.K. Development of image-based water level sensor with high-resolution and low-cost using image processing algorithm: Application to outgassing measurements from gas-enriched polymer. Sensors 2024, 24, 7699. [Google Scholar] [CrossRef] [PubMed]
  65. Jung, J.K.; Lee, J.H.; Jeon, S.K.; Baek, U.B.; Lee, S.H.; Lee, C.H.; Moon, W.J. H2 uptake and diffusion characteristics in sulfur-crosslinked ethylene propylene diene monomer polymer composites with carbon black and silica fillers after high-pressure hydrogen exposure reaching 90 MPa. Polymers 2023, 15, 162. [Google Scholar] [CrossRef] [PubMed]
  66. Jung, J.K.; Kim, K.T.; Baek, U.B.; Nahm, S.H. Volume dependence of hydrogen diffusion for sorption and desorption processes in cylindrical-shaped polymers. Polymers 2022, 14, 756. [Google Scholar] [CrossRef] [PubMed]
  67. Lee, J.H.; Jung, J.K.; Baek, U.B.; Chun, K.S. Development of measurement technology for uptake and diffusivity of hydrogen gas in rubber materials using volumetric analysis. Trans. Korean Hydrog. New Energy Soc. 2022, 33, 67–76. [Google Scholar] [CrossRef]
  68. Chakraborty, S.; Bandyopadhyay, S.; Ameta, R.; Mukhopadhyay, R.; Deuri, A.S. Application of FTIR in characterization of acrylonitrile-butadiene rubber (nitrile rubber). Polym. Test. 2007, 26, 38–41. [Google Scholar] [CrossRef]
  69. Noriman, N.Z.; Ismail, H.; Rashid, A.A. Characterization of styrene butadiene rubber/recycled acrylonitrile-butadiene rubber (SBR/NBRr) blends: The effects of epoxidized natural rubber (ENR-50) as a compatibilizer. Polym. Test. 2010, 29, 200–208. [Google Scholar] [CrossRef]
  70. Irez, A.B.; Bayraktar, E.; Miskioglu, I. Design and mechanical-physical properties of epoxy-rubber based composites reinforced with nanoparticles. Procedia Eng. 2017, 184, 486–496. [Google Scholar] [CrossRef]
  71. Lim, L.P.; Juan, J.C.; Huang, N.M.; Goh, L.K.; Leng, F.P.; Loh, Y.Y. Enhanced tensile strength and thermal conductivity of natural rubber graphene composite properties via rubber-graphene interaction. Mater. Sci. Eng. B 2019, 246, 112–119. [Google Scholar] [CrossRef]
  72. Kralevich, M.L.; Koenig, J.L. FTIR analysis of silica-filled natural rubber. Rubber Chem. Technol. 1998, 71, 300–309. [Google Scholar] [CrossRef]
  73. Gunasekaran, S.; Natarajan, R.K.; Kala, A. FTIR spectra and mechanical strength analysis of some selected rubber derivatives. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2007, 68, 323–330. [Google Scholar] [CrossRef]
  74. Yang, Z.; Peng, H.; Wang, W.; Liu, T. Crystallization behavior of poly(ε-caprolactone)/layered double hydroxide nanocomposites. J. Appl. Polym. Sci. 2010, 116, 2658–2667. [Google Scholar] [CrossRef]
  75. Khimi, S.R.; Pickering, K.L. A new method to predict optimum cure time of rubber compound using dynamic mechanical analysis. J. Appl. Polym. Sci. 2013, 131. [Google Scholar] [CrossRef]
  76. Kararli, T.T.; Hurlbut, J.B.; Needham, T.E. Glass–rubber transitions of cellulosic polymers by dynamic mechanical analysis. J. Pharm. Sci. 1990, 79, 845–848. [Google Scholar] [CrossRef] [PubMed]
  77. Fernández-Berridi, M.J.; González, N.; Mugica, A.; Bernicot, C. Pyrolysis-FTIR and TGA techniques as tools in the characterization of blends of natural rubber and SBR. Thermochim. Acta 2006, 444, 65–70. [Google Scholar] [CrossRef]
  78. Lee, C.H.; Park, S.-H.; Jung, J.K.; Ryu, K.-S.; Nahm, S.H.; Kim, J.; Chen, Y. 11B nuclear magnetic resonance study of boron nitride nanotubes prepared by mechano-thermal method. Solid State Commun. 2005, 134, 419–423. [Google Scholar] [CrossRef]
  79. Bueche, F. Tensile strength of rubbers. J. Polym. Sci. 1957, 24, 189–200. [Google Scholar] [CrossRef]
  80. Debye, P. Zur theorie der spezifischen wärmen. Ann. Phys. 1912, 344, 789–839. [Google Scholar] [CrossRef]
  81. Böhmer, R.; Gainaru, C.; Richert, R. Structure and dynamics of monohydroxy alcohols—Milestones towards their microscopic understanding, 100 years after Debye. Phys. Rep. 2014, 545, 125–195. [Google Scholar] [CrossRef]
  82. Hill, N.E. Dielectric Properties and Molecular Behaviour; The Van Nostrand Series in Physical Chemistry; UT Back-in-Print Service; Van Nostrand Reinhold Company: New York, NY, USA, 2006. [Google Scholar]
  83. Saltas, V.; Vallianatos, F.; Gidarakos, E. Charge transport in diatomaceous earth studied by broadband dielectric spectroscopy. Appl. Clay Sci. 2013, 80–81, 226–235. [Google Scholar] [CrossRef]
  84. George, D.K.; Charkhesht, A.; Vinh, N.Q. New terahertz dielectric spectroscopy for the study of aqueous solutions. Rev. Sci. Instrum. 2015, 86, 123105. [Google Scholar] [CrossRef]
  85. Baxter, J.B.; Guglietta, G.W. Terahertz spectroscopy. Anal. Chem. 2011, 83, 4342–4368. [Google Scholar] [CrossRef] [PubMed]
  86. Delgado, A.; García-Sánchez, M.F.; M’Peko, J.-C.; Ruiz-Salvador, A.R.; Rodríguez-Gattorno, G.; Echevarría, Y.; Fernández-Gutierrez, F. An elementary picture of dielectric spectroscopy in solids: Physical basis. J. Chem. Educ. 2003, 80, 1062. [Google Scholar] [CrossRef]
  87. Tonkoshkur, A.S.; Glot, A.B.; Ivanchenko, A.V. Basic models in dielectric spectroscopy of heterogeneous materials with semiconductor inclusions. Multidiscip. Model. Mater. Struct. 2017, 13, 36–57. [Google Scholar] [CrossRef]
  88. Jonscher, A.K. Dielectric relaxation in solids. J. Phys. D Appl. Phys. 1999, 32, R57–R70. [Google Scholar] [CrossRef]
  89. Macdonald, J.R. Impedance Spectroscopy: Emphasizing Solid Materials and Systems; Wiley: New York, NY, USA, 1987. [Google Scholar]
  90. Irvine, J.T.S.; Sinclair, D.C.; West, A.R. Electroceramics: Characterization by impedance spectroscopy. Adv. Mater. 1990, 2, 132–138. [Google Scholar] [CrossRef]
  91. Kremer, F.; Vallerien, S.U.; Zentel, R. Dielectric spectroscopy: Fashionable again. Adv. Mater. 1990, 2, 145–147. [Google Scholar] [CrossRef]
  92. Asami, K.; Yonezawa, T.; Wakamatsu, H.; Koyanagi, N. Dielectric spectroscopy of biological cells. Bioelectrochem. Bioenerg. 1996, 40, 141–145. [Google Scholar] [CrossRef]
  93. Poplavko, Y. Dielectric Spectroscopy of Electronic Materials: Applied Physics of Dielectrics; Woodhead Publishing Series in Electronic and Optical Materials; Woodhead Publishing: Cambridge, MA, USA, 2021. [Google Scholar]
  94. Lee, S.B.; Smith, R.L.; Inomata, H.; Arai, K. Coaxial probe and apparatus for measuring the dielectric spectra of high pressure liquids and supercritical fluid mixtures. Rev. Sci. Instrum. 2000, 71, 4226–4230. [Google Scholar] [CrossRef]
  95. Gabriel, S.; Lau, R.W.; Gabriel, C. The dielectric properties of biological tissues: II. Measurements in the frequency range 10 Hz to 20 GHz. Phys. Med. Biol. 1996, 41, 2251–2269. [Google Scholar] [CrossRef]
  96. Markus, P.; Martínez-Tong, D.E.; Papastavrou, G.; Alegria, A. Effect of environmental humidity on the ionic transport of poly(ethylene oxide) thin films, investigated by local dielectric spectroscopy. Soft Matter 2020, 16, 3203–3208. [Google Scholar] [CrossRef]
  97. Kuchler, F.; Farber, R.; Franck, C.M. Humidity and temperature effects on the dielectric properties of PET film. In Proceedings of the 2020 IEEE Electrical Insulation Conference (EIC), Knoxville, TN, USA, 22 June–3 July 2020; pp. 179–183. [Google Scholar]
  98. Gennaro, A.; Rosa, A.S.; Cornelis, P.; Pfeiffer, H.; Disalvo, E.A.; Wagner, P.; Wübbenhorst, M. A compact device for simultaneous dielectric spectroscopy and microgravimetric analysis under controlled humidity. Rev. Sci. Instrum. 2019, 90, 125106. [Google Scholar] [CrossRef] [PubMed]
  99. Zafar, R.; Gupta, N. Dielectric spectroscopy of epoxy-based barium titanate nanocomposites: Effect of temperature and humidity. IET Nanodielectr. 2020, 3, 20–27. [Google Scholar] [CrossRef]
  100. Bibi, F.; Guillaume, C.; Vena, A.; Gontard, N.; Sorli, B. Wheat gluten, a bio-polymer layer to monitor relative humidity in food packaging: Electric and dielectric characterization. Sens. Actuators A Phys. 2016, 247, 355–367. [Google Scholar] [CrossRef]
  101. Zouaoui, M.J.; Nait-Ali, B.; Glandut, N.; Smith, D.S. Effect of humidity on the dielectric constant and electrical impedance of mesoporous zirconia ceramics. J. Eur. Ceram. Soc. 2016, 36, 163–169. [Google Scholar] [CrossRef]
  102. Soltani, R.; David, E.; Lamarre, L. Study on the effect of humidity on dielectric response and partial discharges activity of machine insulation materials. In Proceedings of the 2009 IEEE Electrical Insulation Conference, Montreal, QC, Canada, 31 May–3 June 2009; pp. 343–347. [Google Scholar]
  103. Lipare, A.Y.; Vasambekar, P.N.; Vaingankar, A.S. Dielectric behavior and a.c. resistivity study of humidity sensing ferrites. Mater. Chem. Phys. 2003, 81, 108–115. [Google Scholar] [CrossRef]
  104. Wolf, M.; Gulich, R.; Lunkenheimer, P.; Loidl, A. Relaxation dynamics of a protein solution investigated by dielectric spectroscopy. Biochim. Biophys. Acta (BBA)-Proteins Proteom. 2012, 1824, 723–730. [Google Scholar] [CrossRef]
  105. Yew, J.H.; Saha, T.K.; Thomas, A.J. Impact of temperature on the frequency domain dielectric spectroscopy for the diagnosis of power transformer insulation. In Proceedings of the 2006 IEEE Power Engineering Society General Meeting, Montreal, QC, Canada, 18–22 June 2006; p. 7. [Google Scholar]
  106. Paraskevas, C.D.; Vassiliou, P.; Dervos, C.T. Temperature dependent dielectric spectroscopy in frequency domain of high-voltage transformer oils compared to physicochemical results. IEEE Trans. Dielectr. Electr. Insul. 2005, 13, 539–546. [Google Scholar] [CrossRef]
  107. Petzelt, J.; Kozlov, G.V.; Volkov, A.A. Dielectric spectroscopy of paraelectric soft modes. Ferroelectrics 1987, 73, 101–123. [Google Scholar] [CrossRef]
  108. Nahm, S.H. Use of dielectric spectroscopy for real-time in-situ reaction monitoring. J. Coat. Technol. Res. 2006, 3, 257–265. [Google Scholar] [CrossRef]
  109. Diani, J.; Liu, Y.; Gall, K. Finite strain 3D thermoviscoelastic constitutive model for shape memory polymers. Polym. Eng. Sci. 2006, 46, 486–492. [Google Scholar] [CrossRef]
  110. Sorichetti, P.; Matteo, C.; Lambri, O.; Manguzzi, G.; Salvatierra, L.; Herrero, O. Structural changes in EPDM subjected to ageing in high voltage transmission lines. IEEE Trans. Dielectr. Electr. Insul. 2007, 14, 1170–1182. [Google Scholar] [CrossRef]
  111. Majumder, P.S.; Bhowmick, A.K. Structure-property relationship of electron-beam-modified EPDM rubber. J. Appl. Polym. Sci. 2000, 77, 323–337. [Google Scholar] [CrossRef]
  112. Erukhimovich, I.; de la Cruz, M.O. Phase equilibrium and charge fractionation in polyelectrolyte solutions. J. Polym. Sci. Part B Polym. Phys. 2007, 45, 3003–3009. [Google Scholar] [CrossRef]
  113. Nayak, S.; Chaki, T.K.; Khastgir, D. Dielectric relaxation and viscoelastic behavior of polyurethane–titania composites: Dielectric mixing models to explain experimental results. Polym. Bull. 2017, 74, 369–392. [Google Scholar] [CrossRef]
  114. Airinei, A.; Asandulesa, M.; Stelescu, M.D.; Tudorachi, N.; Fifere, N.; Bele, A.; Musteata, V. Dielectric, thermal and water absorption properties of some EPDM/Flax fiber composites. Polymers 2021, 13, 2555. [Google Scholar] [CrossRef]
  115. Min, D.; Li, S.; Hirai, N.; Ohki, Y. Dielectric spectroscopic analysis of degradation in ethylene-propylene-diene copolymer. IEEE Trans. Dielectr. Electr. Insul. 2016, 23, 3620–3630. [Google Scholar] [CrossRef]
  116. Schönhals, A.; Szymoniak, P. Dynamics of Composite Materials; Springer International Publishing: Cham, Switzerland, 2022. [Google Scholar]
  117. Eid, M.A.M.; El-Nashar, D.E. Filling effect of silica on electrical and mechanical properties of EPDM/NBR blends. Polym. Plast. Technol. Eng. 2006, 45, 675–684. [Google Scholar] [CrossRef]
  118. Zhang, F.; Zhao, Q.; Liu, T.; Lei, Y.; Chen, C. Preparation and relaxation dynamics of ethylene–propylene–diene rubber/clay nanocomposites with crosslinking interfacial design. J. Appl. Polym. Sci. 2018, 135, 45553. [Google Scholar] [CrossRef]
  119. Chailan, J.F.; Boiteux, G.; Chauchard, J.; Pinel, B.; Seytre, G. Viscoelastic and dielectric study of thermally aged ethylene—Propylene diene monomer (EPDM) compounds. Polym. Degrad. Stab. 1995, 47, 397–403. [Google Scholar] [CrossRef]
  120. Abd-El-Messieh, S.L.; El-Sabbagh, S.; Abadir, I.F. Dielectric relaxation and mechanical investigation of ethylene propylene diene monomer rubber with some crosslinking additives. J. Appl. Polym. Sci. 1999, 73, 1509–1519. [Google Scholar] [CrossRef]
  121. Dash, B.K.; Achary, P.G.R.; Nayak, N.C.; Choudhary, R.N.P. Dielectric relaxation behavior of exfoliated graphite nanoplatelet-filled EPDM vulcanizates. J. Electron. Mater. 2017, 46, 563–572. [Google Scholar] [CrossRef]
  122. Nayak, N.C.; Dash, B.K.; Parida, B.N.; Padhee, R. Dielectric relaxation behavior of exfoliated graphite nanoplatelets filled ethylene vinyl acetate copolymer and ethylene propylene diene terpolymer blend. J. Mater. Sci. Mater. Electron. 2018, 29, 1955–1963. [Google Scholar] [CrossRef]
  123. Barkoula, N.M.; Alcock, B.; Cabrera, N.O.; Peijs, T. Fatigue Properties of Highly Oriented Polypropylene Tapes and All-Polypropylene Composites. Polym. Polym. Compos 2008, 16, 101–113. [Google Scholar] [CrossRef]
  124. Mocellini, R.R.; Lambri, O.A.; Matteo, C.L.; Sorichetti, P.A. Dielectric properties and viscoelastic response in two-phase polymers. IEEE Trans. Dielectr. Electr. Insul. 2008, 15, 982–993. [Google Scholar] [CrossRef]
  125. Tiwari, A.; Sen, V.; Dhakate, S.R.; Mishra, A.P.; Singh, V. Synthesis, characterization, and hoping transport properties of HCl doped conducting biopolymer-co-polyaniline zwitterion hybrids. Polym. Adv. Technol. 2008, 19, 909–914. [Google Scholar] [CrossRef]
  126. Ismail, M.N.; Turky, G.M. Effect of fillers and vulcanizing systems on the physicomechanical and electrical properties of EPDM vulcanizates. Polym. Plast. Technol. Eng. 2001, 40, 635–652. [Google Scholar] [CrossRef]
  127. Younan, A.F.; Choneim, A.M.; Tawfik, A.A.A.; Abd-El-Nour, K.N. Electrical and physical properties of ethylene—Propylene—Diene monomer (EPDM) rubber loaded with semi-reinforcing furnace black. Polym. Degrad. Stab. 1995, 49, 215–222. [Google Scholar] [CrossRef]
  128. Celette, N.; Stevenson, I.; David, L.; Davenas, J.; Vigier, G.; Seytre, G. Irradiation effects on the relaxation behaviour of EPDM elastomers. Polym. Int. 2004, 53, 495–505. [Google Scholar] [CrossRef]
  129. Ajroldi, G.; Pianca, M.; Fumagalli, M.; Moggi, G. Fluoroelastomers-dependence of relaxation phenomena on composition. Polymer 1989, 30, 2180–2187. [Google Scholar] [CrossRef]
  130. Moni, G.; Jose, T.; Rajeevan, S.; Mayeen, A.; Rejimon, A.; Sarath, P.S.; George, S.C. Influence of exfoliated graphite inclusion on the thermal, mechanical, dielectric and solvent transport characteristics of fluoroelastomer nanocomposites. J. Polym. Res. 2020, 27, 72. [Google Scholar] [CrossRef]
  131. Moni, G.; Mayeen, A.; Mohan, A.; George, J.J.; Thomas, S.; George, S.C. Ionic liquid functionalised reduced graphene oxide fluoroelastomer nanocomposites with enhanced mechanical, dielectric and viscoelastic properties. Eur. Polym. J. 2018, 109, 277–287. [Google Scholar] [CrossRef]
  132. Balachandran, N.A.; Philip, K.; Rani, J. Effect of expanded graphite on thermal, mechanical and dielectric properties of ethylene–propylene–diene terpolymer/hexa fluoropropylene–vinylidinefluoride dipolymer rubber blends. Eur. Polym. J. 2013, 49, 247–260. [Google Scholar] [CrossRef]
  133. Starkweather, H.W.; Avakian, P.; Fontanella, J.J.; Wintersgill, M.C. Dielectric properties of polymers based on hexafluoropropylene. J. Therm. Anal. 1996, 46, 785–794. [Google Scholar] [CrossRef]
  134. Tamir, E.; Srebnik, S.; Sidess, A. Prediction of the relaxation modulus of a fluoroelastomer using molecular dynamics simulation. Chem. Eng. Sci. 2020, 225, 115786. [Google Scholar] [CrossRef]
  135. Zabelina, A.N.; Glushak, M.I.; Shcherbinin, A.S.; Khoroshavina, Y.V.; Ramsh, A.S.; Kurliand, S.K. Dielectric spectroscopy study of butadiene–nitrile rubbers. Russ. J. Phys. Chem. A 2020, 94, 119–124. [Google Scholar] [CrossRef]
  136. Gunasekaran, S.; Natarajan, R.; Kala, A.; Jagannathan, R. Dielectric studies of some rubber materials at microwave frequencies. Indian J. Pure Appl. Phys. 2008, 46, 733–737. [Google Scholar]
  137. Yang, D.; Tian, M.; Dong, Y.; Liu, H.; Yu, Y.; Zhang, L. Disclosed dielectric and electromechanical properties of hydrogenated nitrile–butadiene dielectric elastomer. Smart Mater. Struct. 2012, 21, 035017. [Google Scholar] [CrossRef]
  138. Adachi, H.; Adachi, K.; Kotaka, T. Effects of cross-linking density and of stretching on dielectric α-relaxation in poly(acrylonitrile-co-butadiene) rubber. Polym. J. 1980, 12, 329–334. [Google Scholar] [CrossRef]
  139. Böhm, M.; Soden, W.V.; Heinrich, W.; Yehia, A.A. Influence of crosslinking on mechanical and dielectric properties of nitrile-butadiene-rubber. Colloid Polym. Sci. 1987, 265, 295–303. [Google Scholar] [CrossRef]
  140. Garraza, A.L.R.; Sorichetti, P.; Marzocca, A.J.; Matteo, C.L.; Monti, G.A. Influence of the microstructure of vulcanized polybutadiene rubber on the dielectric properties. Polym. Test. 2011, 30, 657–662. [Google Scholar] [CrossRef]
  141. Yang, D.; Kong, X.; Ni, Y.; Gao, D.; Yang, B.; Zhu, Y.; Zhang, L. Novel nitrile-butadiene rubber composites with enhanced thermal conductivity and high dielectric constant. Compos. Part A Appl. Sci. Manuf. 2019, 124, 105447. [Google Scholar] [CrossRef]
  142. Sturniolo, S.; Pieruccini, M.; Corti, M.; Rigamonti, A. Probing α-relaxation with nuclear magnetic resonance echo decay and relaxation: A study on nitrile butadiene rubber. Solid State Nucl. Magn. Reson. 2013, 51–52, 16–24. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, C.; Zhou, G.; Zhu, W.; Chen, C.; Fu, Y.; Zhang, Z.; Li, H. Study of relaxations in epoxy/rubber composites by thermally stimulated depolarization current and dielectric spectroscopy. Front. Chem. 2022, 10, 874685. [Google Scholar] [CrossRef] [PubMed]
  144. Ardi, M.S.; Dick, W.; Kubát, J. Time-domain dielectric relaxation of poly(methyl methacrylate) and nitrile rubber. Colloid Polym. Sci. 1993, 271, 739–747. [Google Scholar] [CrossRef]
  145. Khalaf, A.I.; Ward, A.A.; El-Kader, A.E.A.; El-Sabbagh, S.H. Effect of selected vegetable oils on the properties of acrylonitrile-butadiene rubber vulcanizates. Polimery 2015, 60, 43–56. [Google Scholar] [CrossRef]
  146. Gaca, M.; Pietrasik, J.; Zaborski, M.; Okrasa, L.; Boiteux, G.; Gain, O. Effect of zinc oxide modified silica particles on the molecular dynamics of carboxylated acrylonitrile-butadiene rubber composites. Polymers 2017, 9, 645. [Google Scholar] [CrossRef]
  147. Roland, C.M. Electrical and dielectric properties of rubber. Rubber Chem. Technol. 2016, 89, 32–53. [Google Scholar] [CrossRef]
  148. Pietrasik, J.; Gaca, M.; Zaborski, M.; Okrasa, L.; Boiteux, G.; Gain, O. Studies of molecular dynamics of carboxylated acrylonitrile-butadiene rubber composites containing in situ synthesized silica particles. Eur. Polym. J. 2009, 45, 3317–3325. [Google Scholar] [CrossRef]
  149. Crossley, J. Dielectric relaxation of 1-alkenes. J. Chem. Phys. 1973, 58, 5315–5318. [Google Scholar] [CrossRef]
  150. Vo, L.T.; Anastasiadis, S.H.; Giannelis, E.P. Dielectric study of poly(styrene-co-butadiene) composites with carbon black, silica, and nanoclay. Macromolecules 2011, 44, 6162–6171. [Google Scholar] [CrossRef]
  151. Huang, M.; Tunnicliffe, L.B.; Thomas, A.G.; Busfield, J.J.C. The glass transition, segmental relaxations and viscoelastic behaviour of particulate-reinforced natural rubber. Eur. Polym. J. 2015, 67, 232–241. [Google Scholar] [CrossRef]
  152. Souillard, C.; Chazeau, L.; Cavaillé, J.-Y.; Brun, S.; Schach, R. On the β relaxations in poly(butadiene) and poly(styrene-butadiene) rubbers. Polymer 2019, 168, 236–245. [Google Scholar] [CrossRef]
  153. Pissis, P.; Fragiadakis, D.; Kanapitsas, A.; Delides, K. Broadband dielectric relaxation spectroscopy in polymer nanocomposites. Macromol. Symp. 2008, 265, 12–20. [Google Scholar] [CrossRef]
  154. Souillard, C.; Cavaillé, J.-Y.; Chazeau, L.; Schach, R. Dynamic mechanical relaxation of cross-linked styrene-butadiene polymers containing free chains: Possibility of reptation. Polymer 2014, 55, 5218–5225. [Google Scholar] [CrossRef]
  155. Abd-El-Messieh, S.L.; Abd-El-Nour, K.N. Effect of curing time and sulfur content on the dielectric relaxation of styrene butadiene rubber. J. Appl. Polym. Sci. 2003, 88, 1613–1621. [Google Scholar] [CrossRef]
  156. Xiang, K.; Wu, S.; Huang, G.; Zheng, J.; Huang, J.; Li, G. Relaxation behavior and time-temperature superposition (TTS) profiles of thermally aged styrene-butadiene rubber (SBR). Macromol. Res. 2014, 22, 820–825. [Google Scholar] [CrossRef]
  157. Wu, S.; Tang, Z.; Guo, B.; Zhang, L.; Jia, D. Effects of interfacial interaction on chain dynamics of rubber/graphene oxide hybrids: A dielectric relaxation spectroscopy study. RSC Adv. 2013, 3, 14549–14559. [Google Scholar] [CrossRef]
  158. Jungk, J.; Klueppel, M.; Meier, J.G. Dielectric relaxation spectroscopy of precipitated silica and elastomer-silica composites. KGK Rubberpoint 2009, 62, 319–325. [Google Scholar]
  159. Kumar, R.P.; Thomas, S. Dielectric properties of short sisal fibre reinforced SBR composites. Sci. Eng. Compos. Mater. 1999, 8, 311–326. [Google Scholar] [CrossRef]
  160. Hanna, F.F.; Yehia, A.A.; Abou-Bakr, A.F. Dielectric properties of styrene—Butadiene rubber/silicon dioxide mixtures. Br. Polym. J. 1973, 5, 83–90. [Google Scholar] [CrossRef]
  161. Gambino, T.; Alegría, A.; Arbe, A.; Colmenero, J.; Malicki, N.; Dronet, S. Modeling the high frequency mechanical relaxation of simplified industrial polymer mixtures using dielectric relaxation results. Polymer 2020, 187, 122051. [Google Scholar] [CrossRef]
  162. Ortega, L.; Cerveny, S.; Sill, C.; Isitman, N.A.; Rodriguez-Garraza, A.L.; Meyer, M.; Westermann, S.; Schwartz, G.A. The effect of vulcanization additives on the dielectric response of styrene-butadiene rubber compounds. Polymer 2019, 172, 205–212. [Google Scholar] [CrossRef]
  163. Ward, A. Dielectric and Mechanical Properties of Filled Rubbers in Dependence on Stress Amplitude and Temperature. Ph.D. Thesis, Cairo University, Giza, Egypt, 2003. [Google Scholar]
  164. Cerveny, S.; Bergman, R.; Schwartz, G.A.; Jacobsson, P. Dielectric α- and β-relaxations in uncured styrene butadiene rubber. Macromolecules 2002, 35, 4337–4342. [Google Scholar] [CrossRef]
  165. Renukappa, N.M.; Siddaramaiah; Sudhaker Samuel, R.D.; Sundara Rajan, J.; Lee, J.H. Dielectric properties of carbon black: SBR composites. J. Mater. Sci. Mater. Electron. 2009, 20, 648–656. [Google Scholar] [CrossRef]
  166. Janik, P.; Paluch, M.; Ziolo, J.; Sulkowski, W.; Nikiel, L. Low-frequency dielectric relaxation in rubber. Phys. Rev. E 2001, 64, 042502. [Google Scholar] [CrossRef]
  167. Lindemann, N.; Schawe, J.E.K.; Lacayo-Pineda, J. Kinetics of the glass transition of styrene-butadiene-rubber: Dielectric spectroscopy and fast differential scanning calorimetry. J. Appl. Polym. Sci. 2021, 138, 49769. [Google Scholar] [CrossRef]
  168. Moon, Y.I.; Jung, J.K.; Kim, G.H.; Chung, K.S. Observation of the relaxation process in fluoroelastomers by dielectric relaxation spectroscopy. Phys. B Condens. Matter 2021, 608, 412870. [Google Scholar] [CrossRef]
  169. Moon, Y.I.; Jung, J.K.; Chung, K.S. Dielectric relaxation spectroscopy in synthetic rubber polymers: Nitrile butadiene rubber and ethylene propylene diene monomer. Adv. Mater. Sci. Eng. 2020, 2020, 8406059. [Google Scholar] [CrossRef]
  170. Jung, J.K.; Moon, Y.I.; Kim, G.H.; Tak, N.H. Characterization of dielectric relaxation process by impedance spectroscopy for polymers: Nitrile butadiene rubber and ethylene propylene diene monomer. J. Spectrosc. 2020, 2020, 8815492. [Google Scholar] [CrossRef]
  171. Jung, J.K.; Moon, Y.I.; Chung, K.S.; Kim, K.-T. Development of a program for analyzing dielectric relaxation and its application to polymers: Nitrile butadiene rubber. Macromol. Res. 2020, 28, 596–604. [Google Scholar] [CrossRef]
  172. Jung, J.K.; Il Moon, Y.; Chung, K.S. Dielectric relaxation in a fluoroelastomer and ethylene propylene diene monomer observed by using impedance spectroscopy. J. Korean Phys. Soc. 2020, 76, 416–425. [Google Scholar] [CrossRef]
  173. Kim, G.-H.; Moon, Y.-I.; Jung, J.-K.; Choi, M.-C.; Bae, J.-W. Influence of carbon black and silica fillers with different concentrations on dielectric relaxation in nitrile butadiene rubber investigated by impedance spectroscopy. Polymers 2022, 14, 155. [Google Scholar] [CrossRef] [PubMed]
  174. Miles, P.A.; Westphal, W.B.; Von Hippel, A. Dielectric spectroscopy of ferromagnetic semiconductors. Rev. Mod. Phys. 1957, 29, 279–307. [Google Scholar] [CrossRef]
  175. Jonscher, A.K. Dielectric characterisation of semiconductors. Solid-State Electron. 1990, 33, 737–742. [Google Scholar] [CrossRef]
  176. Koltunowicz, T.N. Test station for frequency-domain dielectric spectroscopy of nanocomposites and semiconductors. J. Appl. Spectrosc. 2015, 82, 653–658. [Google Scholar] [CrossRef]
  177. Nobre, M.A.L.; Lanfredi, S. Dielectric spectroscopy on Bi3Zn2Sb3O14 ceramic: An approach based on the complex impedance. J. Phys. Chem. Solids 2003, 64, 2457–2464. [Google Scholar] [CrossRef]
  178. Vila, R.; González, M.; Mollá, J.; Ibarra, A. Dielectric spectroscopy of alumina ceramics over a wide frequency range. J. Nucl. Mater. 1998, 253, 141–148. [Google Scholar] [CrossRef]
  179. Ferreira, V.M.; Baptista, J.L.; Kamba, S.; Petzelt, J. Dielectric spectroscopy of MgTiO3-based ceramics in the 109–1014Hz region. J. Mater. Sci. 1993, 28, 5894–5900. [Google Scholar] [CrossRef]
  180. Krohns, S.; Lunkenheimer, P.; Ebbinghaus, S.G.; Loidl, A. Broadband dielectric spectroscopy on single-crystalline and ceramic CaCu3Ti4O12. Appl. Phys. Lett. 2007, 91, 022910. [Google Scholar] [CrossRef]
  181. Tsurumi, T.; Li, J.; Hoshina, T.; Kakemoto, H.; Nakada, M.; Akedo, J. Ultrawide range dielectric spectroscopy of BaTiO3-based perovskite dielectrics. Appl. Phys. Lett. 2007, 91, 182905. [Google Scholar] [CrossRef]
  182. Yassin, A.Y. Dielectric spectroscopy characterization of relaxation in composite based on (PVA–PVP) blend for nickel–cadmium batteries. J. Mater. Sci. Mater. Electron. 2020, 31, 19447–19463. [Google Scholar] [CrossRef]
  183. Badot, J.-C.; Lestriez, B.; Dubrunfaut, O. Interest in broadband dielectric spectroscopy to study the electronic transport in materials for lithium batteries. Mater. Sci. Eng. B 2016, 213, 190–198. [Google Scholar] [CrossRef]
  184. Di Noto, V.; Giffin, G.A.; Vezzù, K.; Piga, M.; Lavina, S. Broadband dielectric spectroscopy: A powerful tool for the determination of charge transfer mechanisms in ion conductors. In Solid State Proton Conductors: Properties and Applications in Fuel Cells; Knauth, P., Vona, M.L.D., Eds.; Wiley: West Sussex, UK, 2012; pp. 109–183. [Google Scholar]
  185. Hanafusa, H.; Sato, Y. Dielectric properties of polymer-TiO2 composites. Kobunshi Ronbunshu 1978, 35, 277–282. [Google Scholar] [CrossRef]
  186. Michl, M.; Bauer, T.; Lunkenheimer, P.; Loidl, A. Nonlinear dielectric spectroscopy in a fragile plastic crystal. J. Chem. Phys. 2016, 144, 114506. [Google Scholar] [CrossRef]
  187. Hedvig, P.; Czvikovszky, T. Dielectric spectroscopic study of wood plastic combinations. Angew. Makromol. Chem. 1972, 21, 79–85. [Google Scholar] [CrossRef]
  188. Strobl, G.R. The Physics of Polymers: Concepts for Understanding Their Structures and Behavior; Springer: Berlin/Heidelberg, Germany, 1997; pp. 143–190. [Google Scholar]
  189. Rosen, S.L. Fundamental Principles of Polymeric Materials; Wiley: New York, NY, USA, 1982. [Google Scholar]
  190. Karak, N. Fundamentals of Polymers: Raw Materials to Finish Products; PHI Learning Pvt Ltd.: New Delhi, India, 2009; pp. 1–30. [Google Scholar]
  191. Watanabe, H. Dielectric relaxation of type-A polymers in melts and solutions. Macromol. Rapid Commun. 2001, 22, 127–175. [Google Scholar] [CrossRef]
  192. Nikonorova, N.A.; Yakimansky, A.V.; Smirnov, N.N.; Kudryavtsev, V.V.; Diaz-Calleja, R.; Pissis, P. Dielectric relaxation in copolymethacrylates containing side-chain nonlinear optical chromophores. Polymer 2007, 48, 556–563. [Google Scholar] [CrossRef]
  193. Ngai, K.L.; Roland, C.M. Chemical structure and intermolecular cooperativity: Dielectric relaxation results. Macromolecules 1993, 26, 6824–6830. [Google Scholar] [CrossRef]
  194. North, A.M. Dielectric relaxation in polymer solutions. Chem. Soc. Rev. 1972, 1, 49–72. [Google Scholar] [CrossRef]
  195. Hammami, H.; Arous, M.; Lagache, M.; Kallel, A. Study of the interfacial MWS relaxation by dielectric spectroscopy in unidirectional PZT fibres/epoxy resin composites. J. Alloys Compd. 2007, 430, 1–8. [Google Scholar] [CrossRef]
  196. Xu, P.; Zhang, X. Investigation of MWS polarization and dc conductivity in polyamide 610 using dielectric relaxation spectroscopy. Eur. Polym. J. 2011, 47, 1031–1038. [Google Scholar] [CrossRef]
  197. Chanmal, C.; Jog, J. Dielectric relaxation spectroscopy for polymer nanocomposites. In Characterization Techniques for Polymer Nanocomposites; Mittal, V., Ed.; Wiley: Weinheim, Germany, 2012; pp. 167–184. [Google Scholar]
  198. Mijović, J.; Lee, H.; Kenny, J.; Mays, J. Dynamics in polymer−silicate nanocomposites as studied by dielectric relaxation spectroscopy and dynamic mechanical spectroscopy. Macromolecules 2006, 39, 2172–2182. [Google Scholar] [CrossRef]
  199. Kremer, F.; Schönhals, A. Broadband Dielectric Spectroscopy; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  200. Runt, J.; Fitzgerald, J.J. Dielectric Spectroscopy of Polymeric Materials: Fundamentals and Applications; American Chemical Society: Washington, DC, USA, 1997. [Google Scholar]
  201. Vassilikou-Dova, A.; Kalogeras, I.M. Dielectric analysis (DEA). In Thermal Analysis of Polymers: Fundamentals and Applications; Menczel, J.D., Prime, R.B., Eds.; Wiley: Hoboken, NJ, USA, 2009; pp. 497–614. [Google Scholar]
  202. Stockmayer, W.H. Dielectric dispersion in solutions of flexible polymers. Pure Appl. Chem. 1967, 15, 539–554. [Google Scholar] [CrossRef]
  203. Block, H. The nature and application of electrical phenomena in polymers. In Electric Phenomena in Polymer Science. Advances in Polymer Science; Abe, A., Albertsson, A.-C., Dusek, K., Genzer, J., Kobayashi, S., Lee, K.-S., Leibler, L., Long, T.E., Manners, I., Möller, M., et al., Eds.; Springer: Berlin/Heidelberg, Germany, 1979; pp. 93–167. [Google Scholar]
  204. Van Krevelen, D.W.; Te Nijenhuis, K. Properties of Polymers: Their Correlation with Chemical Structure; Their Numerical Estimation and Prediction from Additive Group Contributions; Elsevier: Oxford, UK, 2009; pp. 779–786. [Google Scholar]
  205. Kakani, S.L. Electronics Theory and Applications; New Age International (P) Ltd.: New Delhi, India, 2005. [Google Scholar]
  206. Cole, K.S.; Cole, R.H. Dispersion and absorption in dielectrics I. Alternating current characteristics. J. Chem. Phys. 1941, 9, 341–351. [Google Scholar] [CrossRef]
  207. Davidson, D.W.; Cole, R.H. Dielectric relaxation in glycerine. J. Chem. Phys. 1950, 18, 1417. [Google Scholar] [CrossRef]
  208. Havriliak, S.; Negami, S. A complex plane representation of dielectric and mechanical relaxation processes in some polymers. Polymer 1967, 8, 161–210. [Google Scholar] [CrossRef]
  209. Gerstl, C.; Schneider, G.J.; Pyckhout-Hintzen, W.; Allgaier, J.; Richter, D.; Alegría, A.; Colmenero, J. Segmental and normal mode relaxation of poly(alkylene oxide)s studied by dielectric spectroscopy and rheology. Macromolecules 2010, 43, 4968–4977. [Google Scholar] [CrossRef]
  210. Roland, C.M.; Bero, C.A. Normal mode relaxation in linear and branched polyisoprene. Macromolecules 1996, 29, 7521–7526. [Google Scholar] [CrossRef]
  211. Kuttich, B.; Lederle, C.; Stühn, B. Water dependence of the dielectric β-relaxation in poly(ɛ-caprolactone). J. Chem. Phys. 2013, 139, 244907. [Google Scholar] [CrossRef]
  212. Priestley, R.D.; Broadbelt, L.J.; Torkelson, J.M.; Fukao, K. Glass transition and α-relaxation dynamics of thin films of labeled polystyrene. Phys. Rev. E 2007, 75, 061806. [Google Scholar] [CrossRef]
  213. Pissis, P.; Fragiadakis, D. Dielectric studies of segmental dynamics in epoxy nanocomposites. J. Macromol. Sci. Part B 2007, 46, 119–136. [Google Scholar] [CrossRef]
  214. Dunn, A.M.; Hofmann, O.S.; Waters, B.; Witchel, E. Cloaking malware with the trusted platform module. In Proceedings of the 20th USENIX Conference on Security (SEC’11), San Francisco, CA, USA, 8–12 August 2011; USENIX Association: Berkeley, CA, USA, 2011; pp. 395–410. [Google Scholar]
  215. Kyriakos, K.; Raftopoulos, K.N.; Pissis, P.; Kyritsis, A.; Näther, F.; Häußler, L.; Fischer, D.; Vyalikh, A.; Scheler, U.; Reuter, U.; et al. Dielectric and thermal studies of the segmental dynamics of poly(methyl methacrylate)/silica nanocomposites prepared by the sol–gel method. J. Appl. Polym. Sci. 2013, 128, 3771–3781. [Google Scholar] [CrossRef]
  216. Wübbenhorst, M.; van Koten, E.M.; Jansen, J.C.; Mijs, W.; van Turnhout, J. Dielectric relaxation spectroscopy of amorphous and liquid-crystalline side-chain polycarbonates. Macromol. Rapid Commun. 1997, 18, 139–147. [Google Scholar] [CrossRef]
  217. Atawa, B.; Correia, N.T.; Couvrat, N.; Affouard, F.; Coquerel, G.; Dargent, E.; Saiter, A. Molecular mobility of amorphous N-acetyl-α-methylbenzylamine and Debye relaxation evidenced by dielectric relaxation spectroscopy and molecular dynamics simulations. Phys. Chem. Chem. Phys. 2019, 21, 702–717. [Google Scholar] [CrossRef]
  218. Vallerien, S.U.; Kremer, F.; Boeffel, C. Broadband dielectric spectroscopy on side group liquid crystal polymers. Liq. Cryst. 1989, 4, 79–86. [Google Scholar] [CrossRef]
  219. Vikulova, M.; Tsyganov, A.; Bainyashev, A.; Artyukhov, D.; Gorokhovsky, A.; Muratov, D.; Gorshkov, N. Dielectric properties of PMMA/KCTO(H) composites for electronics components. J. Appl. Polym. Sci. 2021, 138, 51168. [Google Scholar] [CrossRef]
  220. Yehia, A.A. Conductivity studies on acrylonitrile butadiene rubber loaded with different types of carbon blacks. Int. J. Mater. Methods Technol. 2013, 1, 22–33. [Google Scholar]
  221. Hiltz, J.A. Characterization of fluoroelastomers by various analytical techniques including pyrolysis gas chromatography/mass spectrometry. J. Anal. Appl. Pyrolysis 2014, 109, 283–295. [Google Scholar] [CrossRef]
  222. Yeo, Y.-G.; Park, H.-H.; Lee, C.-S. A study on the characteristics of a rubber blend of fluorocarbon rubber and hydrogenated nitrile rubber. J. Ind. Eng. Chem. 2013, 19, 1540–1548. [Google Scholar] [CrossRef]
  223. De Angelis, M.G.; Sarti, G.C.; Sanguineti, A.; Maccone, P. Permeation, diffusion, and sorption of dimethyl ether in fluoroelastomers. J. Polym. Sci. Part B Polym. Phys. 2004, 42, 1987–2006. [Google Scholar] [CrossRef]
  224. Kader, M.A.; Bhowmick, A.K. Acrylic rubber-fluorocarbon rubber miscible blends: Effect of curatives and fillers on cure, mechanical, aging, and swelling properties. J. Appl. Polym. Sci. 2003, 89, 1442–1452. [Google Scholar] [CrossRef]
  225. George, S.; Sushama, C.M.; Nair, A.B. Efficiencies of dipolymer rubber blends (EPDM\FKM) using common weight data envelopment analysis. Mater. Res. 2017, 20, 1722–1728. [Google Scholar] [CrossRef]
  226. Koizumi, N.; Tsunashima, K.; Yano, S. Dielectric relaxation in copolymers of vinylidene fluoride and hexafluoropropylene. J. Polym. Sci. Part B Polym. Lett. 1969, 7, 815–820. [Google Scholar] [CrossRef]
  227. Kochervinskii, V.V.; Malyshkina, I.A.; Markin, G.V.; Gavrilova, N.D.; Bessonova, N.P. Dielectric relaxation in vinylidene fluoride–hexafluoropropylene copolymers. J. Appl. Polym. Sci. 2007, 105, 1101–1117. [Google Scholar] [CrossRef]
  228. Axelrod, N.; Axelrod, E.; Gutina, A.; Puzenko, A.; Ishai, P.B.; Feldman, Y. Dielectric spectroscopy data treatment: I. Fre-quency domain. Meas. Sci. Technol. 2004, 15, 755–764. [Google Scholar] [CrossRef]
  229. Nelder, J.A.; Mead, R. A simplex method for function minimization. Comput. J. 1965, 7, 308–313. [Google Scholar] [CrossRef]
Figure 1. Configuration of automatic dielectric spectroscopy system for measuring complex permittivity in polymer specimens.
Figure 1. Configuration of automatic dielectric spectroscopy system for measuring complex permittivity in polymer specimens.
Polymers 17 01539 g001
Figure 2. Screenshot images of commercially available data analysis software for determining dielectric dispersion spectra. (a) “EIS-smart-tool” (source: https://github.com/bmgmelo/EIS-smart-tool, accessed on 1 February 2025) and the different colors of the empty circles and solid lines indicate the measured data and their respective fitting results at each temperature; (b) “WinFit”, Novocontrol Technologies (source: https://www.novocontrol.de/php/winfit.php, accessed on 1 February 2025).
Figure 2. Screenshot images of commercially available data analysis software for determining dielectric dispersion spectra. (a) “EIS-smart-tool” (source: https://github.com/bmgmelo/EIS-smart-tool, accessed on 1 February 2025) and the different colors of the empty circles and solid lines indicate the measured data and their respective fitting results at each temperature; (b) “WinFit”, Novocontrol Technologies (source: https://www.novocontrol.de/php/winfit.php, accessed on 1 February 2025).
Polymers 17 01539 g002
Figure 3. A screenshot showing the execution of the integrated deconvolution analysis program. One typical dataset is loaded as an example, where a 3D graph of complex dielectric constants as a function of frequency and temperature appears in the upper left panel. The red and blue sheets represent the real and imaginary parts of the complex permittivity, respectively. The x- and y-axes represent frequency and temperature, while the double z-axes in Figure 3 represent the real (left) and imaginary (right) permittivity values. In the right half of the main panel, the measured and fitted spectra at a specific temperature are shown. The red and blue cross symbols indicate the real and imaginary values of the measured permittivity, respectively, while the solid lines represent the fitted results for the corresponding colored data using a combination of two HN functions. The ten fitted parameters are listed in the bottom right panel. The fitting accuracy value (1.3%) is shown below the current temperature (29.98 °C), and the details of its calculation procedure are described in the Supplementary Materials.
Figure 3. A screenshot showing the execution of the integrated deconvolution analysis program. One typical dataset is loaded as an example, where a 3D graph of complex dielectric constants as a function of frequency and temperature appears in the upper left panel. The red and blue sheets represent the real and imaginary parts of the complex permittivity, respectively. The x- and y-axes represent frequency and temperature, while the double z-axes in Figure 3 represent the real (left) and imaginary (right) permittivity values. In the right half of the main panel, the measured and fitted spectra at a specific temperature are shown. The red and blue cross symbols indicate the real and imaginary values of the measured permittivity, respectively, while the solid lines represent the fitted results for the corresponding colored data using a combination of two HN functions. The ten fitted parameters are listed in the bottom right panel. The fitting accuracy value (1.3%) is shown below the current temperature (29.98 °C), and the details of its calculation procedure are described in the Supplementary Materials.
Polymers 17 01539 g003
Figure 4. Molecular structure of NBR.
Figure 4. Molecular structure of NBR.
Polymers 17 01539 g004
Figure 5. (a) A three-dimensional plot of the imaginary permittivity of an NBR; (b) the temperature ranges observed for each relaxation process presented as a bar chart.
Figure 5. (a) A three-dimensional plot of the imaginary permittivity of an NBR; (b) the temperature ranges observed for each relaxation process presented as a bar chart.
Polymers 17 01539 g005
Figure 6. Deconvoluted relaxation process of NBR for seven representative temperatures: (a) 404 K, (b) 359 K, (c) 326 K, (d) 293 K, (e) 272 K, (f) 251 K, and (g) 233 K. Contributions made by five relaxation processes—α’ (black dashed line), α (red dashed line), β (blue dashed line), γ (gray dashed line), and conduction (navy dashed line)—to imaginary permittivity data. Black, red, blue, and gray arrows indicate peak position of corresponding α’, α, β, and γ relaxations, respectively.
Figure 6. Deconvoluted relaxation process of NBR for seven representative temperatures: (a) 404 K, (b) 359 K, (c) 326 K, (d) 293 K, (e) 272 K, (f) 251 K, and (g) 233 K. Contributions made by five relaxation processes—α’ (black dashed line), α (red dashed line), β (blue dashed line), γ (gray dashed line), and conduction (navy dashed line)—to imaginary permittivity data. Black, red, blue, and gray arrows indicate peak position of corresponding α’, α, β, and γ relaxations, respectively.
Polymers 17 01539 g006
Figure 7. Assignment of dielectric relaxation process for schematic molecular chain structure and its motion mode in NBR.
Figure 7. Assignment of dielectric relaxation process for schematic molecular chain structure and its motion mode in NBR.
Polymers 17 01539 g007
Figure 8. The reciprocal temperature dependence of the central frequency ( f 0 ) of the imaginary loss peak for the α’ (black circle), α (red circle), β (blue circle), γ (gray circle), and conductivity (navy circle) processes in NBR. Solid lines represent the fitted curves for each relaxation mode. Furthermore, α’ was also fitted using the VFTH model, yielding fitting parameters of f p , = 2.8 × 107 Hz and A = 0.09 eV and a Vogel temperature of 223.239 K.
Figure 8. The reciprocal temperature dependence of the central frequency ( f 0 ) of the imaginary loss peak for the α’ (black circle), α (red circle), β (blue circle), γ (gray circle), and conductivity (navy circle) processes in NBR. Solid lines represent the fitted curves for each relaxation mode. Furthermore, α’ was also fitted using the VFTH model, yielding fitting parameters of f p , = 2.8 × 107 Hz and A = 0.09 eV and a Vogel temperature of 223.239 K.
Polymers 17 01539 g008
Figure 9. (a) The glass transition temperature determined via dielectric spectroscopy, with the dashed line representing the fitted result of the VFTH equation. (b) The glass transition temperature determined via DSC.
Figure 9. (a) The glass transition temperature determined via dielectric spectroscopy, with the dashed line representing the fitted result of the VFTH equation. (b) The glass transition temperature determined via DSC.
Polymers 17 01539 g009
Figure 10. Molecular structure of EPDM.
Figure 10. Molecular structure of EPDM.
Polymers 17 01539 g010
Figure 11. (a) A 3D plot of the imaginary part of complex permittivity for EPDM. (b) The temperature ranges for the observed relaxation phenomena presented as a bar chart.
Figure 11. (a) A 3D plot of the imaginary part of complex permittivity for EPDM. (b) The temperature ranges for the observed relaxation phenomena presented as a bar chart.
Polymers 17 01539 g011
Figure 12. Deconvoluted relaxation process of EPDM for seven representative temperatures: (a) 404 K, (b) 359 K, (c) 326 K, (d) 293 K, (e) 272 K, (f) 251 K, and (g) 233 K. Contributions made by all three relaxation processes—MWS (orange dashed line), α (red dashed line), and β (blue dashed line)—and conduction (navy dashed line) to imaginary permittivity data. Orange, red, and blue arrows indicate peak position of corresponding MWS, α, and β relaxations, respectively.
Figure 12. Deconvoluted relaxation process of EPDM for seven representative temperatures: (a) 404 K, (b) 359 K, (c) 326 K, (d) 293 K, (e) 272 K, (f) 251 K, and (g) 233 K. Contributions made by all three relaxation processes—MWS (orange dashed line), α (red dashed line), and β (blue dashed line)—and conduction (navy dashed line) to imaginary permittivity data. Orange, red, and blue arrows indicate peak position of corresponding MWS, α, and β relaxations, respectively.
Polymers 17 01539 g012
Figure 13. Schematic assignment for molecular structure of EPDM. Each reorientational relaxation is shown with arrows.
Figure 13. Schematic assignment for molecular structure of EPDM. Each reorientational relaxation is shown with arrows.
Polymers 17 01539 g013
Figure 14. The relaxation rate (f₀) as a function of the reciprocal temperature (1000/T) for EPDM. Activation energies were extracted from the slopes of the fitted solid lines. The overlapping circle symbols for the MWS and α relaxation processes indicate repeated measurements. The inset displays the temperature dependence of the relaxation strength (Δε) for each process.
Figure 14. The relaxation rate (f₀) as a function of the reciprocal temperature (1000/T) for EPDM. Activation energies were extracted from the slopes of the fitted solid lines. The overlapping circle symbols for the MWS and α relaxation processes indicate repeated measurements. The inset displays the temperature dependence of the relaxation strength (Δε) for each process.
Polymers 17 01539 g014
Figure 15. (a) The glass transition temperature determined using dielectric spectroscopy, with the dashed line representing the fitted result of the VFTH equation. (b) The glass transition temperature determined using DSC.
Figure 15. (a) The glass transition temperature determined using dielectric spectroscopy, with the dashed line representing the fitted result of the VFTH equation. (b) The glass transition temperature determined using DSC.
Polymers 17 01539 g015
Figure 16. Molecular structure of FKM.
Figure 16. Molecular structure of FKM.
Polymers 17 01539 g016
Figure 17. (a) A three-dimensional plot of the imaginary permittivity of an FKM. (b) The temperature ranges for the observed relaxation phenomena presented as a bar chart.
Figure 17. (a) A three-dimensional plot of the imaginary permittivity of an FKM. (b) The temperature ranges for the observed relaxation phenomena presented as a bar chart.
Polymers 17 01539 g017
Figure 18. The deconvoluted relaxation process of FKM for several representative temperatures. Both analyzing processes use the combination of HN functions and the power law of the DC conductivity contribution to fit the spectra. All three main relaxations—MWS (orange dashed lines), α (red dashed lines), and β (blue dashed lines)—and conductivity (navy dashed line) were observed in FKM compounds. The arrows indicate the peak positions of the corresponding MWS, α, and β relaxations, respectively.
Figure 18. The deconvoluted relaxation process of FKM for several representative temperatures. Both analyzing processes use the combination of HN functions and the power law of the DC conductivity contribution to fit the spectra. All three main relaxations—MWS (orange dashed lines), α (red dashed lines), and β (blue dashed lines)—and conductivity (navy dashed line) were observed in FKM compounds. The arrows indicate the peak positions of the corresponding MWS, α, and β relaxations, respectively.
Polymers 17 01539 g018
Figure 19. Assignments of the dielectric relaxation processes (α and β) for a schematic molecular chain structure and its motional mode in FKM.
Figure 19. Assignments of the dielectric relaxation processes (α and β) for a schematic molecular chain structure and its motional mode in FKM.
Polymers 17 01539 g019
Figure 20. The reciprocal temperature dependence of the central frequency ( f 0 ) of the imaginary loss peak for the α (red circle), β (blue circle), α + β (pink circle), MWS (orange circle), and conductivity (navy circle) processes in FKM. The activation energy for the relaxation process was the slope of the fitted corresponding solid line.
Figure 20. The reciprocal temperature dependence of the central frequency ( f 0 ) of the imaginary loss peak for the α (red circle), β (blue circle), α + β (pink circle), MWS (orange circle), and conductivity (navy circle) processes in FKM. The activation energy for the relaxation process was the slope of the fitted corresponding solid line.
Polymers 17 01539 g020
Figure 21. (a) The glass transition temperature determined using dielectric spectroscopy, with the dashed line representing the fitted result of the VFTH equation. (b) The glass transition temperature determined via DSC.
Figure 21. (a) The glass transition temperature determined using dielectric spectroscopy, with the dashed line representing the fitted result of the VFTH equation. (b) The glass transition temperature determined via DSC.
Polymers 17 01539 g021
Table 1. The chemical composition of the NBR compound.
Table 1. The chemical composition of the NBR compound.
Chemical NameFunction(%)
Acrylonitrile–butadiene rubberPolymer40.0
Carbon black (medium thermal)Filler reinforcing50.0
1,2-Benzenedicarboxylic acidProcessing aid6.0
2-BenzimidazolethiolAntioxidant2.0
SulfurCuring agent2.0
Total 100.0
Table 2. Chemical composition of EPDM rubber compound.
Table 2. Chemical composition of EPDM rubber compound.
Chemical NameFunction(%)
Ethylene propylene rubberPolymer58.0
Carbon blackFiller reinforcement34.0
Zinc oxideProcessing aid3.0
Dicumyl peroxideAntioxidant5.0
Total 100.0
Table 3. Chemical composition of FKM rubber compound.
Table 3. Chemical composition of FKM rubber compound.
Chemical NameFunction(%)
Poly (vinylidene fluoride-co-hexafluoropropylene)Polymer82.0
Carbon blackFiller reinforcement14.0
Calcium dihydroxideCuring agent4.0
Total 100.0
Table 4. Comparison of activation energy values for β and MWS relaxations in FKM measured using dielectric spectroscopy.
Table 4. Comparison of activation energy values for β and MWS relaxations in FKM measured using dielectric spectroscopy.
ReferenceOurs[172][226][133][227]
Ea for β (kJ/mol)50.8863.550.2, 54.4 *4749.8, 40.2
Ea for MWS (kJ/mol)88.57110 151, 81 **
* Referred to as γ relaxation. ** Referred to as α relaxation.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Moon, Y.; Kim, G.; Jung, J. A Study of the Dielectric Relaxation of Nitrile–Butadiene Rubber, Ethylene–Propylene–Diene Monomer, and Fluoroelastomer Polymers with a Self-Developed Deconvolution Analysis Program. Polymers 2025, 17, 1539. https://doi.org/10.3390/polym17111539

AMA Style

Moon Y, Kim G, Jung J. A Study of the Dielectric Relaxation of Nitrile–Butadiene Rubber, Ethylene–Propylene–Diene Monomer, and Fluoroelastomer Polymers with a Self-Developed Deconvolution Analysis Program. Polymers. 2025; 17(11):1539. https://doi.org/10.3390/polym17111539

Chicago/Turabian Style

Moon, Youngil, Gyunghyun Kim, and Jaekap Jung. 2025. "A Study of the Dielectric Relaxation of Nitrile–Butadiene Rubber, Ethylene–Propylene–Diene Monomer, and Fluoroelastomer Polymers with a Self-Developed Deconvolution Analysis Program" Polymers 17, no. 11: 1539. https://doi.org/10.3390/polym17111539

APA Style

Moon, Y., Kim, G., & Jung, J. (2025). A Study of the Dielectric Relaxation of Nitrile–Butadiene Rubber, Ethylene–Propylene–Diene Monomer, and Fluoroelastomer Polymers with a Self-Developed Deconvolution Analysis Program. Polymers, 17(11), 1539. https://doi.org/10.3390/polym17111539

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

Article Metrics

Back to TopTop