Next Article in Journal
Optimization of Chromium Removal Conditions from Tanned Leather Waste for Collagen Valorization
Previous Article in Journal
Potential of Compost-Derived Actinomycetes for Low-Density Polyethylene Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design-Oriented Degradation Mapping and Hyperelastic Model-Switch Guidelines for Nitrile-Butadiene Rubber Seals

1
Digital Twin Laboratory, Dong-Eui University, 176 Eomgwang-ro, Busan 47340, Republic of Korea
2
Center for Brain Busan 21, Dong-Eui University, 176 Eomgwang-ro, Busan 47340, Republic of Korea
3
Korea Atomic Energy Research Institute, Gyeongju-si 38180, Republic of Korea
*
Authors to whom correspondence should be addressed.
Polymers 2025, 17(17), 2316; https://doi.org/10.3390/polym17172316
Submission received: 30 June 2025 / Revised: 15 August 2025 / Accepted: 21 August 2025 / Published: 27 August 2025
(This article belongs to the Special Issue Mechanical and Durability Properties of Polymer Materials)

Abstract

Nitrile-butadiene rubber (NBR) seals used in automotive and energy equipment undergo pronounced mechanical degradation at elevated temperatures, yet a quantitative rule for switching between hyperelastic models remains unclear. Here, accelerated thermal aging tests were linked to service conditions by estimating the activation energy via Flynn–Wall–Ozawa analysis and applying an Arrhenius-based equivalence. Tensile testing, dynamic mechanical analysis, and thermogravimetric analysis were combined to track embrittlement and crosslinking, and finite element simulations were benchmarked against experiments using an L2-norm metric. The outcome is a degradation map with a model-switching guideline. The Neo-Hookean model is preferred in the less-embrittled regime, whereas the five-parameter Mooney–Rivlin model is recommended as embrittlement progresses. This framework improves stress-prediction fidelity while keeping model complexity commensurate with the aging state, enabling faster and more reliable design of NBR seals for high-temperature automotive and renewable-energy applications.

1. Introduction

Elastomeric materials, commonly known as rubbers, are indispensable in modern engineering due to their unique properties, such as high elasticity, resilience, and durability [1]. Their applications are vast, ranging from automotive tires and seals, where materials like Styrene-Butadiene Rubber (SBR) and Ethylene Propylene Diene Monomer (EPDM) are common, to advanced aerospace and biomedical devices. Among these versatile materials, Nitrile-Butadiene Rubber (NBR) is widely used as a key sealing material to prevent gas or oil leakage in high-temperature, high-pressure moving parts due to its excellent oil and abrasion resistance [2,3].
However, long-term exposure of NBR to environments with a combination of oxygen, heat, mechanical load, and chemicals accelerates chain scission, further cross-linking, and oxidative degradation, causing a rapid deterioration of its mechanical properties and durability [4,5,6]. As this degradation can lead to the failure of the entire system through pressure loss and lubricant depletion [7], predicting degradation and incorporating it into design is essential.
Given the extremely slow rate of thermal aging in rubber under real-world conditions, significant research has been dedicated to simulating equivalent states of degradation using accelerated aging tests [8,9]. Existing literature can be broadly categorized into three streams: (1) Studies of material degradation mechanisms focusing on changes in chemical structure and crosslink density [10,11,12,13,14,15]; (2) reliability engineering approaches that use accelerated tests to quantify long-term performance and predict service life [16,17,18,19,20,21]; and (3) simulation approaches that seek to enhance the accuracy of Finite Element Analysis (FEA) by fitting material model parameters to experimental data at various stages of degradation [22,23,24,25,26].
Although each research area holds significant merit, a critical gap persists: insights gained at the molecular level are not directly integrated into the selection process for engineering design models. First, molecular or network metrics are characterized but rarely translated into a quantitative rule that selects among constitutive models. Second, accelerated-aging studies establish time–temperature relations yet do not prescribe how specific service conditions should determine model choice. Third, model comparisons are commonly performed at discrete aging states with local fit metrics and limited uncertainty reporting, leaving no clear boundary for a model switch. Consequently, in practical applications, the choice of an optimal model for a given level of degradation often relies on empirical judgment rather than a systematic guideline.
To address this gap, we introduce and experimentally validate an end-to-end workflow that maps service time–temperature inputs to a quantitative rule for model selection in aged NBR. Initially, the activation energy derived from TGA-based Flynn–Wall–Ozawa analysis was input into the Arrhenius equation. This allowed us to design a predictable accelerated aging protocol at 150 °C that simulates equivalent service conditions at 70, 80, and 90 °C for periods of 1, 3, and 5 years. Following this, the crosslink density and elongation at break for each degradation stage were quantitatively determined through tensile, DMA, and TGA tests. These experimental data were then used to reverse-engineer the parameters for the Neo-Hookean, Mooney–Rivlin, and 2nd-order Yeoh hyperelastic models. The accuracy of each model was then evaluated by an L2-norm comparison.
Our findings verify that the Neo-Hookean model provides the best fit during the initial degradation phases, whereas the Mooney–Rivlin model excels in accuracy beyond the brittle transition point. Building on these findings, we derived a practical design guideline presented as a contour map, which allows for the selection of the most suitable material model based on two key criteria: crosslink density and strain at maximum tensile strength.
The overall workflow of this research is outlined in Figure 1. The structure of this paper is as follows: Section 2 reviews the Arrhenius-based theory for lifetime prediction; Section 3 presents the results from accelerated aging tests and material characterization; Section 4 details the suitability analysis of each model; Section 5 establishes the degradation boundaries and model transition criteria and discusses the broader engineering implications. Ultimately, this work intends to offer practical guidance for designing NBR seals for automotive and renewable energy applications that operate in high-temperature and high-pressure conditions.

2. Determination of Equivalent Service Life of NBR Using Activation Energy

The long-term evolution of mechanical properties and the lifetime of rubber materials can be determined via accelerated aging experiments founded on kinetic models, which leverage the laws governing time- and temperature-dependent property changes [27,28]. In this work, the Arrhenius equation was employed to determine the required duration for accelerated aging tests by establishing a conversion relationship between actual and accelerated service conditions.
As the long-term degradation of rubber follows a temperature-dependent reaction rate, it can be modeled using the Arrhenius equation [29,30]. The Arrhenius equation is given as follows in Equation (1).
k = A e x p ( E a R T )
where k represents the chemical reaction rate, E a is the activation energy (kJ mol−1), R is the universal gas constant, T denotes the absolute temperature (K), and A is the pre-exponential factor. A key prerequisite for calculating the accelerated ageing duration for a target service life with the Arrhenius equation is determining the material’s characteristic activation energy. To this end, we conducted thermogravimetric analysis (TGA) on NBR specimens and applied the isoconversional Flynn–Wall–Ozawa (FWO) method using multiple heating rates (5, 10, 20, and 50 °C/min) [31,32,33]. The FWO equation is given as follows in Equation (2).
l o g F α = log A E a R l o g β 2.315 0.4567 E a R T
where F α is the integral form of the conversion function, A is the pre-exponential factor, E a is the activation energy (kJ mol−1), R is the universal gas constant (8.314 J mol−1 K−1), β is the constant heating rate (°C/min), and T is the absolute temperature (K). The numerical constants, −2.315 and −0.4567, are empirically derived coefficients from Doyle’s approximation of the temperature integral, which is a key step in the derivation of the FWO method [34].
The activation energy, E a , was determined from the slope of the regression line on a plot of ln(β) against 1/T. The coefficients of determination (R2) for the regression lines at four distinct conversion rates (α = 5, 10, 20, and 50%) were all above 0.99, indicating a strong linear fit. As the conversion (α) progressed, the calculated Ea increased from 60 to 232 kJ mol−1, yielding a mean value of 131 kJ mol−1. Detailed weight-loss curves and regression results for the Flynn–Wall–Ozawa (FWO) method are provided in Supporting Information Figure S1 and Table S1.
By substituting the calculated activation energy into the Arrhenius equation, we determined the necessary durations for the accelerated aging tests conducted at 150 °C, selected based on TGA evidence of no meaningful mass loss below ~300 °C and the validated upper limit of our DMA/TMA workflow (≤150 °C). The resulting accelerated aging times, which are designed to simulate 1, 3, and 5 years of service life at operating temperatures of 70 °C, 80 °C, and 90 °C, are detailed in Table 1. The specimen identification scheme used in this study, #1 through #10, corresponds to these incremental aging durations, beginning with the un-aged specimen (#1) and culminating with the specimen aged for a maximum of 91.8 h (#10). This nomenclature is maintained consistently in all following sections.

3. Accelerated Ageing Test and Evaluation of Material Property Changes in NBR with Degradation

3.1. Experimental Equipment and Test Conditions

All experimental equipment and key test conditions used in this study are summarized in Table 2. To evaluate the mechanical changes in the acceleratedly aged NBR specimens, tensile tests were first performed on five specimens for each condition in accordance with the ASTM D638 standard [35]. From the five resulting stress-strain curves, the specimen exhibiting the median behavior was selected as the representative specimen. Subsequently, thermogravimetric analysis (TGA), dynamic mechanical analysis (DMA), and thermomechanical analysis (TMA) were all performed sequentially on this same specimen. This in-depth analysis of a single representative specimen was intended to ensure data consistency while simultaneously enhancing both experimental efficiency and reliability.
To ensure the appropriateness of selecting the median specimen for subsequent TGA, DMA, and TMA tests, a statistical reproducibility analysis was conducted on the tensile test results. The coefficients of variation for ultimate tensile strength, elongation at break, and elastic modulus across all groups were below 2.6%, indicating low variability among specimens. Moreover, the deviation of the median value from the group mean was within 0.77%, confirming that the median specimen is statistically representative of the mechanical behavior of the entire group.

3.2. Results and Discussion

The mechanical properties of NBR were evaluated as a function of aging time. The stress-strain curves for Samples #1, #4, #7, and #10 are presented in Figure 2, while the ultimate tensile strength, elastic modulus, and elongation at break are summarized in Table 3. As the aging time increased, the elongation at break decreased sharply from 273% (a strain of 2.73) to 5.5% (a strain of 0.055), whereas the elastic modulus increased linearly from 6.01 MPa to 76.71 MPa. This behavior is a clear demonstration of typical embrittlement, where the material loses flexibility and becomes stiffer. Such changes are characteristic of the thermal-oxidative aging process in NBR, during which oxidative cross-linking reactions restrict the mobility of the polymer network [36,37]. On the other hand, the tensile strength showed no significant change in the initial stages of aging but exhibited a decreasing trend in the later stages (see Supporting Information Figure S2).
To elucidate the underlying cause of the observed changes in mechanical properties, DMA was carried out. Storage modulus and crosslink density values extracted from the rubbery plateau region (Tg + 50 °C) are summarized in Table 4. Detailed DMA measurement results are shown in Figure 3. With progressive aging, an increasing trend in the storage modulus and a decreasing trend in the tan δ peak were observed. This behavior indicates an increase in the crosslink density between the polymer chains. Indeed, the crosslink density, calculated based on Flory’s theory, was found to increase linearly with respect to aging time.
Figure 4 visually represents the significant relationship between the rise in crosslink density and the decline of mechanical properties. As depicted in Figure 4a, the crosslink density derived from DMA analysis shows a clear linear increase as a function of accelerated aging time. Figure 4b illustrates the correlation between crosslink density and tensile strength. A Pearson correlation analysis confirmed a statistically significant, strong negative correlation between tensile strength and crosslink density (r = −0.986, p < 0.001). Similarly, a very strong and statistically significant negative correlation was found between elongation at break and crosslink density (r = −0.965, p < 0.001). These statistical findings confirm that the deterioration in the material’s strength and flexibility is significantly associated with the increase in crosslink density caused by thermal degradation.
TGA and TMA were conducted to investigate further changes in thermal characteristics due to degradation. According to the TMA results, neither the glass transition temperature (Tg) nor the coefficient of thermal expansion (CTE) displayed a distinct trend as a function of aging time. Thermal expansion coefficients and Tg values for all aging conditions are listed in Supporting Information Table S2. Similarly, TGA revealed no significant shift in the primary decomposition temperature of the aged specimens; however, a tendency for reduced weight loss was observed during the initial dehydration phase (refer to Supporting Information Table S3). This implies that some chemical transformations had already taken place during the accelerated aging protocol.

4. Selection and Validation of Optimal Hyperelastic Models for Design Reliability

4.1. Hyperelastic Constitutive Models

Materials characterized by non-linear stress-strain behavior are known as hyperelastic materials. These materials, of which rubber is a typical example, can sustain large elastic deformations and fully recover their initial configuration upon the removal of the load. The mechanical response of such materials is described using hyperelastic models, where the stress-strain relationship is derived from a scalar function known as the strain energy density function. The strain energy density function, W, quantifies the energy stored per unit of undeformed volume, and the specific mathematical expression for W constitutes the hyperelastic model itself [38].
The strain energy density function can be generally approximated as follows in Equation (3) [39].
W = W ( I 1 , I 2 , I 3 )
where I 1 , I 2 , I 3 are the three invariants of the Green stress tensor;
I 1 = λ 1 2 + λ 2 2 + λ 3 2
I 2 = λ 1 2 λ 2 2 + λ 2 2 λ 3 2 + λ 3 2 λ 1 2
I 3 = λ 1 2 λ 2 2 λ 3 2
Here, λ 1 2 λ 2 2 λ 3 2 is represented by the main elongation.
To characterize the hyperelastic constitutive response of the rubber, this research employed three models: the Neo-Hookean model, the 5-parameter Mooney–Rivlin model, and the second-order Yeoh model. A brief overview of these three models is provided below.

4.1.1. Neo-Hookean Model

The Neo-Hookean model, which is founded on molecular theory, represents the simplest form of hyperelasticity. It is considered an extension of Hooke’s law to large deformations [40] and characterizes the material’s hyperelastic response with just one independent material constant. The mathematical expression for the Neo-Hookean model is given as Equation (7) [41].
W = C 1 ( I 1 3 )
where C1 is a material constant, and I1 is the first invariant of the left Cauchy-Green deformation tensor.

4.1.2. Yeoh Secondary Model

The Yeoh model is a phenomenological model similar to the Mooney–Rivlin model, but with the defining characteristic that its strain energy density function depends exclusively on the first strain invariant, I 1 [42]. The mathematical form of the second-order Yeoh model is presented as Equation (8) [43,44].
W = C 10 I 1 3 + C 20 I 3 2 + 1 d 10 J 1 2 + 1 d 20 J 1 4
where I 1 is the first principal strain invariant (sum of the principal strains), J is the strain Jacobian (a measure of the volume change produced by deformation), C10 and C20 are material constants (units of stress), and d 10 and d 20 are compressibility parameters (units of inverse stress).

4.1.3. Mooney–Rivlin 5-Parameter Model

The Mooney–Rivlin model is considered one of the foundational hyperelastic models, alongside the Yeoh model. Owing to its relative simplicity and effectiveness, it has become one of the most popular models for analyzing nonlinear elastic behavior. The model’s versatility allows it to be extended to capture more complex deformation modes by increasing the number of parameters [45]. The specific form of the 5-parameter Mooney–Rivlin model employed in this research is given by Equation (9). The material constants C10, C01, C11, C20, and C02 are determined from the experimental data.
W = 2 C 10 ( λ 1 λ ) + 2 C 01 λ 1 λ 3 + 6 C 11 λ 2 λ 1 + 1 λ 2 + 1 λ 3 1 λ 4 + 4 C 20 λ 1 1 λ 3 λ 2 + 2 λ 3 + 4 C 02 ( 2 λ + 1 λ 2 3 ) ( 1 1 λ 3 )
The fitted hyperelastic constants for the Neo-Hookean, Yeoh, and Mooney–Rivlin models were obtained from tensile test data. These values are summarized in Table 5. They were then used as input parameters for subsequent FEA simulations to evaluate model accuracy.

4.2. FEA Model and Accuracy Validation Method

To assess the accuracy of the selected hyperelastic models, we constructed a finite element (FE) simulation model for mechanical property tests. Figure 5 illustrates the FE model of the test specimen, including the generated mesh and the applied boundary conditions. The model’s geometry was designed in accordance with the ASTM D 638 Type IV standard, making it identical to the physical specimens used in the experimental tensile tests. To ensure precise calculations of deformation and stress, a more refined mesh was utilized in the central gauge section. The experimental conditions were simulated by constraining one end of the specimen with a fixed support and applying a prescribed displacement to the opposite end. The numerical analysis was conducted using ANSYS Academic Research software (ANSYS Workbench 2024 R2), made available through the Convergence Parts and Materials Core Research Support Center.
A mesh sensitivity evaluation was undertaken before validating the finite element models. Maximum von Mises stress in the gauge section was calculated for element counts between 1584 and 32,224. The 32,224-element mesh was used as the reference. The relative error fell to 0.5 percent at around 6720 elements, which was taken as the point of convergence. A mesh of 6720 elements was used for all further simulations to provide a practical balance between accuracy and computational efficiency. Figure 6 shows the convergence behavior with relative error in maximum von Mises stress versus element count and the 0.5 percent threshold for optimal mesh density.
To quantitatively evaluate which hyperelastic model best fits the experimental data, the L2-Norm was employed as a measure of error. In linear algebra, a norm is defined as a function that assigns a strictly positive length or magnitude to each vector in a vector space. The L2-Norm, often associated with the principle of Least Squares Error, works by minimizing the sum of the squared differences between the target (experimental) and estimated (model-predicted) values [46].
The mathematical expression for the L2-Norm is given in Equation (10).
S = n = 1 i y 1 f x 1 2
where y 1 represents the target (experimental) value and x 1 denotes the estimated (model-predicted) value. For each sample in this study, ten discrete data points were selected from its experimental stress-strain curve. The L2-Norm was then used to evaluate which of the Mooney–Rivlin, Yeoh, and Neo-Hookean models more closely approximated the actual experimental results.

4.3. Model Accuracy Assessment and Optimal Model Selection

As representative examples, Figure 7 illustrates a comparison between the experimental stress-strain curves for two specimens, Sample #2 (less aged) and Sample #9 (more aged), and the corresponding curves predicted by the hyperelastic models through finite element analysis. The degree of approximation for each model was quantified by calculating the L2-Norm at the points indicated by markers. A comprehensive summary of the L2-Norm values for all specimens across the different aging durations is provided in Table 6. A lower L2-Norm value signifies a better agreement between the model prediction and the experimental results.
The goodness-of-fit analysis for the hyperelastic models across various degradation conditions demonstrated that during the initial phase of degradation (up to approximately 7 h of accelerated aging), the Neo-Hookean model yielded the best agreement with experimental data. This suggests that a simpler constitutive model is sufficient for effectively predicting the mechanical response in the early stages. Conversely, for advanced stages of degradation (16.2 h and beyond), the 5-parameter Mooney–Rivlin model exhibited superior accuracy, indicating the necessity of a more sophisticated model incorporating higher-order terms to capture the material’s behavior.
These findings underscore the importance of a strategic approach to selecting a suitable hyperelastic model according to the material’s degradation state. Specifically, an efficient methodology would involve applying the Neo-Hookean model for materials in the early stages of degradation and transitioning to the Mooney–Rivlin model as the material becomes more severely degraded.

5. Proposal for a Comprehensive Hyperelastic Model Selection Guideline

5.1. Defining the Degradation Inflection Point and Embrittlement Boundary

Based on the experimental results from Section 4, this section synthesizes the analysis of mechanical and thermal property changes, derived from Arrhenius-based NBR accelerated aging tests, with the accuracy evaluation results of hyperelastic models from Finite Element Analysis (FEA). Based on this, an integrated hyperelastic model selection guideline is proposed to improve the reliability of NBR components in real-world engineering design.
A key finding of this research is the observation of a critical inflection point at 16.2 h of accelerated aging (at 150 °C), which corresponds to 3 years of service life at 80 °C. At this juncture, the material’s embrittlement significantly intensifies. This was evidenced by a sharp drop in the elongation at break to 124.6% (a strain of 1.246) from 181.3% (a strain of 1.813) recorded at the prior stage (7.27 h). The underlying cause for this shift in behavior is a critical increase in crosslink density. Upon reaching 16.2 h, the crosslink density surged to over 3800 mol m−3 from 3007 mol m−3 at the previous stage, thereby restricting polymer chain mobility and stripping the material of its flexibility. This qualitative transformation in the material directly influenced the selection of the FEA model; it was confirmed that at this turning point, the optimal constitutive model transitions from the simpler Neo-Hookean to the more complex 5-parameter Mooney–Rivlin model.

5.2. Guideline Based on Service Conditions: The Degradation Map

To establish a clear guideline for selecting hyperelastic models, we conducted a linear regression analysis based on our findings, targeting three key parameters: strain at maximum tensile strength, ultimate tensile strength, and crosslink density. The objective was to quantify a predictive model for these material properties as a function of service conditions (temperature and time). The outcomes of the linear regression analysis for these three variables are summarized in Table 7.
Upon comparing the three properties, strain at maximum tensile strength exhibited the strongest regression fit and the highest predictive capability. Crosslink density, in turn, proved to be a valuable auxiliary indicator for interpreting behavioral changes arising from structural transformations in the material. While ultimate tensile strength is a conventional metric for performance assessment, its predictive power in this context was found to be relatively weak. Consequently, this study established model selection criteria primarily based on strain at maximum tensile strength and crosslink density.
In this research, we define the embrittlement boundary for NBR by the conditions where the strain at maximum tensile strength is ≤1.5 (mm/mm) or the crosslink density is ≥3300 mol m−3. This threshold marks the onset of brittle behavior in the material. Critically, it also corresponds to the domain where the more complex Mooney–Rivlin model begins to provide a superior fit to the experimental data.
This concept is visually synthesized in Figure 8, which presents a contour map of strain at maximum tensile strength against service conditions (years and temperature). The boundary line for Strain = 1.5 (mm/mm), indicated by a dashed line, delineates the zone of rapid ductility loss. The superposition of this boundary with the high crosslink density region identifies a critical zone of concurrent structural and mechanical degradation. This suggests that within this zone, material embrittlement must be assumed, and a transition in the constitutive model is warranted.
However, note that the embrittlement boundary delineated in this map (i.e., Crosslink Density ≥ 3300 mol m−3) was determined for the specific NBR compound tested in this study. This value can vary depending on the rubber’s formulation, such as the type and amount of additives. Therefore, applying this map directly to other NBR systems would require preliminary validation.
Therefore, the hyperelastic model selection guideline framework proposed herein can serve as a foundational basis for developing future intelligent model selection systems that account for material degradation. Such advancements are expected to enhance the accuracy of nonlinear analyses for rubber components, thereby contributing to more robust and reliable design optimization.

5.3. Discussion

This study successfully establishes an integrated workflow to quantitatively analyze the state of degradation in NBR and, based on this analysis, select the optimal hyperelastic model. The central finding is that as degradation progresses, the most accurate predictive model transitions from the simpler Neo-Hookean model to the more complex Mooney–Rivlin model. This demonstrates that a data-driven methodology can fill the technical gap identified in the introduction, where model selection has traditionally relied on empirical engineering judgment.
This model transition can be rationalized by the physical evolution of the material. In its early degradation stages, NBR possesses a relatively uniform polymer network, the behavior of which can be sufficiently captured by the statistical mechanics-based Neo-Hookean model [47]. However, with continued aging, additional cross-linking reactions lead to a linear increase in crosslink density. This increased density constrains polymer chain mobility and enhances the network’s non-ideality, making its behavior too complex to be described by a single-parameter ( I 1 ) model. As a result, the Mooney–Rivlin model [48], which incorporates the second strain invariant ( I 2 ) into its strain energy function, demonstrates superior predictive accuracy in these later stages.
The ‘degradation map’ (Figure 8), which is the culmination of this analysis, provides designers with a tangible and practical guideline. For instance, an engineer can input a component’s target service life and temperature into the map to instantly assess whether it falls within the embrittlement zone. This allows for the informed adoption of a more sophisticated model, such as the Mooney–Rivlin model, for Finite Element Analysis (FEA), thereby significantly enhancing design reliability.
Although this study puts forth a successful framework for model selection for NBR, it is crucial to recognize its limitations to generalize the findings and broaden their applicability. The main limitations are that accelerated aging was conducted at a single temperature (150 °C), which may not fully represent the influence of diverse thermal conditions, and that the study did not account for chemo-mechanical interactions with substances like oil, which are prevalent in actual service environments. To address these limitations, future research should focus on validating the Arrhenius model’s predictive power across multiple temperature points. Moreover, the framework needs to be expanded by conducting combined-environment tests (e.g., simultaneous thermal and oil exposure) to build more robust degradation models and design guidelines that more faithfully represent real-world powertrain conditions.

6. Conclusions

This study sought to analyze the degradation behavior of industrial Nitrile-butadiene rubber (NBR) and to propose an integrated workflow for high-fidelity Finite Element Analysis. To this end, a TGA-derived activation energy (Ea = 131.34 kJ mol−1) was applied to the Arrhenius model to systematically design accelerated aging tests. Subsequent characterization via DMA and tensile tests quantitatively confirmed that as degradation proceeds, crosslink density increases linearly (from 2.0 × 103 to 6.8 × 103 mol m−3). This was identified as the primary cause of embrittlement, evidenced by a drastic reduction in elongation at break from 273% (a strain of 2.73) to 4% (a strain of 0.04). Moreover, an L2-Norm error analysis demonstrated that the Neo-Hookean model provides the most accurate predictions for early-stage degradation, whereas the Mooney–Rivlin model excels for advanced degradation states. By integrating these findings, this research culminates in a practical design guideline, presented as a degradation map, that enables the selection of an optimal hyperelastic model based on real-world service conditions (temperature and time).
However, despite the validity of the proposed framework, this study has several limitations that suggest directions for future research. First, the accelerated aging was conducted at a single temperature (150 °C), meaning the Arrhenius model’s predictive accuracy was not fully validated across a wider range of thermal conditions. Second, this study focused solely on thermo-oxidative aging, whereas real-world service environments for NBR often involve synergistic effects from contact with substances like oil or from sustained mechanical stress. Furthermore, the brittle boundary value presented in this study (crosslink density ≥ 3300 mol m−3) is based on the experimental results of the specimens used, and it may vary depending on the actual rubber formulation, so further research on more diverse NBR specimens is needed. Finally, as this study’s characterization was centered on mechanical properties, further investigation into the chemical and morphological changes on the material’s surface, for instance, through techniques like FTIR or SEM, is warranted for future research.
The predictive accuracy of FEA for rubber components is critically dependent on the suitability of the chosen material model and its input parameters. In light of this, the integrated approach proposed herein facilitates a paradigm shift from conventional, experience-based design to a data-driven, predictive methodology. This work is therefore expected to make a substantial contribution to enhancing the accuracy of long-term reliability assessments and the overall design of components based on rubber materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym17172316/s1, Figure S1: TGA Weight Loss and Flynn–Wall–Ozawa (FWO) Linear Regression for Activation Energy Determination; Figure S2: Evolution of Mechanical Properties of NBR with Accelerated Aging Time; Table S1: Activation Energy of NBR at Different Conversion Levels Based on FWO Method; Table S2: Thermal Expansion Coefficients and Glass Transition Temperatures of Aged NBR Specimens (TMA Results); Table S3: Thermogravimetric Analysis of Aged NBR Specimens: DTG Peak, Ignition, and Burnout Characteristics.

Author Contributions

N.-Y.C., D.-S.K., and S.-U.Z.; methodology, N.-Y.C.; software, N.-Y.C. (ANSYS Workbench 2024 R2); validation, N.-Y.C., D.-S.K., and S.-U.Z.; formal analysis, N.-Y.C.; investigation, N.-Y.C.; resources, N.-Y.C.; data curation, N.-Y.C.; writing—original draft preparation, N.-Y.C.; writing—review and editing, N.-Y.C., D.-S.K., and S.-U.Z.; visualization, N.-Y.C.; supervision, D.-S.K. and S.-U.Z.; project administration, D.-S.K. and S.-U.Z.; funding acquisition, D.-S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Industrial Strategic Technology Development Program-Alchemist Project (RS-2024-00432559, Development of Space-Grade Monolithic AlGaN/Diamond Ultimate Semiconductor) funded by the Ministry of Trade, Industry and Energy (MOTIE, Republic of Korea).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
NBRNitrile-butadiene rubber
FEAFinite Element Analysis
TGAThermogravimetric analysis
DMADynamic mechanical analysis
TMAThermomechanical analysis
FWOFlynn–Wall–Ozawa
CTECoefficient of thermal expansion
TgGlass transition temperature
WStrain energy density function
EaActivation energy
I1, I2, I3Invariants of the Green stress tensor
kChemical reaction rate
RUniversal gas constant
R2Coefficient of determination
TAbsolute temperature
αConversion rate
βConstant heating rate
ϵbreakStrain at break/Elongation at break
λMain elongation
ρxCrosslink density
σmaxUltimate tensile strength

References

  1. Li, Z.; Cai, D.; Xu, H.; Sun, Y.; Zhao, X.; Hu, W.; Zhang, X.; Zhou, Q.; Zhang, X.; Xu, X. Healable and recyclable elastomers with record-high mechanical robustness, unprecedented crack tolerance, and superhigh elastic restorability. Adv. Mater. 2021, 33, 2101498. [Google Scholar] [CrossRef]
  2. Jiang, B.; Jia, X.; Wang, Z.; Wang, T.; Guo, F.; Wang, Y. Influence of thermal aging in oil on the friction and wear properties of nitrile butadiene rubber. Tribol. Lett. 2019, 67, 86. [Google Scholar] [CrossRef]
  3. Ahmed, F.S.; Shafy, M.; Abd El-Megeed, A.A.; Hegazi, E.M. The effect of γ-irradiation on acrylonitrile–butadiene rubber NBR seal materials with different antioxidants. Mater. Des. 2012, 36, 823–828. [Google Scholar] [CrossRef]
  4. Bhaumik, S.; Kumaraswamy, A.; Guruprasad, S.; Bhandari, P. Investigation of friction in rectangular nitrile-butadiene rubber hydraulic rod seals for defence applications. J. Mech. Sci. Technol. 2015, 29, 4793–4799. [Google Scholar] [CrossRef]
  5. Lou, W.; Zhang, W.; Wang, H.; Jin, T.; Liu, X. Influence of hydraulic oil on degradation behavior of nitrile rubber O-rings at elevated temperature. Eng. Fail. Anal. 2018, 92, 1–11. [Google Scholar] [CrossRef]
  6. Ito, M.; Nagai, K. Degradation issues of polymer materials used in railway field. Polym. Degrad. Stab. 2008, 93, 1723–1735. [Google Scholar] [CrossRef]
  7. Mo, Y.M.; Gong, Y.; Yang, Z.G. Failure analysis on the O-ring of radial thrust bearing room of main pump in a nuclear power plant. Eng. Fail. Anal. 2020, 115, 104673. [Google Scholar] [CrossRef]
  8. Liu, J.; Li, X.; Xu, L.; Zhang, P. Investigation of aging behavior and mechanism of nitrile-butadiene rubber in the accelerated thermal aging environment. Polym. Test. 2016, 54, 59–66. [Google Scholar] [CrossRef]
  9. Morrell, P.R.; Patel, M.; Skinner, A.R. Accelerated thermal ageing studies on nitrile rubber O-rings. Polym. Test. 2003, 22, 651–656. [Google Scholar] [CrossRef]
  10. Zhu, M.; Huang, H.; Li, H.; Huang, G.; Lan, J.; Fu, J.; Fan, J.; Liu, Y.; Ke, Z.; Guo, X.; et al. A study on the aging mechanism and anti-aging properties of nitrile butadiene rubber: Experimental characterization and molecular simulation. Polymers 2025, 17, 1446. [Google Scholar] [CrossRef]
  11. Lou, W.; Zhang, W.; Wang, H.; Jin, T.; Liu, X. Stress–thermal oxidative aging behavior of hydrogenated nitrile rubber seals. J. Appl. Polym. Sci. 2019, 136, 47014. [Google Scholar] [CrossRef]
  12. Lou, W.; Zhang, W.; Jin, T.; Liu, X.; Dai, W. Synergistic effects of multiple environmental factors on degradation of hydrogenated nitrile rubber seals. Polymers 2018, 10, 897. [Google Scholar] [CrossRef]
  13. Shao, H.; Guo, Q.; He, A. Evolution of crosslinking network structure and thermo-oxidative aging behavior of unfilled NR/BR blends with TBIR as extra functional compatibilizer. Polym. Test. 2022, 115, 107715. [Google Scholar] [CrossRef]
  14. da Silva Ruiz, N.M.; de Sousa, A.M.; Calderari, M.R.; de Figueiredo, M.A.; Lima, A.P.; de Azevedo, E.R. Time domain NMR evaluation of thermal and thermochemical aging of nitrile rubber on crosslinking and mechanical properties. Polym. Degrad. Stab. 2024, 224, 110727. [Google Scholar] [CrossRef]
  15. Wang, S.; Wang, C.; He, A. Accelerated aging behavior and degradation mechanism of nitrile rubber in thermal air and hydraulic oil environments. Polym. Eng. Sci. 2023, 63, 2218–2232. [Google Scholar] [CrossRef]
  16. Zanzi, M.S.; de Souza, E.L.; Dutra, G.B.; Paiva, K.V.; Oliveira, J.L.; Cunha, T.V.; Monteiro, A.S. Service lifetime prediction of nitrile butadiene rubber gaskets used in plate heat exchangers. J. Appl. Polym. Sci. 2022, 139, e52523. [Google Scholar] [CrossRef]
  17. Ammineni, S.P.; Lingaraju, D.; Nagaraju, C. Estimating shelf life and degradation mechanisms of nitrile butadiene rubber for viscoelastic dampers. J. Elastomers Plast. 2024, 56, 434–453. [Google Scholar] [CrossRef]
  18. Tayefi, M.; Eesaee, M.; Hassanipour, M.; Elkoun, S.; David, E.; Nguyen-Tri, P. Thermal aging behavior and lifetime prediction of industrial elastomeric compounds based on styrene–butadiene rubber. Polym. Eng. Sci. 2025, 65, 3226–3246. [Google Scholar] [CrossRef]
  19. Liu, Q.; Shi, W.; Chen, Z.; Li, K.; Liu, H.; Li, S. Rubber accelerated ageing life prediction by Peck model considering initial hardness influence. Polym. Test. 2019, 80, 106132. [Google Scholar] [CrossRef]
  20. Kömmling, A.; Jaunich, M.; Goral, M.; Wolff, D. Insights for lifetime predictions of O-ring seals from five-year long-term aging tests. Polym. Degrad. Stab. 2020, 179, 109278. [Google Scholar] [CrossRef]
  21. Shi, J.; Yuan, S.; Zhang, W.; Zhang, J.; Chen, H. Hydrothermal aging mechanisms and service life prediction of twisted bamboo fiber wound composites. Mater. Des. 2023, 227, 111716. [Google Scholar] [CrossRef]
  22. Xie, Z.; Huang, X.; Zhang, K.; Yan, S.; Chen, J.; He, R.; Li, J.; Zhong, W. Thermo-oxidative aging effects on hyperelastic behavior of EPDM rubber: A constitutive modeling approach. Materials 2025, 18, 2236. [Google Scholar] [CrossRef]
  23. Jiang, D.; Wang, Z.; Wang, X. A general hyperelastic model for rubber-like materials incorporating strain-rate and temperature. J. Elastomers Plast. 2024, 56, 624–651. [Google Scholar] [CrossRef]
  24. Zhu, M.; Ma, D.; Zhou, Y.; Huang, H.; Shao, Z.; Wu, F.; Li, B. The change of sealing property in the aging process of NBR sealing equipment based on finite element analysis. Coatings 2024, 14, 1178. [Google Scholar] [CrossRef]
  25. Shao, Z.; Zhu, M.; Wang, H.; Li, M. Effect of aging of foam rubber on properties and constitutive models. J. Phys. Conf. Ser. 2021, 2011, 107715. [Google Scholar] [CrossRef]
  26. Marl, S.; Ni, X.; Hornig, T.; Spieker, C.; Giesen, R.-U.; Heim, H.-P.; Fister, M. Correlations between the aging behavior and finite element method simulation of three silicone elastomers. Materials 2024, 17, 3961. [Google Scholar] [CrossRef] [PubMed]
  27. Korba, A.G.; Kumar, A.; Barkey, M.E. A model for hyper-elastic material behavior under thermal aging with an application to natural rubber. In Proceedings of the ASME 2018 Manufacturing Science and Engineering Conference (MSEC2018), College Station, TX, USA, 18–22 June 2018. [Google Scholar]
  28. Park, J.; Choun, Y.S.; Kim, M.K.; Hahm, D. Revaluation of the aging property modification factor of lead rubber bearings based on accelerated aging tests and finite element analysis. Nucl. Eng. Des. 2019, 347, 59–66. [Google Scholar] [CrossRef]
  29. Woo, C.S.; Park, H.S. Useful lifetime prediction of rubber component. Eng. Fail. Anal. 2011, 18, 1645–1651. [Google Scholar] [CrossRef]
  30. Brown, R.P. Practical Guide to the Assessment of the Useful Life of Rubbers; iSmithers Rapra: Shropshire, UK, 2001. [Google Scholar]
  31. Ozawa, T. Thermal analysis. Bull. Chem. Soc. Jpn. 1965, 38, 1881–1886. [Google Scholar] [CrossRef]
  32. Chen, S.; Yu, H.; Ren, W.; Zhang, Y. Thermal degradation behavior of hydrogenated nitrile-butadiene rubber/clay nanocomposites. Thermochim. Acta 2009, 491, 103–108. [Google Scholar] [CrossRef]
  33. Flynn, J.H.; Wall, L.A. A quick, direct method for the determination of activation energy from thermogravimetric data. Polym. Lett. 1966, 4, 323–328. [Google Scholar] [CrossRef]
  34. Doyle, C.D. Kinetic analysis of thermogravimetric data. J. Appl. Polym. Sci. 1961, 5, 285–292. [Google Scholar] [CrossRef]
  35. ASTM D638; Standard Test Method for Tensile Properties of Plastics. ASTM International: West Conshohocken, PA, USA, 2014.
  36. Zhao, J.; Yang, R.; Iervolino, R.; Barbera, S. Investigation of crosslinking in the thermooxidative aging of nitrile–butadiene rubber. J. Appl. Polym. Sci. 2015, 132, 3. [Google Scholar] [CrossRef]
  37. Zhao, J.; Yang, R.; Iervolino, R.; Barbera, S. Changes of chemical structure and mechanical property levels during thermo-oxidative aging of NBR. Rubber Chem. Technol. 2013, 86, 591–603. [Google Scholar] [CrossRef]
  38. Ogden, R.W. Non-Linear Elastic Deformations; Courier Corporation: Amherst, MA, USA, 1997. [Google Scholar]
  39. Tejasri, P.; Basak, A. Comparative study of computational material models. Int. J. Appl. Eng. Res. 2018, 13, 8588–8595. [Google Scholar]
  40. Boulanger, P.; Hayes, M. Finite-amplitude waves in Mooney-Rivlin and Hadamard materials. In Topics in Finite Elasticity; Springer: Vienna, Austria, 2001; pp. 131–167. [Google Scholar]
  41. Kim, B.; Lee, S.B.; Lee, J.; Cho, S.; Park, H.; Yeom, S.; Park, S.H. A comparison among Neo-Hookean, Mooney-Rivlin, and Ogden models for chloroprene rubber. Int. J. Precis. Eng. Manuf. 2012, 13, 759–764. [Google Scholar] [CrossRef]
  42. Dhrangdhariya, P.; Maiti, S.; Rai, B. Effect of spoke design and material nonlinearity on non-pneumatic tire stiffness and durability performance. arXiv 2021, arXiv:2103.03637. [Google Scholar] [CrossRef]
  43. Amatoury, J.; Cheng, S.; Kairaitis, K.; Wheatley, J.R.; Amis, T.C.; Bilston, L.E. Development and validation of a computational finite element model of the rabbit upper airway: Simulations of mandibular advancement and tracheal displacement. J. Appl. Physiol. 2016, 120, 743–757. [Google Scholar] [CrossRef]
  44. Hejazi, F.; Farahpour, H.; Ayyash, N.; Chong, T. Development of a volumetric compression restrainer for structures subjected to vibration. J. Build. Eng. 2022, 46, 103735. [Google Scholar] [CrossRef]
  45. Okeke, C.P.; Thite, A.N.; Durodola, J.F.; Greenrod, M.T. Hyperelastic polymer material models for robust fatigue performance of automotive LED lamps. Procedia Struct. Integr. 2017, 5, 600–607. [Google Scholar] [CrossRef]
  46. York, D. Least squares fitting of a straight line with correlated errors. Earth Planet. Sci. Lett. 1968, 5, 320–324. [Google Scholar] [CrossRef]
  47. Treloar, L.R.G. The elasticity of a network of long-chain molecules—II. Trans. Faraday Soc. 1943, 39, 241–246. [Google Scholar] [CrossRef]
  48. Rivlin, R.S. Large elastic deformations of isotropic materials IV. Philos. Trans. R. Soc. Lond. A 1948, 241, 379–397. [Google Scholar]
Figure 1. Workflow from accelerated-aging design to hyper-elastic model-selection guideline for NBR seals.
Figure 1. Workflow from accelerated-aging design to hyper-elastic model-selection guideline for NBR seals.
Polymers 17 02316 g001
Figure 2. Representative stress-strain curves of NBR specimens at selected aging stages.
Figure 2. Representative stress-strain curves of NBR specimens at selected aging stages.
Polymers 17 02316 g002
Figure 3. Temperature-Dependent Dynamic Mechanical Properties of NBR during Thermal Aging.
Figure 3. Temperature-Dependent Dynamic Mechanical Properties of NBR during Thermal Aging.
Polymers 17 02316 g003
Figure 4. (a) Linear increase in crosslink density as a function of accelerated aging time. (b) Corresponding decrease in tensile strength as a function of crosslink density.
Figure 4. (a) Linear increase in crosslink density as a function of accelerated aging time. (b) Corresponding decrease in tensile strength as a function of crosslink density.
Polymers 17 02316 g004
Figure 5. Finite element analysis (FEA) model for the validation of hyperelastic models.
Figure 5. Finite element analysis (FEA) model for the validation of hyperelastic models.
Polymers 17 02316 g005
Figure 6. Mesh convergence analysis showing relative error in maximum von Mises stress versus element count. The red dashed line indicates the relative error trend, and the gray dashed line represents the 0.5% error threshold.
Figure 6. Mesh convergence analysis showing relative error in maximum von Mises stress versus element count. The red dashed line indicates the relative error trend, and the gray dashed line represents the 0.5% error threshold.
Polymers 17 02316 g006
Figure 7. Validation of hyperelastic models against experimental stress-strain curves at different degradation stages: (a) early-stage ageing (1.4 h) and (b) late-stage ageing (55.12 h).
Figure 7. Validation of hyperelastic models against experimental stress-strain curves at different degradation stages: (a) early-stage ageing (1.4 h) and (b) late-stage ageing (55.12 h).
Polymers 17 02316 g007
Figure 8. Design-oriented degradation map and model selection guideline for NBR based on service conditions (temperature and time). The arrows indicate the transition boundary and direction of the applicable hyperelastic model (from Neo-Hookean to Mooney–Rivlin).
Figure 8. Design-oriented degradation map and model selection guideline for NBR based on service conditions (temperature and time). The arrows indicate the transition boundary and direction of the applicable hyperelastic model (from Neo-Hookean to Mooney–Rivlin).
Polymers 17 02316 g008
Table 1. Accelerated-aging durations corresponding to the target service times and temperatures.
Table 1. Accelerated-aging durations corresponding to the target service times and temperatures.
Test #12345678910
Target service time [year]Unaged131531535
Target service temperature [°C]Unaged707080708090809090
Accelerated-ageing time [h]Unaged1.44.455.357.2716.218.3526.7755.1291.77
Table 2. Experimental equipment and test conditions.
Table 2. Experimental equipment and test conditions.
CategoryEquipment/ModelManufacturer (Country)Main Test ConditionsStandard/Mode and Application
Accelerated AgeingTemperature Chamber SH-662ESPEC (Osaka, Japan)150 °C, constant high-temperature exposureAgeing steps from 1.4 to 91.8 h based on Arrhenius equivalent lifetime design
Tensile TestUniversal Testing Machine KDMT-156KDMT (Seoul, Republic of Korea)ASTM D638 type IV specimens
cross-head 300 mm min−1, measured until fracture
≥5 specimens per condition
-
TGATGA N-1000Sinco (Seoul, Republic of Korea)Sample 10–11 mg,
heating rate 20 °C min−1
N2 atmosphere
Data acquisition for FWO activation energy calculation
DMAQ800 DMATA Instruments (New Castle, DE, USA)Tension mode, −40–150 °C
heating rate 5 °C min−1
sinusoidal loading
Calculation of Storage/Loss modulus, tan δ, and cross-link density
TMAQ400 TMATA Instruments (New Castle, DE, USA)Expansion mode, −50–150 °C
heating rate 5 °C min−1
50 mN load
Measurement of Tg and CTE, rubbery plateau analysis
Table 3. Key mechanical properties of NBR specimens at different accelerated aging times.
Table 3. Key mechanical properties of NBR specimens at different accelerated aging times.
Sample #Ageing Time (h)σ Max (MPa)ϵ Break (%)Elastic Modulus (MPa)
1unaged16.4 ± 1.16273.46 ± 20.086.01 ± 0.14
47.215.11 ± 0.90187.60 ± 12.898.06 ± 0.32
726.813.02 ± 0.32135.51 ± 3.909.62 ± 0.30
1091.84.17 ± 0.175.55 ± 0.6676.71 ± 5.98
Table 4. Storage Modulus and Crosslink Density of Aged NBR Specimens from DMA at Tg + 50 °C.
Table 4. Storage Modulus and Crosslink Density of Aged NBR Specimens from DMA at Tg + 50 °C.
Test #Using YearUsing TemperatureAccelerated Aging TimeStorage Modulus at Rubbery Plateau Region (Tg + 50 °C)Crosslink Density
[year][°C][h·min][MPa][mol/m3]
100014.532006.46
21701.416.292269.97
33704.4516.702334.16
41805.3521.092946.25
55707.2721.453007.15
638016.227.263807.24
719018.3524.993483.47
858026.7729.784146.81
939055.1241.395747.51
1059091.7749.166794.17
Table 5. Hyperelastic constants of aged NBR for Neo-Hookean, Yeoh 2nd-order, and Mooney–Rivlin models at different accelerated aging times.
Table 5. Hyperelastic constants of aged NBR for Neo-Hookean, Yeoh 2nd-order, and Mooney–Rivlin models at different accelerated aging times.
Hyperelastic Constant of NBR Neo-Hookean and Yeoh 2nd Order Model at Different Aging Time
Test #Neo-Hookean ModelYeoh Secondary Order
Initial Shear ModulusC10C20
[Mpa][Mpa][Mpa]
10.3190.1350.004
20.3390.1420.006
30.3580.1490.008
40.3650.1580.007
50.3800.1590.008
60.4020.1900.005
70.4340.1990.007
80.4730.243−0.004
90.8720.607−0.735
103.9673.266−145.391
Hyperelastic Constant of NBR Mooney–Rivlin 5 Parameter at Different Aging Time
Test #C10C01C20C11C02
[Mpa][Mpa][Mpa][Mpa][Mpa]
1−1.652.040.04−0.260.96
2−2.282.740.14−0.721.71
3−2.593.080.23−1.092.22
4−3.233.800.38−1.693.05
5−2.763.300.21−1.022.20
6−4.875.601.17−4.696.49
7−4.635.350.96−3.945.76
8−7.508.413.86−13.8415.43
9−44.4146.60467.01−1213.50818.52
10−1140.141151.151,978,220.32−4,174,395.942,205,309.53
Table 6. Comparison of Predictive Accuracy (L2-Norm Error) of Hyperelastic Models for Each Degradation Stage.
Table 6. Comparison of Predictive Accuracy (L2-Norm Error) of Hyperelastic Models for Each Degradation Stage.
Test #Neo-Hookean ModelMooney–Rivlin
5 Parameter
Yeoh Secondary Order
10.01880.01290.0130
20.00190.00960.0053
30.00560.01650.0151
40.00460.00920.0102
50.00620.01580.0062
60.00690.00280.0095
70.00870.00680.0144
80.01380.00370.0126
90.01610.00170.0067
100.00150.00240.0064
Table 7. Regression analysis results for predicting mechanical properties of NBR as a function of service time (Y) and temperature (T).
Table 7. Regression analysis results for predicting mechanical properties of NBR as a function of service time (Y) and temperature (T).
Response VariableRegression EquationR2Adj. R2Pred. R2p-Value (Year)p-Value (Temp)
Strain at Maximum Tensile Strength6.57 − 0.3292Y − 0.0519T80.2%75.8%59.6%0.0010.008
Ultimate Tensile Strength37.85 − 1.266Y − 0.2718T75.9%70.5%46.8%0.0030.006
Crosslink Density−6196 + 517Y + 105.278.4%73.6%54.2%0.0020.005
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

Choi, N.-Y.; Kim, D.-S.; Zhang, S.-U. Design-Oriented Degradation Mapping and Hyperelastic Model-Switch Guidelines for Nitrile-Butadiene Rubber Seals. Polymers 2025, 17, 2316. https://doi.org/10.3390/polym17172316

AMA Style

Choi N-Y, Kim D-S, Zhang S-U. Design-Oriented Degradation Mapping and Hyperelastic Model-Switch Guidelines for Nitrile-Butadiene Rubber Seals. Polymers. 2025; 17(17):2316. https://doi.org/10.3390/polym17172316

Chicago/Turabian Style

Choi, Na-Yeon, Dong-Seok Kim, and Sung-Uk Zhang. 2025. "Design-Oriented Degradation Mapping and Hyperelastic Model-Switch Guidelines for Nitrile-Butadiene Rubber Seals" Polymers 17, no. 17: 2316. https://doi.org/10.3390/polym17172316

APA Style

Choi, N.-Y., Kim, D.-S., & Zhang, S.-U. (2025). Design-Oriented Degradation Mapping and Hyperelastic Model-Switch Guidelines for Nitrile-Butadiene Rubber Seals. Polymers, 17(17), 2316. https://doi.org/10.3390/polym17172316

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