Next Article in Journal
A Spatiotemporal–Semantic Coupling Intelligent Q&A Method for Land Use Approval Based on Knowledge Graphs and Intelligent Agents
Previous Article in Journal
Circular Fertilization Strategy Using Sulphur with Orange Waste Enhances Soil Health and Broccoli Nutritional and Nutraceutical Quality in Mediterranean Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Insights into the Insulating and Pyrolysis Properties of Environmentally Friendly PMVE/CO2 Mixtures: A Collaborative Analysis Based on Density Functional Theory and Reaction Kinetics

School of Electrical Engineering, China University of Mining and Technology, Xuzhou 221116, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2025, 15(16), 9011; https://doi.org/10.3390/app15169011
Submission received: 29 July 2025 / Revised: 13 August 2025 / Accepted: 14 August 2025 / Published: 15 August 2025

Abstract

Perfluoromethyl vinyl ether (PMVE) has recently emerged as a promising environmentally friendly insulating gas with potential for practical applications in the power industry. When mixed with CO2, the PMVE/CO2 mixture exhibits an elevated liquefaction temperature and enhanced insulation performance, making it suitable for engineering use. In this study, density functional theory (DFT) calculations were employed to investigate the reactive sites of PMVE molecules. The results indicate that the C2–O and C3–O bonds are the most susceptible to breakage, highlighting their high reactivity. The optimal insulation performance of the PMVE/CO2 mixture is achieved at a CO2 concentration of approximately 60%, with significant molecular decomposition observed at temperatures exceeding 2600 K. The primary decomposition products include C2F2, COF3, COF2, F, C2F3, CO, CF3, and C2F4. Both high temperature and elevated CO2 content accelerate the decomposition process. These findings provide valuable insights into the insulation properties and thermal stability of the PMVE/CO2 system, offering theoretical support for its potential application in eco-friendly high-voltage insulation technologies.

1. Introduction

Sulfur hexafluoride (SF6) is a colorless, odorless gas with superior dielectric and arc-quenching properties and has been extensively utilized in medium- and high-voltage electrical equipment, including gas-insulated circuit breakers (GCBs) and cubicle-type gas-insulated switchgear (C-GIS), for over six decades [1,2]. However, SF6 is recognized as one of the most potent greenhouse gases, with a global warming potential (GWP) of 23,900 and an atmospheric lifetime exceeding 3200 years [3,4,5]. In 1997, the Kyoto Protocol designated SF6 as one of the six greenhouse gases subject to international regulation [6]. As global climate policies have tightened, especially in the power sector, the reduction in SF6 emissions has become an environmental imperative. Consequently, the development of low-GWP alternatives with comparable insulation performance is a critical pathway toward sustainable and climate-resilient electrical infrastructure.
In recent years, several high-performance alternative insulating media have attracted increasing attention in the power sector [3,7,8,9,10]. These include gases such as C4F7N, CF3I, perfluorocarbons (PFCs) like c-C4F8, C2F6, C3F8, and perfluoroketones (CnF2nO). Among them, C4F7N demonstrates superior dielectric strength and excellent arc-quenching capability compared to SF6 and has already been deployed in 145 kV gas-insulated switchgear (GIS) and 420 kV gas-insulated lines (GILs) [11]. However, its relatively high liquefaction temperature and ongoing concerns regarding chronic toxicity have limited broader adoption [12]. CF3I exhibits better insulation properties than SF6, with a significantly lower GWP and shorter atmospheric lifetime, making it a promising candidate. Nevertheless, its high biological toxicity and classification by the International Agency for Research on Cancer (IARC) as a Group 3 mutagen have raised safety concerns and constrained its practical use [3]. Perfluorocarbon-based gases generally exhibit high GWPs and long atmospheric lifetimes, posing challenges to their compatibility with low-carbon, environmentally sustainable development goals. Although perfluoroketones such as C6F12O and C5F10O show favorable environmental characteristics, their high liquefaction temperatures of 49 °C and 24 °C mean that they exist as liquids under ambient conditions and therefore cannot be used as standalone gas insulating media [13]. These limitations underscore the urgent need to develop novel insulating gases that offer a balanced combination of high dielectric performance, low liquefaction temperature, and minimal environmental impact.
In November 2023, the Korea Electrical Engineering Research Institute (KERI) announced a newly developed environmentally friendly insulating gas, referred to as “K6.” This gas successfully passed the International Electrotechnical Commission (IEC) circuit-breaking performance test and has since undergone pilot deployment in 145 kV high-voltage circuit breakers [14]. Subsequent studies by Xiao Song and collaborators identified the primary component of K6 as C3F6O, also known as PMVE. Notably, pure PMVE exhibits an extremely low global warming potential (GWP), estimated at less than one millionth of that of SF6, along with a significantly shorter atmospheric lifetime and zero ozone depletion potential, underscoring its excellent environmental compatibility. In addition to its environmental advantages, PMVE demonstrates promising electrical insulation characteristics. Theoretical investigations by South Korean researchers have highlighted its potential as a viable SF6 substitute. For instance, Yeunsoo Park et al. measured the total electron scattering cross-section (TCS) of PMVE and confirmed its favorable electron interaction properties [15]. Furthermore, Nidhi Sinha’s group examined the dielectric strength of PMVE through electron collision experiments, revealing that its critical dielectric strength (475 Td) exceeds that of SF6 (355 Td) [16]. An experimental study by Xiao Song and Zhang Xiaoxing further evaluated the AC breakdown behavior, partial discharge characteristics, and dielectric recovery performance of PMVE. Their results showed that PMVE exhibits stable breakdown voltage under non-uniform electric fields and a narrow gap between partial discharge inception voltage (PDIV) and partial discharge extinction voltage (PDEV), which is advantageous for suppressing discharge activity [17].
Although PMVE exhibits excellent insulation performance, its tendency toward spontaneous combustion and relatively high liquefaction temperature limit its effectiveness as a standalone insulating medium. To address these challenges, PMVE is typically blended with noble gases to enhance its safety and modify its physical properties. Among potential additives, CO2 has attracted particular interest due to its favorable insulating and arc-quenching characteristics, as well as its low boiling point, making it a suitable buffer gas in gas mixtures [18]. Nidhi Sinha and colleagues investigated the synergistic effects of combining PMVE with buffer gases and reported a notable enhancement in insulation performance when CO2 was used as the additive. Specifically, when the PMVE/CO2 mixture ratio approached 60%, the dielectric strength of the blend was found to be comparable to that of pure SF6 [16]. Despite these promising findings, current research remains predominantly focused on the properties of pure PMVE, while investigations into the behavior and performance of PMVE/CO2 mixtures are still limited.
Partial discharge, arc discharge, and spark discharge are common fault phenomena in power equipment, typically originating from insulation defects [19]. Among these, the local temperature during partial discharge can range from 700 to 1200 K, while arc discharges may generate temperatures as high as 3000 to 12,000 K [20]. These extreme thermal conditions can induce the decomposition of insulating gases, leading to the formation of reactive free radicals and various degradation by-products. The accumulation of such decomposition products may deteriorate the dielectric properties of the insulating medium and pose serious risks to the operational safety of electrical systems [21]. Given these concerns, it is of both theoretical and practical importance to investigate the high-temperature decomposition behavior and underlying mechanisms of PMVE/CO2 gas mixtures. Such studies are essential for assessing the long-term reliability, thermal stability, and environmental suitability of these emerging insulation alternatives in high-voltage applications. It is worth noting that a certain discrepancy exists between the simulation temperature in ReaxFF-MD and the actual experimental temperature. By fine-tuning the force field parameters, the time-dependent decomposition rate of the PMVE/CO2 mixture could be monitored within the simulation range, effectively narrowing the gap between the simulated and actual decomposition behavior.
This study employs density functional theory (DFT) to investigate the fundamental properties and molecular stability of PMVE, identifying its potential decomposition sites. Ab initio molecular dynamics (AIMDs) simulations were used to evaluate the ionization energy, electron affinity, and excitation energy of PMVE/CO2 mixtures, aiding in the selection of optimal mixing ratios for engineering use. Thermal decomposition behaviors under varying temperatures and ratios were further explored via ReaxFF molecular dynamics, focusing on key decomposition products and concentrations. Transition state theory and thermodynamic analyses revealed reaction pathways and product distributions, supporting the use of gas chromatography for monitoring equipment status and developing safety measures. These findings establish a link between molecular properties and macroscopic insulation performance, providing theoretical support for PMVE/CO2 as a sustainable insulating gas alternative.

2. Calculation Methods

2.1. Density Functional Theory Simulation

2.1.1. PMVE Model Construction

According to the Gaussian PMVE model, the bond structure of PMVE consists of C=C double bonds, C-F single bonds, and C-O single bonds. The molecular configuration of PMVE is shown in Figure 1. The structure was optimized using the B3LYP/cc-pVTZ basis set. These key bond length parameters are consistent with those reported in reference [22], confirming that the computational results obtained using the B3LYP/cc-pVTZ basis set are reliable.

2.1.2. Calculation Method

Density functional theory (DFT) is a quantum mechanical computational method that focuses on the electron density outside the atomic nucleus. By calculating the ground-state energy, DFT allows for the determination of molecular system properties and the acquisition of molecular electrical characteristics. This theoretical approach has been widely used in studies of SF6 substitute gases [23,24,25]. The fundamental insulating properties of these gases are closely related to their microscopic structures. The stability of dissociation reactions can be evaluated by calculating bond strengths and charge distributions. Furthermore, to validate the decomposition pathways of PMVE, transition state searches and reaction enthalpy calculations are performed.
In this study, the Gaussian 16 software package [26] was used to optimize all molecular structures employing the B3LYP [27] functional along with the cc-pVTZ basis set [28,29], and Grimme’s D3BJ dispersion [30] was also considered. To further improve the accuracy of the results, the zero-point energy (ZPE) and enthalpy corrections were carried out [31]. The formula used to calculate the reaction energy is as follows:
Δ E = E P r o d u c t s E R e a c t a n t s
where ΔE represents the reaction energy, EProducts is the energy of the products calculated after geometric optimization, and EReactants is the energy of the reactants.

2.2. Molecular Dynamics and Reaction Dynamics Simulation

2.2.1. PMVE/CO2 Model Construction

To investigate the macroscopic properties of PMVE/CO2 gas mixture at various temperatures, a periodic cubic structure model was constructed using Materials Studio. The model consists of multiple PMVE and CO2 molecules, with the number of PMVE molecules set to 100 to balance computational efficiency. Table 1 presents the key system parameters and simulation conditions, including density, CO2 composition, and box dimensions. These parameters were chosen to replicate the conditions of the gas at 25 °C and 0.1 MPa. The constructed gas mixture model is shown in Figure 2.

2.2.2. Calculation Method

Reaction dynamics is a simulation method that combines the ReaxFF reaction force field and molecular dynamics developed by van Duin and Goddard [32]. This method makes up for the defect that traditional force fields cannot be applied to complex systems and has been widely used to explore the decomposition process of insulating gases [33]. The ReaxFF reaction force field divides the system energy into several parts for calculation and describes the energy of each part except the bond level on the basis of the bond level. The calculation formula of the system energy is as follows:
E system = E bond + E over + E under + E val      + E pen + E tors + E coa + E H bond      + E conj + E vdWaals + E Coulomb
where Esystem represents the total energy of the system, Ebond is the bond energy, and Eover and Eunder are the energy correction terms for overcoordination and undercoordination, respectively. Eval, Epen, and Ecoa correspond to the bond angle energy term, penalty energy term, and three-body conjugation term, respectively. EH-bond represents the hydrogen bonding energy, Etors refers to the torsion energy, and Econj denotes the four-body energy conjugation. EvdWaals and ECoulomb represent the non-bonding van der Waals and Coulomb interactions, respectively.
In reaction kinetics simulations, an appropriate increase in temperature can accelerate the reaction rate without altering the reaction path or process [34]. Since the time scale in simulations differs significantly from that of real-world reactions, the reaction is typically accelerated by raising the temperature.
To investigate the decomposition mechanism of the PMVE/CO2 mixture, the NVE ensemble was first used to optimize the system for 5 ps at 5 K. This was followed by equilibration using NVT and NPT ensembles for 5 ps at 300 K. Molecular dynamics simulations were then performed for 1000 ps using the NVT ensemble at various temperatures, with the 1000 ps simulation time deemed sufficient to analyze the main cleavage products and cleavage pathways. Periodic boundary conditions were applied throughout the simulation, with temperature controlled using the Berendsen method and a temperature coupling coefficient of 0.1 ps. The simulation time step was set to 0.1 fs [35]. Finally, the enthalpy of the cleavage pathway of the PMVE/CO2 mixture was calculated based on reaction kinetics (ReaxFF-MD), and the potential transition states of the reaction were identified. The difficulty of the cleavage path was assessed from a thermodynamic perspective, validating the reliability of the reaction kinetics simulation results.

2.3. First Principles Simulation

First-principles simulations were conducted using CP2K software to investigate the insulation properties of PMVE/CO2 systems at various mixing ratios. Input files for CP2K were generated using Multiwfn software developed by Dr. Tian Lu [36]. Geometric optimizations employed the PBE functional with D3 dispersion correction and the MOLOPT basis set (DZVP-MOLOPT-SR-GTH). Subsequent energy calculations utilized the same functional but with a more precise basis set (TZVP-MOLOPT-SR-GTH) [37]. The orbital transformation (OT) method was selected for self-consistent field (SCF) convergence [38]. In this study, the ionization energy, electron affinity, and excitation energy of PMVE/CO2 gas mixtures were calculated using the aforementioned methods. Subsequently, wavefunction analyses were performed on all structural models using the Multiwfn software [36]. These analyses aimed to evaluate the insulation properties of the mixed systems.

3. Results and Analysis

3.1. Molecular Structure and Properties of PMVE

3.1.1. Bond Strength Analysis

The bond strength of PMVE can be assessed through factors such as bond length, bond angle, bond type, and Mayer bond order (MBO). The bond length provides an indication of the intermolecular forces, with the longest bonds in the PMVE molecule being the C2–O bond (1.359 Å) and the C3–O bond (1.380 Å). Mayer bond order is commonly used to describe the relative strength of chemical bonds. The C1=C2 bond has the highest bond order of 1.694, indicating a strong interaction between the C1 and C2 atoms. In contrast, the C2–O and C3–O bonds have the smallest bond orders of 1.027 and 1.060, suggesting weak interactions between the C2 and O atoms and the C3 and O atoms. These weak interactions make the C2–O and C3–O bonds more susceptible to breaking under chemical reactions or high-temperature conditions, leading to the formation of radicals such as COF3, C2F3, CF3, and C2F3 (Figure 3).

3.1.2. Charge Distribution

The Fukui function, based on electron density, is commonly used to predict the reactivity of molecules. It reveals the sensitivity of specific points within a molecule to electron attack. The Fukui function is defined as follows:
f + = ( ρ N + δ ( r ) ρ N ( r ) ) δ
f = ( ρ N δ ( r ) ρ N ( r ) ) δ
where N denotes the number of electrons in the reference state of the molecule, and δ is the fraction of the electron [39]. Figure 4 demonstrates the values of Fukui functions mapped onto the surface of the electron density image.
In the case of a chemical reaction, the f+ function indicates regions that are more receptive to electrons and thus more susceptible to nucleophilic attack. As shown in Figure 4a, electrons are more readily acquired near the C1=C2 bond compared to other regions. For the f—function, the red region corresponds to areas where the electron density decreases, making them more vulnerable to electrophilic attack. From Figure 4b, it can be observed that the C2–O and C1–F2 bond positions are more likely to lose electrons.
Based on the previous calculations of bond length and bond order, along with the current results, it is clear that the reactivity of the C2–O and C2–F3 bonds is significantly higher than that of other sites. As a result, the initial cleavage of the PMVE molecule occurs at the C2–O and C2–F3 bond positions.

3.2. Insulating Performance Analysis of PMVE/CO2

During the gas discharge process, ionization, electron affinity, and electron transition play crucial roles [40]. Ionization energy and electron affinity together represent the binding strength of atoms to electrons. A higher ionization energy indicates a stronger atomic binding to electrons. Therefore, the gas is more resistant to breakdown. Excitation energy, however, reflects the energy required for a gas molecule to transition from the ground state to an excited state. By determining the excitation energy of molecules, the difficulty of gas discharge can be analyzed. To investigate the insulation properties of PMVE/CO2, the ionization potential (IP), electron affinity (EA), and excitation energy (EX) of PMVE/CO2 were calculated under different CO2 content conditions. As shown in Table 2, the ionization energy of the gas mixture is highest when the CO2 content is 40%, followed by the mixture with 60% CO2, while pure PMVE gas exhibits the lowest ionization energy. Furthermore, gas mixtures with varying CO2 content do not exhibit electron affinity, and their excitation energies (EX) remain similar, indicating that the addition of CO2 has little effect on the electron excitation process, which is almost negligible.
When the CO2 content is 40%, the PMVE/CO2 gas mixture exhibits significant electrical properties. At this mixing ratio, the ionization energy of the gas mixture is the highest, and its electron affinity is second only to that of pure PMVE gas, indicating that its atoms or molecules have the strongest electron-binding ability. This significantly enhances the insulation performance of the gas. Additionally, the excitation energy of gas mixtures with varying CO2 content remains similar, suggesting that the addition of CO2 has little effect on the electron excitation process. Based on a comprehensive analysis, the PMVE/CO2 gas mixture achieves the best insulation performance at a CO2 content of 40%, making the gas more resistant to breakdown.

3.3. Decomposition Process of PMVE/CO2 Mixture

3.3.1. Effect of Temperature on the Decomposition Process

To further investigate the effect of temperature on the decomposition characteristics of PMVE, molecular dynamics simulations were conducted for the PMVE/CO2 mixture system in the temperature range of 2200 K to 3200 K. Figure 5 illustrates the decomposition behavior of PMVE and CO2 over this temperature range. The results show that the thermal decomposition rate of PMVE increases with temperature. At 2200 K, neither PMVE nor CO2 undergoes significant decomposition. When the temperature rises to 2400 K, the decomposition rate remains slow, with only 8 PMVE molecules and 7 CO2 molecules decomposing within the simulation time. As the temperature increases to 2800 K, the decomposition rate accelerates significantly, and the decomposition of both PMVE and CO2 intensifies. By 3200 K, the decomposition reaction becomes even more pronounced, with 91 PMVE molecules decomposed by the end of the simulation, indicating that the decomposition of PMVE is nearly complete.
Figure 6 shows the variation in total potential energy of the PMVE/CO2 system over time within the temperature range of 2200 K to 3200 K, revealing the energy characteristics during the decomposition reaction. At all temperatures, the total potential energy of the system increases over time, indicating that the decomposition of the PMVE/CO2 mixture is an endothermic process, which aligns with real-world behavior. Below 2600 K, changes in potential energy are minimal, reflecting a low degree of decomposition and a relatively slow process. However, as the temperature rises to 2800 K, the potential energy increases more rapidly between 0 and 200 ps, signifying an acceleration in the decomposition of the gas mixture. At 3000 K, the potential energy rises sharply within the first 500 ps and stabilizes after 700 ps, indicating that the decomposition process is nearly complete. In summary, significant decomposition of the PMVE/CO2 mixture mainly occurs at temperatures of 2800 K and above.
Figure 7 presents the distribution of the maximum quantities of PMVE/CO2 decomposition products across the temperature range of 2200 K to 3200 K. The results show that the decomposition of the PMVE/CO2 gas mixture primarily produces C2F2, COF3, COF2, F, C2F3, CO, CF3, C2F3, and C2F4. Statistical analysis reveals that the overall number of decomposition products increases significantly with rising temperature. At 2200 K, the PMVE/CO2 mixture shows minimal decomposition. Between 2200 K and 2800 K, the main decomposition products are C2F2, COF3, F, and C2F3. As the temperature increases to 3000 K, the production of COF2 begins to rise sharply. By 3200 K, the yields of CO and COF increase significantly, indicating more extensive decomposition.
Figure 8 illustrates the evolution of the quantity of each decomposition product over time at different temperatures. As the temperature increases, the decomposition rate of PMVE accelerates, and the yield of most decomposition products increases accordingly. Below 2400 K, the decomposition rate is extremely slow, with almost no decomposition occurring. However, at 2600 K, the decomposition process begins to accelerate, gradually forming COF3, C2F3, and other products. When the temperature exceeds 2800 K, the decomposition of PMVE intensifies significantly, and the quantities of COF2, COF, C2F2, and CO show a linear growth trend. At 3000 K, the formation rates of COF3 slow down significantly at around 800 ps, indicating a saturation point. At 3200 K, the amount of COF3 starts to decrease at 450 ps, suggesting that COF3 redecomposes at high temperatures. Concurrently, the quantities of COF2 and COF increase substantially, pointing to the correlation between the decrease in COF3 and the formation of COF2 and COF. Additionally, the quantities of F, C2F3, C2F3 and CF3 first increase and then decrease significantly during the reaction. These fluctuations indicate the recombination and redecomposition of free radicals during PMVE decomposition. From a kinetic perspective, this reflects a dynamic equilibrium between the free radicals generated by PMVE decomposition, which helps maintain the insulation performance of the system to some extent. A small amount of C2F4 is produced during the reaction, aligning with experimental findings from previous studies [17], further confirming the reliability of the simulation.

3.3.2. Effect of CO2 Content on the Decomposition Process

To compensate for the high liquefaction temperature of PMVE and enhance its insulation performance, insulating gases are often mixed with inert gases in engineering applications. In this section, ReaxFF-MD simulations are conducted for PMVE/CO2 gas mixtures with CO2 content of 30%, 40%, 50%, and 60% at 3000 K to investigate the effect of CO2 content on the decomposition characteristics of the PMVE/CO2 mixture. Figure 9 presents the time evolution of PMVE decomposition for different CO2 contents at 3000 K. The results show that as the CO2 content in the system increases, the decomposition process of PMVE intensifies. For instance, in the PMVE/CO2 system with 60% CO2, 74 PMVE molecules decompose, while only 37 PMVE molecules decompose in the 30% CO2 system. This trend can be attributed to the increasing density of the gas mixture as PMVE content rises (as shown in Table 1). A denser mixture results in a higher number of molecules per unit volume, which increases the probability of effective molecular collisions, thereby accelerating the reaction.
From Figure 10 and Figure 11, a comparative analysis of product yields under different CO2 concentrations reveals that the yields of COF3, COF2, and C2F2 increase linearly as the CO2 content rises from 30% to 50%. However, when the CO2 content reached 60%, the generation rate of these products began to show signs of saturation. Additionally, under the 60% CO2 condition, the yields of CF3, COF3, and COF2 were significantly higher compared to other mixtures. Notably, CO was only detected when the CO2 content exceeded 50%. Furthermore, the quantities of C2F3 and CF3 fluctuated dynamically over time at different mixing ratios. Overall, the trends in the formation of primary products across different CO2 content conditions were consistent with those observed under varying temperature conditions.

3.4. Decomposition Mechanism of PMVE/CO2

The analysis results in Section 3.3.1 and Section 3.3.2 show that changes in CO2 content do not affect the decomposition mechanism of PMVE. To further elucidate the reaction mechanism of the PMVE/CO2 mixture, the decomposition pathways and reaction enthalpy are presented in Figure 12 and Table 3, respectively. The results reveal that the generation of certain products aligns with prior literature reports [14]. Specifically, three main cleavage pathways for PMVE have been identified. Among these, pathways A1 and A2 require less energy compared to pathway A3. The findings indicate that the C–O bond fracture occurs more easily than the C–F bond fracture during the initial decomposition of PMVE. Additionally, C2F3 transforms into C2F2 upon absorbing 302.51 kJ/mol of energy. Moreover, the B1 to C2 pathways are exothermic, suggesting that the formation of C2F4, C3F6, CF4, and C2F6 is spontaneous. Previous experimental data have confirmed that the primary decomposition products of PMVE include C2F6, C3F6, C3F8, and CF4 [17]. Therefore, the results obtained from the ReaxFF-MD simulation method are highly consistent with observed phenomena, validating the effectiveness and reliability of the simulation approach.

4. Conclusions

Based on density functional theory (DFT) and dynamic simulations, the electrical performance and decomposition characteristics of PMVE/CO2 mixtures are studied theoretically. The effects of temperature and CO2 content on the decomposition process are explored, revealing key insights into the decomposition behavior. The primary decomposition pathways of PMVE are identified, and the simulation results are further validated from a thermodynamic perspective. The conclusions are as follows:
(1) When the CO2 content is 40%, the ionization energy of the gas mixture is highest, indicating that the system is most stable at this mixing ratio. Furthermore, gas mixtures with different CO2 content do not exhibit electron affinity. The excitation energy of different content is similar, which indicates that the addition of CO2 does not affect the electron excitation process.
(2) The primary cleavage of PMVE molecules during thermal decomposition predominantly occurs at the C2–O and C–O bond positions, following three main cleavage pathways. The decomposition rate of the PMVE/CO2 gas mixture increases significantly with rising temperature and CO2 content, and the formation trends of the main decomposition products under different CO2 content conditions are consistent with those observed under varying temperatures. Additionally, a dynamic equilibrium process involving free radicals occurs during PMVE decomposition, which helps maintain the insulation performance of the system.
(3) Compared with SF6/CO2, the PMVE/CO2 mixture exhibits a significant advantage in terms of its extremely low global warming potential (GWP), which is less than one ten-thousandth of that of SF6/CO2, thereby contributing to a much smaller greenhouse effect under normal operating conditions. Furthermore, during the gas decomposition process, SF6 produces highly toxic and corrosive by-products, particularly SO2 and SOF2, which pose greater environmental and safety hazards than those from PMVE. It is important to note that both gases can generate highly corrosive hydrogen fluoride (HF) when in contact with moisture. However, SF6 demonstrates excellent equipment compatibility and is widely used in practice, whereas PMVE still faces challenges related to equipment compatibility before its large-scale application.

Author Contributions

Conceptualization, H.C. and W.Y.; methodology, H.C.; software, W.Y.; validation, H.D. and W.Y.; formal analysis, H.C.; investigation, H.C. and W.Y.; resources, H.D., W.Y. and W.Z.; data curation, H.C.; writing—original draft preparation, H.C. and W.Z.; writing—review and editing, W.Y. and W.Z.; visualization, H.C.; supervision, S.L.; project administration, H.C. and W.Y.; funding acquisition, H.D. and W.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Graduate Innovation Program of China University of Mining and Technology (SJCX25_1459 and SJCX25_1454) and the China Postdoctoral Science Foundation (2024M763542).

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors sincerely thank Tian Lu at the Beijing Kein Research Center for Natural Sciences for his help with computer simulations.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Monges, Y.; Cabañas, J.; Lopez, M.; Morel, S.; Chaparro, E. An integrated testbed IEC61850-IIoT for Gas Insulated SF6 Monitoring in SCADA systems. In Proceedings of the 2022 IEEE ANDESCON, Barranquilla, Colombia, 16–19 November 2022; pp. 1–6. [Google Scholar]
  2. Rokunohe, T.; Yagihashi, Y.; Aoyagi, K.; Oomori, T.; Endo, F. Development of SF6-Free 72.5 kV GIS. IEEE Trans. Power Deliv. 2007, 22, 1869–1876. [Google Scholar] [CrossRef]
  3. Kieffel, Y.; Irwin, T.; Ponchon, P.; Owens, J. Green gas to replace SF6 in electrical grids. IEEE Power Energy Mag. 2016, 14, 32–39. [Google Scholar] [CrossRef]
  4. Nechmi, H.E.; Beroual, A.; Girodet, A.; Vinson, P. Fluoronitriles/CO2 gas mixture as promising substitute to SF6 for insulation in high voltage applications. IEEE Trans. Dielectr. Electr. Insul. 2016, 23, 2587–2593. [Google Scholar] [CrossRef]
  5. Xiao, S.; Zhang, X.; Tang, J.; Liu, S. A review on SF6 substitute gases and research status of CF3I gases. Energy Rep. 2018, 4, 486–496. [Google Scholar] [CrossRef]
  6. Aldy, J.E.; Stavins, R.N. Post-Kyoto International Climate Policy: Implementing Architectures for Agreement; Cambridge University Press: Cambridge, UK, 2009. [Google Scholar]
  7. Stoller, P.; Doiron, C.; Tehlar, D.; Simka, P.; Ranjan, N. Mixtures of CO2 and C5F10O perfluoroketone for high voltage applications. IEEE Trans. Dielectr. Electr. Insul. 2017, 24, 2712–2721. [Google Scholar] [CrossRef]
  8. Beroual, A.; Haddad, A. Recent advances in the quest for a new insulation gas with a low impact on the environment to replace sulfur hexafluoride (SF6) gas in high-voltage power network applications. Energies 2017, 10, 1216. [Google Scholar] [CrossRef]
  9. Rabie, M.; Franck, C.M. Assessment of eco-friendly gases for electrical insulation to replace the most potent industrial greenhouse gas SF6. Environ. Sci. Technol. 2018, 52, 369–380. [Google Scholar] [CrossRef]
  10. Reilly, J.; Prinn, R.; Harnisch, J.; Fitzmaurice, J.; Jacoby, H.; Kicklighter, D.; Melillo, J.; Stone, P.; Sokolov, A.; Wang, C. Multi-gas assessment of the Kyoto Protocol. Nature 1999, 401, 549–555. [Google Scholar] [CrossRef]
  11. Kieffel, Y.; Biquez, F.; Vigouroux, D.; Ponchon, P.; Schlernitzauer, A.; Magous, R.; Cros, G.; Owens, J.G. Characteristics of g3–an alternative to SF6. CIRED 2017, 2017, 54–57. [Google Scholar] [CrossRef]
  12. Preve, C.; Maladen, R.; Piccoz, D. Method for validation of new eco-friendly insulating gases for medium voltage equipment. In Proceedings of the 2016 IEEE International Conference on Dielectrics (ICD), Montpellier, France, 3–7 July 2016; pp. 235–240. [Google Scholar]
  13. Zhang, X.; Li, Y.; Xiao, S.; Tang, J.; Tian, S.; Deng, Z. Decomposition mechanism of C5F10O: An environmentally friendly insulation medium. Environ. Sci. Technol. 2017, 51, 10127–10136. [Google Scholar] [CrossRef] [PubMed]
  14. Sinha, N.; Choi, H.; Song, M.-Y.; Jang, H.-J.; Oh, Y.-H.; Song, K.-D. Perfluoro-methyl-vinyl-ether as SF6 alternative in insulation applications: A DFT study on the physiochemical properties and decomposition pathways. Comput. Theor. Chem. 2023, 1225, 114159. [Google Scholar] [CrossRef]
  15. Park, Y.; Choi, Y.R.; Kim, D.-C.; Kim, Y.; Song, M.-Y.; Kim, Y.-W.; Cho, H.; Jang, H.-J.; Oh, Y.-H.; Song, K.-D. Total electron scattering cross section of C3F6O at the intermediate-energy region for developing an alternative insulation gas to SF6. Curr. Appl. Phys. 2022, 41, 111–115. [Google Scholar] [CrossRef]
  16. Sinha, N.; Song, M.-Y.; Chang, H.; Choi, H.; Jang, H.-J.; Oh, Y.-H.; Song, K.-D. Electron impact cross sections and transport studies of C3F6O. Appl. Sci. 2023, 13, 12612. [Google Scholar] [CrossRef]
  17. Xiao, S.; Chen, Y.; Tang, M.; Tian, S.; Xia, H.; Wang, Y.; Tang, J.; Li, Y.; Zhang, X. Characteristics of perfluoromethyl vinyl ether: A new eco-friendly alternative gas for SF6. High Volt. 2024, 9, 509–517. [Google Scholar] [CrossRef]
  18. Zhao, H.; Li, X.; Zhu, K.; Wang, Q.; Lin, H.; Guo, X. Study of the arc interruption performance of SF6-CO2 mixtures as a substitute for SF6. IEEE Trans. Dielectr. Electr. Insul. 2016, 23, 2657–2667. [Google Scholar] [CrossRef]
  19. Zhuang, X.; Zhi, L.; Huang, C. The resuming insulation discharge failure and conductor overheating failure judgment method of SF6 electrical equipment. Energy Power Eng. 2009, 1, 122. [Google Scholar] [CrossRef]
  20. Ono, R.; Oda, T. Measurement of gas temperature and OH density in the afterglow of pulsed positive corona discharge. J. Phys. D Appl. Phys. 2008, 41, 035204. [Google Scholar] [CrossRef]
  21. Zhang, Y.; Li, Y.; Zhang, X.; Xiao, S.; Tang, J. Insights on decomposition process of c-C4F8 and c-C4F8/N2 mixture as substitutes for SF6. R. Soc. Open Sci. 2018, 5, 181104. [Google Scholar] [CrossRef]
  22. Kondo, Y.; Ishikawa, K.; Hayashi, T.; Sekine, M.; Hori, M. Electron impact ionization of perfluoro-methyl-vinyl-ether C3F6O. Plasma Sources Sci. Technol. 2018, 27, 015009. [Google Scholar] [CrossRef]
  23. Yu, X.; Hou, H.; Wang, B. Prediction on dielectric strength and boiling point of gaseous molecules for replacement of SF6. J. Comput. Chem. 2017, 38, 721–729. [Google Scholar] [CrossRef]
  24. Sulbaek Andersen, M.P.; Kyte, M.; Andersen, S.T.; Nielsen, C.J.; Nielsen, O.J. Atmospheric Chemistry of (CF3)2CF–CN: A Replacement Compound for the Most Potent Industrial Greenhouse Gas, SF6. Environ. Sci. Technol. 2017, 51, 1321–1329. [Google Scholar] [CrossRef]
  25. Fu, Y.; Rong, M.; Yang, K.; Yang, A.; Wang, X.; Gao, Q.; Liu, D.; Murphy, A.B. Calculated rate constants of the chemical reactions involving the main byproducts SO2F, SOF2, SO2F2 of SF6 decomposition in power equipment. J. Phys. D Appl. Phys. 2016, 49, 155502. [Google Scholar] [CrossRef]
  26. Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Petersson, G.; Nakatsuji, H. Gaussian 16 Rev. C. 01, 2016; b) DJ Wales, JPK Doye. J. Phys. Chem. A 1997, 101, 5111–5116. [Google Scholar]
  27. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  28. Kendall, R.A.; Dunning, T.H., Jr.; Harrison, R.J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef]
  29. Rusakov, Y.Y.; Rusakovai, L. Getaway from the Geometry Factor Error in the Molecular Property Calculations: Efficient pecG-n (n = 1, 2) Basis Sets for the Geometry Optimization of Molecules Containing Light p Elements. J. Chem. Theory Comput. 2024, 20, 6661–6673. [Google Scholar] [CrossRef]
  30. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  31. Lu, T.; Chen, Q. Shermo: A general code for calculating molecular thermochemistry properties. Comput. Theor. Chem. 2021, 1200, 113249. [Google Scholar] [CrossRef]
  32. Van Duin, A.C.; Dasgupta, S.; Lorant, F.; Goddard, W.A. ReaxFF: A reactive force field for hydrocarbons. J. Phys. Chem. A 2001, 105, 9396–9409. [Google Scholar] [CrossRef]
  33. Zhao, D.; Yan, J.; He, R.; Geng, Y.; Liu, Z.; Wang, J. Decomposition mechanism of C4F7N/CO2 gas mixture based on molecular dynamics and effect of O2 content. J. Appl. Phys. 2024, 135, 024401. [Google Scholar] [CrossRef]
  34. Chenoweth, K.; Cheung, S.; Van Duin, A.C.; Goddard, W.A.; Kober, E.M. Simulations on the thermal decomposition of a poly (dimethylsiloxane) polymer using the ReaxFF reactive force field. J. Am. Chem. Soc. 2005, 127, 7192–7202. [Google Scholar] [CrossRef] [PubMed]
  35. Berendsen, H.J.; Postma, J.v.; Van Gunsteren, W.F.; DiNola, A.; Haak, J.R. Molecular dynamics with coupling to an external bath. J. Chem. Phys. 1984, 81, 3684–3690. [Google Scholar] [CrossRef]
  36. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  37. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  38. VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. Quickstep: Fast and accurate density functional calculations using a mixed Gaussian and plane waves approach. Comput. Phys. Commun. 2005, 167, 103–128. [Google Scholar] [CrossRef]
  39. Yang, W.; Parr, R.G. Hardness, softness, and the fukui function in the electronic theory of metals and catalysis. Proc. Natl. Acad. Sci. USA 1985, 82, 6723–6726. [Google Scholar] [CrossRef] [PubMed]
  40. Zhan, C.-G.; Nichols, J.A.; Dixon, D.A. Ionization potential, electron affinity, electronegativity, hardness, and electron excitation energy: Molecular properties from density functional theory orbital energies. J. Phys. Chem. A 2003, 107, 4184–4195. [Google Scholar] [CrossRef]
Figure 1. Molecular structure of PMVE.
Figure 1. Molecular structure of PMVE.
Applsci 15 09011 g001
Figure 2. Structure of PMVE/CO2 system.
Figure 2. Structure of PMVE/CO2 system.
Applsci 15 09011 g002
Figure 3. MBO of PMVE molecules.
Figure 3. MBO of PMVE molecules.
Applsci 15 09011 g003
Figure 4. Fukui functions of the PMVE molecule.
Figure 4. Fukui functions of the PMVE molecule.
Applsci 15 09011 g004
Figure 5. Time evolution of PMVE and CO2 decomposition at 2200–3200 K.
Figure 5. Time evolution of PMVE and CO2 decomposition at 2200–3200 K.
Applsci 15 09011 g005
Figure 6. Time evolution of potential energy at 2200–3200 K.
Figure 6. Time evolution of potential energy at 2200–3200 K.
Applsci 15 09011 g006
Figure 7. Maximum number of decomposition products of PMVE/CO2 at 2200–3200 K.
Figure 7. Maximum number of decomposition products of PMVE/CO2 at 2200–3200 K.
Applsci 15 09011 g007
Figure 8. Time evolution of PMVE/CO2 decomposition products at 2000−3000 K.
Figure 8. Time evolution of PMVE/CO2 decomposition products at 2000−3000 K.
Applsci 15 09011 g008aApplsci 15 09011 g008b
Figure 9. Time evolution of PMVE decomposition at 3000 K with 30–60% CO2 content.
Figure 9. Time evolution of PMVE decomposition at 3000 K with 30–60% CO2 content.
Applsci 15 09011 g009
Figure 10. Time evolution of PMVE/CO2 decomposition products at 3000 K with 30–60% CO2 content.
Figure 10. Time evolution of PMVE/CO2 decomposition products at 3000 K with 30–60% CO2 content.
Applsci 15 09011 g010aApplsci 15 09011 g010b
Figure 11. Maximum number of decomposition products of PMVE/CO2 at 3000 K with 30–60% CO2 content.
Figure 11. Maximum number of decomposition products of PMVE/CO2 at 3000 K with 30–60% CO2 content.
Applsci 15 09011 g011
Figure 12. Decomposition mechanism of the PMVE/CO2 gas mixture (gray, red, and blue for C, O, and F atoms, respectively).
Figure 12. Decomposition mechanism of the PMVE/CO2 gas mixture (gray, red, and blue for C, O, and F atoms, respectively).
Applsci 15 09011 g012
Table 1. Parameters of PMVE/CO2 system.
Table 1. Parameters of PMVE/CO2 system.
No.CO2 ContentPMVECO2Density (g/cm3)Box Length (Å)
10%10000.00690214.1
230%1001600.00297236.4
340%1002500.00217276.4
450%1003800.00194301.5
560%1005700.00177339.4
Table 2. Ionization potential, electron affinity, and excitation energy of PMVE/CO2.
Table 2. Ionization potential, electron affinity, and excitation energy of PMVE/CO2.
CO2 ContentIP (eV)EA (eV)EX (eV)
0%5.29−0.425.16
30%5.75−1.285.06
40%11.72−0.855.03
50%6.16−2.275.06
60%7.04−2.435.10
Table 3. Reaction pathways and reaction enthalpy, barrier of PMVE/CO2 gas mixture.
Table 3. Reaction pathways and reaction enthalpy, barrier of PMVE/CO2 gas mixture.
NoReaction PathwaysReaction Enthalpy (KJ/mol)
A1C3F6O → C2F3 + COF3433.91
A2C3F6O → C2OF3 + CF3197.16
A3C3F6O → C3F5O + F488.14
B1C2F3 → C2F2 + F302.51
B2C2F3 + F → C2F4−506.77
B3C2F3 + CF3 → C3F6−431.58
C1CF3 + F → CF4−509.67
C22CF3 → C2F6−372.48
D1C3F5O → COF3 + C2F2248.28
E1COF3 → COF2 + F93.90
E2COF2 → COF + F482.59
E3COF → CO + F155.29
F1CO2 → CO + O801.64
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

Dong, H.; Chu, H.; Zeng, W.; Liu, S.; Ye, W. Molecular Insights into the Insulating and Pyrolysis Properties of Environmentally Friendly PMVE/CO2 Mixtures: A Collaborative Analysis Based on Density Functional Theory and Reaction Kinetics. Appl. Sci. 2025, 15, 9011. https://doi.org/10.3390/app15169011

AMA Style

Dong H, Chu H, Zeng W, Liu S, Ye W. Molecular Insights into the Insulating and Pyrolysis Properties of Environmentally Friendly PMVE/CO2 Mixtures: A Collaborative Analysis Based on Density Functional Theory and Reaction Kinetics. Applied Sciences. 2025; 15(16):9011. https://doi.org/10.3390/app15169011

Chicago/Turabian Style

Dong, Haibo, Haonan Chu, Wentian Zeng, Shicheng Liu, and Wenyu Ye. 2025. "Molecular Insights into the Insulating and Pyrolysis Properties of Environmentally Friendly PMVE/CO2 Mixtures: A Collaborative Analysis Based on Density Functional Theory and Reaction Kinetics" Applied Sciences 15, no. 16: 9011. https://doi.org/10.3390/app15169011

APA Style

Dong, H., Chu, H., Zeng, W., Liu, S., & Ye, W. (2025). Molecular Insights into the Insulating and Pyrolysis Properties of Environmentally Friendly PMVE/CO2 Mixtures: A Collaborative Analysis Based on Density Functional Theory and Reaction Kinetics. Applied Sciences, 15(16), 9011. https://doi.org/10.3390/app15169011

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