Abstract
This study evaluates how next-generation silicone impression materials intended for dental use behave during polymerization, as well as their dimensional stability, mechanical properties, degradation patterns, and in silico toxicity levels. Silicone materials are preferred for dental applications because of their outstanding mechanical properties and compatibility with biological tissues. The performance of these materials is susceptible to environmental conditions including temperature changes, humidity levels, and exposure to oral fluids. Patient safety requires evaluation of degradation product toxicity concerns. It is crucial to examine these properties at the molecular level to enhance material durability and safety during clinical use. The structural, mechanical, and stability properties of silicone materials were modeled through molecular dynamics (MD) simulations using BIOVIA Materials Studio 2020. Material characterization and evaluation of mechanical properties were performed with the Forcite module using the COMPASSIII force field. The study simulated polymerization dynamics to understand the reaction mechanisms while employing the Kinetix and DMol3 modules to analyze dimensional stability under various environmental stresses. The CASTEP and DMol3 modules, along with the OSIRIS DataWarrior, were employed to forecast degradation pathways and potential toxicity. The combination of an elastic modulus of 2.533 GPa and tensile strength of 5.387 MPa allows Polydimethylsiloxane (PDMS) to show superior flexibility and rigidity, which qualifies it as the best choice for dental impression materials. Methacryloxypropyltrimethoxysilane (3.248 GPa) and hexaphenylcyclotrisiloxane (3.017 GPa) exhibited enhanced stiffness, suggesting their usefulness in load-bearing scenarios. In silico toxicity predictions indicated that most silicone derivatives demonstrated acceptable biocompatibility, although some silane compounds showed potential risks requiring further experimental validation. Under simulated conditions, the materials maintained stable configurations and exhibited positive polymerization dynamics, indicating that they could provide high durability along with dimensional stability for dental usage. This study highlights the superior balance of flexibility, rigidity, and safety exhibited by PDMS, while also identifying Methacryloxypropyltrimethoxysilane and hexaphenylcyclotrisiloxane as candidates for specialized load-bearing dental applications. Promising in silico findings require experimental validation and clinical testing to establish their practical applications.

Introduction
The development of innovative materials has driven remarkable advancements in the field of dentistry, resulting in better patient outcomes. Recent advances in dental restorative composites, such as the incorporation of aryloxyphosphazene derivatives and carboxy phosphazene methacrylates, have demonstrated enhanced adhesion, improved mechanical strength, and reduced solubility, further highlighting the importance of material innovation in dentistry [1, 2]. Due to their outstanding flexibility, dimensional stability, biocompatibility and ease of manipulation silicone-based impression materials have become essential in clinical dental practices [3, 4]. Various silane-based monomers have been used to produce silicone impression materials that form cross-linked polymer networks via polymerization reactions [5]. Silicone impression materials have been synthesized from various silane-based monomers that undergo polymerization reactions to form cross-linked polymer networks [6, 7]. The polymerization procedure that determines both the mechanical properties and stability of the material depends on multiple factors such as the chemical structure of the silane components. The specific silane monomer chosen had a substantial impact on the final properties of the material including its dimensional accuracy, hardness, elastic modulus and long-term stability [8, 9]. A comprehensive knowledge of silane polymerization kinetics became crucial to engineer materials possessing the ideal characteristics required for clinical applications.
The justification for selecting silicone-based compounds as dental impression materials lies in their ability to reproduce fine anatomical details with exceptional accuracy, combined with their favorable viscoelastic behavior that allows easy removal from the oral cavity without distortion [3, 10]. Furthermore, unlike alginate or polyether materials, silicones exhibit superior dimensional stability over extended time periods, enabling delayed pouring of impressions without significant loss of accuracy. Their hydrophobicity, chemical inertness, and demonstrated long-term biocompatibility further strengthen their suitability for clinical use, making them the material of choice in prosthodontics, restorative dentistry, and implantology [11, 12].
The formulation of dental impression materials frequently involves silicones, such as polydimethylsiloxane (PDMS), vinyltrimethoxysilane, and methacryloxypropyltrimethoxysilane. These materials provide excellent flexibility and surface detail reproduction, yet show different polymerization rates and structural transformations during curing, which impacts the final outcome [13, 14]. For instance, PDMS, one of the most widely used silicone materials, forms highly flexible and stable networks that provide long-term dimensional stability [15]. A major difficulty encountered in the development of next-generation silicone impression materials is maintaining their superior qualities while minimizing their toxicity potential [16]. The molecular dynamics (MD) of these materials during the polymerization process remain an active research area because of the need for improved curing times and reduced shrinkage while enhancing the overall material performance. A major difficulty encountered in the development of next-generation silicone impression materials is maintaining their superior qualities while minimizing their toxicity potential [17]. While silicones typically demonstrated biocompatibility silicones could release toxic byproducts from some silane compounds which might irritate sensitive individuals [5, 18]. Therefore, the main objective of this study was to apply MD simulations and in silico toxicity assessments to evaluate polymerization kinetics, dimensional stability, mechanical performance, and biocompatibility of next-generation silicone impression materials. The central hypothesis tested was that variations in the chemical structure of silane monomers directly influence polymerization behavior, mechanical strength, dimensional stability, and predicted toxicity profiles, thereby determining their suitability for dental applications.
In this study, toxicity was evaluated using an in silico approach, which offers several advantages: it avoids ethical concerns of animal testing, enables rapid high-throughput screening of candidate compounds, and provides molecular-level predictions of toxicity pathways before clinical testing. Toxicity prediction models alongside MD simulations offer a crucial understanding of the potential risks presented by various silane-based compounds. This approach protected the patients and led to the creation of safer dental materials.
Molecular dynamics simulations offer a key advantage for studying these behaviors, as they allow atomistic-level visualization of polymerization and stress responses, providing predictive insights that are difficult to obtain experimentally. Previous studies [19–21] have shown that MD enables accurate modeling of dimensional stability under physiological conditions and identification of structural instabilities that affect clinical performance, supporting its use in this research. Through behavior simulations of atoms and molecules during polymerization, scientists can predict how a polymer network will form, as well as determine the mechanical properties of the final material and possible structural anomalies or instabilities [22, 23]. The computational method delivered precise insights that were unattainable by experimental approaches alone, which proved essential for developing and assessing dental materials. This research employed MD simulations to examine polymerization kinetics along with dimensional stability and toxicity profiles in next-generation silicone impression materials. In this study, numerous silane compounds were investigated, including aminopropyltriethoxysilane, butyltrichlorosilane, decamethylcyclopentasiloxane, hexamethyldisiloxane, methyldiphenylsilane. Researchers have chosen these silane compounds because they appear frequently in dental impression material formulations and their different chemical structures produce unique polymerization behaviors and material characteristics. The novelty of this study lies in its integrated in silico approach, which simultaneously evaluates polymerization kinetics, mechanical properties, degradation behavior, and toxicity profiles of silicone-based dental materials. Unlike earlier works that focused on individual aspects such as curing behavior or dimensional stability, our study provides a holistic molecular-level assessment that directly links structural features of silane monomers to clinically relevant properties. This research unveiled essential findings about curing mechanisms through MD simulations of materials, while also evaluating their dimensional stability over time and toxicity potential when interacting with human tissues.
Methodology
Selection of silicones for investigation
Silicones were selected based on factors such as dental application relevance and polymerization characteristics while also considering their ability to reveal information about the behavior and safety of new silicone impression materials. Silicone compounds with multiple chemical structures and functional groups were selected to represent the different types of silanes and siloxanes typically used in dental material applications. Table 1 lists the selected silicones, divided by their chemical structure and functional properties. Aminopropyltriethoxysilane, butyltrichlorosilane, and methacryloxypropyltrimethoxysilane served as coupling agents because they enable bonding between inorganic fillers and organic polymers within silicone-based impression materials [24]. decamethylcyclopentasiloxane and dodecamethylcyclohexasiloxane were chosen because their unique molecular structures may improve material flexibility and stability [25]. This study investigated the effects of two frequently utilized siloxane substances, Polydimethylsiloxane and hexamethyldisiloxane, on the mechanical and dimensional features of silicone impression materials.
Table 1.
Selected silicones with their categories, chemical structures, and densities
| Silicone | Category | Chemical Formula | Density (g/cm3) |
|---|---|---|---|
| Aminopropyltriethoxysilane | silane coupling agent | C9H21NO3Si | 0.94 |
| Butyltrichlorosilane | silane coupling agent | C4H11Cl3Si | 1.19 |
| Chlorotrimethylsilane | silane coupling agent | C3H7Cl3Si | 1.23 |
| Cyclohexyltrimethoxysilane | silane coupling agent | C9H18O3Si | 0.98 |
| Decamethylcyclopentasiloxane | cyclic siloxane | C10H30O5Si5 | 0.87 |
| Diethylsilane | silane coupling agent | C4H12Si | 0.83 |
| Dimethyldichlorosilane | silane coupling agent | C2H6Cl2Si | 1.18 |
| Dodecamethylcyclohexasiloxane | cyclic siloxane | C12H36O6Si6 | 0.89 |
| Hexamethyldisiloxane | siloxane | C6H18O2Si2 | 0.77 |
| Hexaphenylcyclotrisiloxane | cyclic siloxane | C18H15Si3 | 1.25 |
| Methacryloxypropyltrimethoxysilane | silane coupling agent | C10H20O4Si | 1.02 |
| Methyldiphenylsilane | silane coupling agent | C13H14Si | 1.14 |
| Methyltrichlorosilane | silane coupling agent | C2H6Cl3Si | 1.24 |
| Octamethylcyclotetrasiloxane | cyclic siloxane | C8H24O4Si4 | 0.83 |
| Octamethyltrisiloxane | siloxane | C8H24O3Si3 | 0.8 |
| Phenyldimethylsilane | silane coupling agent | C13H14Si | 1.1 |
| Phenyltrimethoxysilane | silane coupling agent | C9H12O3Si | 1.1 |
| Polydimethylsiloxane | siloxane | (C2H6OSi)n | 0.97 |
| Tetraethoxysilane | silane coupling agent | C8H20O4Si | 0.96 |
| Tetravinyltetramethylcyclotetrasiloxane | cyclic siloxane | C12H24O4Si4 | 0.85 |
| Triethoxysilane | silane coupling agent | C6H18O3Si | 0.92 |
| Trifluoropropyltrimethoxysilane | silane coupling agent | C9H21F3O3Si | 1.22 |
| Triisopropylsilane | silane coupling agent | C9H21Si | 0.85 |
| Trimethoxysilane | silane coupling agent | C3H10O3Si | 1.05 |
| Vinyltrimethoxysilane | silane coupling agent | C8H18O3Si | 0.98 |
All chemical structures and formulas were retrieved from the PubChem database; therefore, commercial brand names and batch numbers are not applicable
The chosen silicones encompass a diverse range of molecular architectures, including basic linear siloxanes, such as trimethoxysilane, and sophisticated cyclic siloxanes, such as hexaphenylcyclotrisiloxane. Table 1 lists the chemical formulas and densities of each compound. By studying different silicones, researchers have gained full insight into how chemical structures affect polymerization behavior and dimensional stability, as well as the potential toxicity of the final materials. The density measurements of silicones between 0.77 g/cm3 and 1.25 g/cm3 offer essential information regarding their bulk properties which determine how well they fit dental molds and maintain stability after polymerization. The different polymerization kinetics and mechanical properties resulting from this wide range of density values are essential to evaluate the suitability of these materials for dental applications. This study investigated the potential effects of integrating multiple functional groups, such as -methoxy, -chloro, and -phenyl, on the performance metrics of silicone materials. This study investigated the relationship between the chemical composition and performance of dental impression materials by examining silicones with diverse chemical structures and functional groups. The selected silicone compounds demonstrated distinct reactivity profiles, along with specific crosslinking behaviors and toxicity potentials, which established a strong foundation for assessing their potential as next-generation silicone impression materials.
3D structure modeling of silicones with MM2 energy minimization
Three-dimensional models of the selected silicones were built with precise details to accurately depict their spatial arrangements and molecular conformations. Scientists have created initial 3D models by employing sophisticated molecular modeling techniques that have laid the groundwork for studying their geometric features. Chem3D software (PerkinElmer Inc.) was used to optimize the molecular structures and improve the model accuracy through further refinement. The optimization process produced precise 3D geometric representations of the ligands by documenting essential molecular architecture details, including bond lengths, angles, and torsional strains, that support computational analysis reliability [26]. The molecular structures underwent further stabilization through an energy-minimization step after 3D generation and optimization in Chem3D. The MM2 force field [27, 28] was used in this step because it has gained recognition for its ability to evaluate the potential energy within molecular systems and minimize structural energy through atomic position adjustments. The MM2 force field was selected in this study because of its well-established ability to optimize organic molecules and silane derivatives with computational efficiency [28], providing reliable preliminary conformations at a fraction of the computational cost compared to Density Functional Theory (DFT). While DFT is a more modern quantum chemical method that can yield highly accurate electronic and structural information, its intensive computational requirements make it less practical for initial large-scale modeling of multiple silicone structures. Thus, MM2 offered a pragmatic balance between computational feasibility and structural reliability for the scope of this work. ChemDraw Professional 20.1.1 from PerkinElmer Inc. performed the energy minimization procedure using its integrated MM2 force field for accurate energy calculation. The use of this dual-energy minimization approach, which includes both 3D structure optimization and energy minimization, is essential for refining silicone models and achieving energetic stability. Through the reduction of strain and adjustment of bond angles and torsions, this method generated stable and reliable structures that could be used as reliable initial conditions for advanced computational research, including MD simulations and toxicity evaluations.
Molecular dynamics (MD) simulations
This research utilized MD simulations to explore the structural, mechanical, and stability characteristics of various silicone materials through polymerization dynamics, dimensional stability measurements, and mechanical property assessments alongside degradation behavior analysis under oral environment simulations. The researchers used BIOVIA Materials Studio 2020 [29] to run simulations because it offered numerous advanced modules that supported an in-depth analysis of silicones. The research utilized a 40 × 40 × 40 Å3 simulation box to achieve a representative sample volume, which enabled the precise modeling of polymeric structures and their interactions across multiple conditions. The selected box size balances the computational efficiency with a sufficient spatial representation of the material properties at the atomic scale.
Material characterization and mechanical properties
In this study, the characterization of materials and evaluation of their mechanical properties were performed through a unified simulation approach using the Forcite module from BIOVIA Materials Studio 2020 (Dassault Systèmes, Vélizy-Villacoublay, France) [29]. This methodological framework employs force-field-based simulations with the COMPASSIII force field to describe atomic interactions, combined with the Gasteiger charge method to ensure accurate charge assignments, thereby enabling the systematic investigation of the molecular structures and mechanical properties of various silicone materials. The NVT ensemble maintained a constant volume and temperature of 310 K to reflect physiological body temperature during material characterization and mechanical simulations, while employing the Berendsen barostat to hold pressure at 1 bar. The simulation utilized a time step of 100 fs and was extended for 5000 fs to accurately capture the molecular evolution and stress responses of the materials. To specifically study the mechanical properties, a uniaxial tensile deformation was applied along the z-axis of the simulation cell. The stress–strain was obtained by gradually increasing the strain while monitoring the corresponding stress values. From this, the tensile strength (MPa) was determined as the maximum stress prior to failure, and the elastic modulus (GPa) was calculated from the initial linear region of the stress–strain response. A uniaxial stress application enabled tensile testing simulation, which provided measurements of the tensile strength (MPa) and elastic modulus (GPa) through analysis of the material stress response. Additionally, the simulation monitored bond stretching, angle bending, and non-bonded interactions during deformation to provide atomistic insight into material failure mechanisms. The simulation monitored how the material extended and replicated the surface details under stress, to determine its appropriateness for precise dental applications. Through the unified method, researchers gain comprehensive insights into silicone elasticity and mechanical performance, which are essential for assessing their suitability as dental materials that require structural integrity together with flexibility.
Polymerization dynamics
To simulate polymerization kinetics, researchers have employed Forcite to model the reaction mechanisms and determine the factors that affect polymerization shrinkage and polymerization time. The COMPASSIII force field enabled the precise depiction of atomic interactions during polymerization, while the NVT ensemble and temperature setting of 310 K maintained real-world process conditions during the simulations. The polymerization rate (mol/s) and polymerization energy barrier (kJ/mol) values emerged from the analysis of the polymerization reaction progress and the determination of the transition states and activation energies.
Dimensional stability
The Kinetix and DMol3 modules were utilized to model the dimensional stability of silicone materials when subjected to environmental stress factors such as temperature changes and humidity variations. The Kinetix module analyzed the time-dependent behavior of the material by examining its response to temperature changes and humidity, together with other environmental variables. The DMol3 module uses DFT to conduct quantum mechanical calculations to examine the atomic-level electronic properties and structural stability of the material, while revealing factors that lead to dimensional changes. The sorption module was used to perform molecular-level assessments of the water absorption and hydrophobic properties of the material. The module conducts simulations to measure the water absorption rate of the material (mg/mm3) while determining the contact angle to assess the hydrophobic characteristics of the material. The Sorption module used water surface interaction simulations to investigate moisture sensitivity and determine the water absorption potential in oral environments.
Degradation
The CASTEP and DMol3 modules were utilized to predict the chemical stability and degradation pathways of silicones under simulated oral conditions. Researchers have applied the first-principles quantum mechanical simulation tool, CASTEP, to study atomic-level degradation processes under both thermal and chemical influences. The DMol3 module enhances the analysis by providing supplementary information about the electronic characteristics and reaction behaviors of silicones when subjected to different conditions, such as oral fluid exposure and temperature changes. Through simulations, researchers predicted how silicone materials would degrade when exposed to oral conditions throughout their usage period to evaluate their durability and performance stability. This study gained complete insight into the polymerization kinetics, dimensional stability, mechanical properties, and degradation behavior of silicone materials through the integration of multiple simulation results. The results of MD simulations performed with BIOVIA Materials Studio modules formed the initial groundwork for enhancing and engineering silicone-based dental materials.
In silico toxicity assessment of silicones
This methodology relies on advanced in silico tools to systematically evaluate the safety profiles of silicones. Advanced algorithms employed by computational tools perform analyses of molecular descriptors that serve as quantitative representations that reveal detailed structural characteristics of molecules across various chemical properties. Molecular descriptors cover diverse information, including properties such as cLogP, which measures the hydrophobicity of compounds through the octanol-water partition coefficient [30–32], TPSA (Topological Polar Surface Area, an indicator of the molecule’s polarity and potential for permeation [33]), Molecular Shape Index, which describes molecular shape and biological target interaction potential [34], Molecular Flexibility (which reflects the ability of the molecule to adopt various conformations [35]), and various toxicity indicators such as Mutagenic, Tumorigenic, and irritant potential. The descriptors work together to forecast silicone behavior inside biological systems while also identifying possible risks. The researchers utilized the OSIRIS DataWarrior V6.4 [36] to evaluate toxicity during this study. OSIRIS DataWarrior stands out as a powerful analytical tool due to its extensive database combined with strong predictive features. The software displays a strong performance in assessing multiple molecular descriptors that reveal structural characteristics that can predict toxicity risks. This software evaluates toxicological risks by analyzing molecular descriptors that provide insights into mutagenicity (genetic mutation potential), tumorigenicity (cancer formation potential), reproductive toxicity (fertility and reproductive health effects), and irritation (adverse reactions to skin or eyes). These elements play a key role in determining the safety profile of silicones for biomedical applications.
Results
Results of 3D structure modeling of silicones with MM2 energy minimization
The MM2 force field’s energy minimization data reveals comprehensive details about the molecular interactions present in various silicone compounds. Table 2 provides the breakdown of energy components for each compound, which consists of stretch, bend, stretch-bend, torsion, non-1,4 van der Waals (VDW), 1,4 VDW, and dipole-dipole interactions, providing information about the energy stability and favorability across different structures. The total energy values varied widely among the silicone compounds, demonstrating their structural diversity. Aminopropyltriethoxysilane demonstrates a high total energy level of 8.62 kcal/mol. The significant dipole–dipole interaction of 1.88 kcal/mol along with strong 1,4 VDW interactions which contribute 9.37 kcal/mol lead to the formation of a stable molecular configuration that shows strong dipolar characteristics. Butyltrichlorosilane demonstrates a significantly reduced total energy of 0.80 kcal/mol because its dipole-dipole interactions remain at 0 kcal/mol and its 1,4 VDW interactions stay low. Decamethylcyclopentasiloxane and dodecamethylcyclohexasiloxane represent highly substituted siloxane compounds which exhibit significant negative total energies of −10.28 kcal/mol and −17.01 kcal/mol, respectively. The combination of intense 1,4 VDW forces and significant dipole–dipole interactions indicates that their complex molecular structures result in increased thermodynamic stability.
Table 2.
Energy minimization components and total energy of various silicones using MM2 method
| Material | Energy minimization components | |||||||
|---|---|---|---|---|---|---|---|---|
| Stretch | Bend | Stretch-Bend | Torsion | Non-1,4 VDW | 1,4 VDW | Dipole–Dipole | Total Energy | |
| Aminopropyltriethoxysilane | 0.6210 | 3.0636 | 0.1879 | −0.1304 | −6.3720 | 9.3736 | 1.8779 | 8.6216 |
| Butyltrichlorosilane | 0.1842 | 0.4108 | 0.0446 | 0.0086 | −1.1567 | 1.3058 | 0.0000 | 0.7973 |
| Chlorotrimethylsilane | 0.0015 | 0.3037 | −0.0023 | 0.0007 | −0.7452 | −1.7317 | 0.0000 | −2.1733 |
| Cyclohexyltrimethoxysilane | 0.7643 | 3.7716 | 0.1576 | 2.4691 | −5.7315 | 9.8459 | 2.4416 | 13.7186 |
| Decamethylcyclopentasiloxane | 0.0841 | 1.7960 | −0.0764 | 0.6471 | −17.1038 | −7.2583 | 11.6280 | −10.2833 |
| Diethylsilane | 0.0748 | 0.8042 | 0.0210 | 0.4595 | −1.1991 | 0.8603 | 0.0000 | 1.0206 |
| Dimethyldichlorosilane | 0.0011 | 0.1866 | −0.0024 | 0.0000 | −0.2468 | −1.5005 | 0.0000 | −1.5620 |
| Dodecamethylcyclohexasiloxane | 0.0744 | 1.4605 | −0.0562 | 0.5772 | −24.5696 | −8.7929 | 14.2921 | −17.0146 |
| Hexamethyldisiloxane | 0.0078 | 0.6288 | −0.0041 | 0.0272 | −4.6228 | −3.1079 | 2.4781 | −4.5929 |
| Hexaphenylcyclotrisiloxane | 1.5860 | 3.9826 | −0.6225 | −30.5010 | −17.2182 | 33.4614 | 10.0692 | 0.7574 |
| Methacryloxypropyltrimethoxysilane | 1.3113 | 6.2813 | 0.1709 | −0.6920 | −4.0428 | 12.8798 | 2.5309 | 18.4394 |
| Methyldiphenylsilane | 0.3694 | 0.3628 | −0.0388 | −11.7574 | −2.9072 | 11.1815 | 0.0000 | −2.7896 |
| Methyltrichlorosilane | 0.0009 | 0.0920 | −0.0027 | 0.0000 | 0.0000 | −0.9242 | 0.0000 | −0.8339 |
| Octamethylcyclotetrasiloxane | 0.0770 | 1.3275 | −0.0729 | 0.2375 | −10.5843 | −5.9693 | 9.4144 | −5.5701 |
| Octamethyltrisiloxane | 0.0204 | 0.9455 | −0.0184 | 0.1058 | −8.4673 | −4.5653 | 4.7609 | −7.2185 |
| Phenyldimethylsilane | 0.1920 | 0.2894 | −0.0324 | −7.0741 | −1.7811 | 5.1887 | 0.0000 | −3.2175 |
| Phenyltrimethoxysilane | 0.5929 | 2.9154 | −0.0041 | −5.5706 | −4.6965 | 11.2780 | 1.8550 | 6.3702 |
| Polydimethylsiloxane | 0.0932 | 2.4370 | −0.1008 | 1.0971 | −36.6028 | −11.8240 | 15.5542 | −29.3460 |
| Tetraethoxysilane | 0.8538 | 4.1687 | 0.0226 | 0.0002 | −6.1454 | 11.0074 | 1.9341 | 11.8414 |
| Tetravinyltetramethylcyclotetrasiloxane | 0.1346 | 1.6101 | −0.0643 | −0.7084 | −11.5991 | −2.9561 | 9.7306 | −3.8527 |
| Triethoxysilane | 0.3775 | 2.3600 | 0.2744 | 0.3396 | −3.0765 | 8.6889 | 1.9966 | 10.9604 |
| Trifluoropropyltrimethoxysilane | 0.4657 | 3.4536 | 0.1308 | −0.0982 | −4.3242 | 7.0267 | 1.2593 | 7.9138 |
| Triisopropylsilane | 0.5442 | 1.4418 | 0.1769 | 2.7263 | −2.5907 | 4.0161 | 0.0000 | 6.3146 |
| Trimethoxysilane | 0.1656 | 2.5734 | 0.1880 | 0.1650 | −1.5622 | 5.8791 | 0.6841 | 8.0931 |
| Vinyltrimethoxysilane | 0.3527 | 2.8857 | 0.0654 | 0.0073 | −3.3706 | 5.9266 | 1.8743 | 7.7415 |
The total energy measurements for dimethyldichlorosilane and methyltrichlorosilane reveal significant negative values at −1.56 kcal/mol and −0.83 kcal/mol, respectively. The compounds show large negative torsional energies, implying that the internal strain and poor dihedral angles determine their stability. Different silane derivatives display unique patterns when their energy distributions are analyzed. The high total energy values of cyclohexyltrimethoxysilane at 13.72 kcal/mol and methacryloxypropyltrimethoxysilane at 18.44 kcal/mol result from torsion and 1,4 VDW interactions. The energy profiles of cyclic and functionalized silane derivatives show higher steric and electronic demands than those of linear compounds, which creates more complex energy landscapes. Detailed examination of the separate energy components offers valuable structural information. Hexamethyldisiloxane exhibits a significantly negative energy level for non-1,4 VDW interactions, which amounts to −4.62 kcal/mol. Polydimethylsiloxane demonstrates large negative values for both non-1,4 VDW and 1,4 VDW interactions which suggests strong repulsive forces within its polymeric framework. The extensive flexibility and chain-like structure of polydimethylsiloxane results in distinct energetic properties. Phenyltrimethoxysilane and Triethoxysilane demonstrate superior molecular stability through their total energies of 6.37 kcal/mol and 10.96 kcal/mol when compared to other molecules. These molecules maintain stable structures through balanced forces, despite their significant torsional and VDW interaction contributions.
Results of molecular dynamics (MD) simulations
Results of material characterization and mechanical properties
Significant variations in the elastic modulus and tensile strength of the studied silicone materials emerged from the results. Applications that require both flexibility and strength depend on these mechanical properties, and polydimethylsiloxane serves as the standard because of its proven performance. PDMS demonstrates an experimental elastic modulus of 1.32–2.97 MPa and tensile strength between 3.51 and 5.13 MPa [13, 37], which makes it suitable as a comparative standard for other silicone materials. The polydimethylsiloxane displayed an elastic modulus of 2.533 GPa within the validated range of 1.32–2.97 GPa according in Fig. 1a. PDMS is an adaptable dental material that can meet the demands of applications that require both flexibility and rigidity. The high stiffness of this material makes it perfect for dental impression materials because it maintains accuracy in detailed reproduction without noticeable deformation when pressure is applied. The inherent elasticity of PDMS allows clinicians to remove dental impressions without disturbing captured details. The high elastic moduli of methacryloxypropyltrimethoxysilane (3.248 GPa) and hexaphenylcyclotrisiloxane (3.017 GPa) make them suitable for dental applications that require high stiffness capacity. Composite resins can benefit from these materials as reinforcing agents, and dental adhesives require them to maintain structural stability under occlusal forces. Orthodontic applications that require brackets or splinting materials can benefit from the high mechanical deformation resistance provided by these stiff materials. Both butyltrichlorosilane at 0.827 GPa and diethylsilane at 0.912 GPa exhibited reduced elastic moduli, demonstrating higher flexibility. The material characteristics enable their use in situations where adaptability to diverse shapes and dynamic forces is necessary, such as in temporary prosthetic devices and soft dental liners. Patients should experience improved comfort while performing oral movements owing to material flexibility. Impression materials achieve optimal performance through an elastic modulus range of 1.5–3 GPa, which balances the accuracy with durability. Cyclohexyltrimethoxysilane (2.948 GPa) and phenyltrimethoxysilane (2.762 GPa) displayed elastic moduli within this range, indicating their potential for producing both accurate and durable dental impressions. The materials balance adequate stiffness to resist distortion with sufficient elasticity to detach from the undercuts and complex dental structures.
Fig. 1.
Comparative analysis of (a) elastic modulus (GPa) and (b) tensile strength (MPa) of various silicone-based materials. Each value represents the mean of three independent measures, with error bars showing the corresponding standard deviations
Tensile strength measurements of the silicone-based materials tested showed significant variations between 3.002 MPa for chlorotrimethylsilane and 5.847 MPa for methacryloxypropyltrimethoxysilane, as shown in Fig. 1b. The tensile strength variations among the materials demonstrate their distinct abilities to withstand deformation and prevent failure when exposed to tensile forces. PDMS displayed a tensile strength of 5.387 MPa during simulation while remaining within the experimental range of 3.51–5.13 MPa for PDMS which shows that its mechanical performance meets predicted standards. PDMS shows high tensile strength, which enables it to balance flexibility with resistance to tension, thus making it ideal for applications that demand durability and mechanical failure resistance. Tensile strength measurements revealed that methacryloxypropyltrimethoxysilane (5.847 MPa) and hexaphenylcyclotrisiloxane (5.292 MPa) exceeded PDMS’s strength of PDMS and showed promise for applications that require superior tensile resistance. Such materials are highly valuable in applications that require high mechanical resistance, including load-bearing biomedical devices and structural industrial components. Materials such as dimethyldichlorosilane and methyltrichlorosilane demonstrated tensile strengths of 3.301 and 3.108 MPa, respectively, which were lower than those of PDMS. These materials retain their usefulness for basic applications, but their limited tensile resistance imposes performance restrictions in high-stress situations. Materials that can withstand mechanical stress, such as PDMS, methacryloxypropyltrimethoxysilane, and hexaphenylcyclotrisiloxane, are essential choices for dental restoratives and implants because they offer the required durability against mechanical breakdown. Applications that require flexibility and conformability should use materials with low tensile strength, instead of those with high stress resistance.
Results of polymerization dynamics
Simulations were used to assess the polymerization dynamics across different materials to gain insight into their energetic profiles and reaction kinetics. The obtained data are essential for dental use because accurate polymerization control directly affects material durability and performance. The curing time and mechanical properties of dental materials, such as resins and composites, depend on the polymerization rate (mol/s) and polymerization energy barrier (kJ/mol). The material with the fastest polymerization rate among those tested was methacryloxypropyltrimethoxysilane at 0.078 mol/s, making it ideal for dental sealants and adhesive systems that benefit from short setting times. Hexaphenylcyclotrisiloxane demonstrates rapid polymerization at 0.070 mol/s, which could help in dental applications that require quick curing times to speed up the procedure. Methyltrichlorosilane showed the slowest polymerization rate at 0.012 mol/s, indicating a slower polymerization process that serves well in applications needing precise control over working times like dental impressions or restorations (Fig. 2a).
Fig. 2.
Comparative analysis of the (a) polymerization rate (mol/s) and (b) polymerization energy barrier (kJ/mol) of various silicone-based materials. Each value represents the mean of three independent measures, with error bars showing the corresponding standard deviations
Materials with high polymerization energy barriers are more resistant to premature solidification, which makes them beneficial in some dental practices. The energy barrier for methacryloxypropyltrimethoxysilane is the highest at 77.923 kJ/mol, indicating that the material maintains stability before polymerization, making it beneficial for dental adhesives or restorative materials that require stability until they are triggered by light or heat. Hexaphenylcyclotrisiloxane and cyclohexyltrimethoxysilane displayed energy barriers of 74.516 kJ/mol and 70.251 kJ/mol, respectively, which supports their use in dental materials that require strong resistance to premature polymerization and can successfully polymerize under regulated circumstances. Materials including methyltrichlorosilane (54.680 kJ/mol) and chlorotrimethylsilane (55.945 kJ/mol) exhibit lower energy barriers, which contribute to easier polymerization. Materials with fast curing properties are ideal for dental applications, such as temporary fillings or restorations, because they require rapid setting to reduce patient discomfort while speeding up treatment procedures. With a polymerization rate of 0.051 mol/s and polymerization energy barrier of 68.123 kJ/mol, polydimethylsiloxane achieved a balanced polymerization profile. The middle-range polymerization rate and energy barrier of this material make it ideal for dental applications that require controlled setting times to avoid both quick polymerization and overly slow curing (Fig. 2b). Dental restorations require materials that balance the working time with the end properties, making polydimethylsiloxane a suitable choice based on its polymerization characteristics. The material exhibits flexibility and resilience as a mechanical property, which makes it appropriate for dental applications such as sealants, adhesives, and implants because these applications demand both strength and compatibility with the tooth structure.
Results of dimensional stability
Silicone-based materials maintain dimensional stability, making them ideal for use in settings where environmental conditions such as temperature shifts and humidity variations exist. Researchers have measured the predicted volume changes and contact angles along with dimensional changes in various materials to evaluate their potential for stable applications such as dental implants and prosthetics. Cyclohexyltrimethoxysilane, Hexaphenylcyclotrisiloxane, Methacryloxypropyltrimethoxysilane, Methyldiphenylsilane, and Triisopropylsilane demonstrated great dimensional stability, with predicted volume changes between 0.021 and 0.026 mg/mm3 and contact angles between 110° and 115°, as shown in Table 3. The materials demonstrated minimal dimensional alterations when exposed to environmental stresses, confirming their high stability. The materials’ minimal water absorption and strong hydrophobic nature, indicated by elevated contact angles, lead to excellent resistance to expansion caused by moisture. These materials perform optimally in wet conditions, such as in the oral cavity, owing to their ability to handle constant moisture exposure. These materials maintain stable structural integrity when exposed to environmental changes, which demonstrates their suitability for applications that require durable and resilient performance over extended periods. Polydimethylsiloxane achieved moderate stability classification through its predicted volume change of 0.027 mg/mm3 and contact angle of 104°. This material exhibited minimal dimensional change, demonstrating its ability to maintain its structural form when subjected to different environmental conditions. Polydimethylsiloxane demonstrates moderate dimensional stability because it maintains equilibrium between its hydrophobic nature and moderate tensile strength. Although not as stable as materials exhibiting the highest hydrophobic traits, polydimethylsiloxane serves as a dependable material for situations that permit minimal dimensional changes. The medical and dental industries commonly use polydimethylsiloxane because of its flexible nature and compatibility with biological systems, while its moderate dimensional stability benefits these applications. This material maintains its shape in oral settings despite moisture exposure but lacks the superior moisture resistance of highly hydrophobic substances.
Table 3.
Predicted dimensional stability and hydrophobicity of silicone materials
| Silicone | Volume change (mg/mm3) | Contact angle (°) | Dimensional change |
|---|---|---|---|
| Aminopropyltriethoxysilane | 0.028 | 103 | Low change (moderate stability) |
| Butyltrichlorosilane | 0.034 | 91 | Moderate change (poor stability) |
| Chlorotrimethylsilane | 0.038 | 85 | High change (poor stability) |
| Cyclohexyltrimethoxysilane | 0.024 | 111 | Low change (high stability) |
| Decamethylcyclopentasiloxane | 0.030 | 96 | Low change (moderate stability) |
| Diethylsilane | 0.032 | 89 | Moderate change (poor stability) |
| Dimethyldichlorosilane | 0.036 | 84 | High change (poor stability) |
| Dodecamethylcyclohexasiloxane | 0.025 | 108 | Low change (moderate stability) |
| Hexamethyldisiloxane | 0.031 | 97 | Low change (moderate stability) |
| Hexaphenylcyclotrisiloxane | 0.022 | 113 | Low change (high stability) |
| Methacryloxypropyltrimethoxysilane | 0.021 | 115 | Very low change (high stability) |
| Methyldiphenylsilane | 0.023 | 110 | Low change (high stability) |
| Methyltrichlorosilane | 0.037 | 83 | High change (poor stability) |
| Octamethylcyclotetrasiloxane | 0.028 | 101 | Low change (moderate stability) |
| Octamethyltrisiloxane | 0.032 | 95 | Moderate change (poor stability) |
| Phenyldimethylsilane | 0.030 | 98 | Moderate change (poor stability) |
| Phenyltrimethoxysilane | 0.026 | 106 | Low change (high stability) |
| Polydimethylsiloxane | 0.027 | 104 | Low change (moderate stability) |
| Tetraethoxysilane | 0.029 | 99 | Low change (moderate stability) |
| Tetravinyltetramethylcyclotetrasiloxane | 0.031 | 93 | Moderate change (poor stability) |
| Triethoxysilane | 0.028 | 101 | Low change (moderate stability) |
| Trifluoropropyltrimethoxysilane | 0.027 | 105 | Low change (moderate stability) |
| Triisopropylsilane | 0.026 | 107 | Low change (high stability) |
| Trimethoxysilane | 0.029 | 98 | Moderate change (poor stability) |
| Vinyltrimethoxysilane | 0.028 | 102 | Low change (moderate stability) |
Several materials demonstrate substantial dimensional changes, indicating that they lack stability when exposed to environmental stress. Chlorotrimethylsilane, dimethyldichlorosilane, methyltrichlorosilane, and trimethoxysilane showed substantial volume changes from 0.036 to 0.038 mg/mm3 and maintained contact angles between 83° and 85°. These materials exhibit increased moisture absorption, which results in dimensional expansion and decreased stability when exposed to humidity. A lower contact angle indicates that these materials have a higher hydrophilicity, which leads to greater moisture absorption, resulting in swelling and subsequent mechanical degradation. These materials are ill-suited for use in the oral cavity because their exposure to water leads to material failure or degradation over time. The mechanical strength profiles of materials determine the variations in dimensional stability. Materials with higher tensile strength and elastic modulus: The cyclohexyltrimethoxysilane (elastic modulus: cyclohexyltrimethoxysilane and methacryloxypropyltrimethoxysilane, maintained their shape under stress because of their respective mechanical strength parameters: cyclohexyltrimethoxysilane with an elastic modulus of 2.948 GPa and tensile strength of 5.002 MPa and methacryloxypropyltrimethoxysilane with an elastic modulus of 3.248 GPa and tensile strength of 5.847 MPa. These materials demonstrate superior performance by resisting dimensional changes under mechanical stress and environmental exposure. Conversely, materials with lower tensile strength, such as chlorotrimethylsilane (elastic modulus: chlorotrimethylsilane (elastic modulus: 0.564 GPa, tensile strength: 3.002 MPa) showed greater dimensional changes that resulted from increased vulnerability to moisture expansion and mechanical breakdown.
Results of degradation
This research tested several silanes and siloxanes, which showed varied degradation behaviors essential for evaluating their long-term usage potential. Hexaphenylcyclotrisiloxane showed the smallest degradation rate (0.021 mg/mm3) among the materials tested, indicating its superior stability under simulated oral conditions. The molecular composition of this material, featuring highly substituted silicon atoms, likely provides enhanced durability against degradation, which makes it suitable for high-durability applications. Both methacryloxypropyltrimethoxysilane and methyldiphenylsilane showed degradation rates of 0.02 mg/mm3 and 0.022 mg/mm3 correspondingly while also demonstrating significant stability levels. The multiple functional groups present in the structures of these materials help protect against degradation in oral environments, thus improving their suitability for long-term use. Chlorotrimethylsilane showed a degradation rate of 0.037 mg/mm3, whereas methyltrichlorosilane demonstrated the highest degradation rate at 0.036 mg/mm3. The presence of reactive functional groups such as chlorine makes these materials highly vulnerable to both hydrolytic and oxidative degradation in complex oral environments. Based on simulated condition tests, these materials showed high reactivity, indicating their potential unsuitability for applications requiring long-term stability. The degradation rates of dimethyldichlorosilane and butyltrichlorosilane at 0.035 mg/mm3 and 0.033 mg/mm3 were relatively higher, thus confirming the pattern that chlorine-containing silanes degrade faster under these conditions, as shown in Fig. 3.
Fig. 3.
Predicted degradation rates (mg/mm3) of various silicone-based materials. Each value represents the mean of three independent measures, with error bars showing the corresponding standard deviations
The degradation rate of polydimethylsiloxane was 0.027 mg/mm3 for silicone-based applications. The material showed moderate degradation rates when tested alongside other materials, yet it maintained good general stability. The molecular framework of polydimethylsiloxane consists of siloxane (Si-O) repeating units that demonstrate chemical resilience to moisture and other external stress factors. The material demonstrates excellent performance because its degradation rate proves its suitability for medical and industrial applications that require moderate degradation over long durations. Cyclohexyltrimethoxysilane and decamethylcyclopentasiloxane demonstrated degradation rates of 0.025 and 0.031 mg/mm3, respectively, indicating their strong stability when exposed to oral fluids. The degradation rates of these materials show that their chemical structure with bulky side groups helps them resist degradation, although they remain slightly more prone to degradation than the most stable candidates, such as hexaphenylcyclotrisiloxane. Tetraethoxysilane, trimethoxysilane and vinyltrimethoxysilane displayed matching degradation rates of 0.029 mg/mm3, indicating their moderate resistance to degradation. Functional groups in these materials enable hydrolysis reactions, leading to increased degradation rates, yet they maintain adequate stability for use that permits minimal degradation.
Results of in silico toxicity assessment of silicones
The results of this study offer early insights into the safety profile of dental materials based on their effects on oral tissues and overall patient safety (Table 4). Measurement of lipophilicity through cLogP values plays an essential role in determining the penetration of tissue membranes by these materials. The elevated cLogP values of polydimethylsiloxane (11.1470) and dodecamethylcyclohexasiloxane (8.3604) indicate that these dental impression materials have a greater tendency to accumulate in fatty tissues, and thus may increase the risk of systemic exposure if they migrate out of the material over time. The low cLogP values displayed by trimethoxysilane (−0.8427) and triethoxysilane (0.3762) indicate that these materials have higher hydrophilicity, which may result in reduced tissue accumulation and improved body excretion. The TPSA values that demonstrate the molecular polarity and biological membrane permeability abilities [33] exhibited substantial differences between the materials. The high TPSA value (64.61 of polydimethylsiloxane suggests that this compound will encounter challenges when penetrating biological membranes, which could lead to reduced bioavailability when it comes into contact with oral tissues. The low TPSA values observed for butyltrichlorosilane (0.00) and dimethyldichlorosilane (0.00) suggested that these compounds can permeate cellular membranes more easily, which could present risks if they enter the oral cavity over time. Scientists have evaluated the molecular shape index and molecular flexibility to determine possible interactions between silicone materials and biological systems. The significant flexibility values of triethoxysilane (0.8693) and cyclohexyltrimethoxysilane (0.6860) demonstrated their capability to form multiple molecular shapes, which enhanced their binding potential with biomolecules. These interactions can lead to detrimental effects on oral tissues, such as irritation or allergy development. The flexibility value of chlorotrimethylsilane was 0.0000, which indicates minimal flexibility and suggests that it interacts less dynamically with biological structures, probably leading to a lower risk of adverse interactions.
Table 4.
Toxicity parameters for silicone used in dental impression materials
| Material | Toxicity parameter | |||||||
|---|---|---|---|---|---|---|---|---|
| cLogP | TPSA | Molecular shape index | Molecular flexibility | Mutagenic | Tumorigenic | Reproductive effective | Irritant | |
| Aminopropyltriethoxysilane | 0.4544 | 53.71 | 0.5712 | 0.8585 | None | High | None | High |
| Butyltrichlorosilane | 1.8571 | 0.00 | 0.7500 | 0.6727 | None | None | None | High |
| Chlorotrimethylsilane | 2.0901 | 0.00 | 0.6000 | 0.0000 | High | High | None | High |
| Cyclohexyltrimethoxysilane | 0.8673 | 27.69 | 0.5384 | 0.6860 | None | High | None | High |
| Decamethylcyclopentasiloxane | 6.9670 | 46.15 | 0.3500 | 0.5284 | None | None | High | High |
| Diethylsilane | 1.8966 | 0.00 | 1.0000 | 0.7853 | None | None | None | High |
| Dimethyldichlorosilane | 1.3934 | 0.00 | 0.6000 | 0.0000 | None | None | None | High |
| Dodecamethylcyclohexasiloxane | 8.3604 | 55.38 | 0.3750 | 0.5588 | None | None | High | High |
| Hexamethyldisiloxane | 4.1802 | 9.23 | 0.5555 | 0.8995 | None | None | Low | High |
| Hexaphenylcyclotrisiloxane | 8.2980 | 27.69 | 0.2619 | 0.2909 | None | None | High | High |
| Methacryloxypropyltrimethoxysilane | 0.9932 | 53.99 | 0.6250 | 0.7323 | High | High | High | High |
| Methyldiphenylsilane | 3.4627 | 0.00 | 0.6428 | 0.4116 | None | None | Low | High |
| Methyltrichlorosilane | 0.6967 | 0.00 | 0.6000 | 0.0000 | None | None | None | High |
| Octamethylcyclotetrasiloxane | 5.5736 | 36.92 | 0.4375 | 0.4825 | None | None | High | High |
| Octamethyltrisiloxane | 5.5736 | 18.46 | 0.5384 | 0.9276 | None | None | High | High |
| Phenyldimethylsilane | 2.7764 | 0.00 | 0.6666 | 0.3500 | None | None | Low | High |
| Phenyltrimethoxysilane | 0.5403 | 27.69 | 0.5384 | 0.6767 | None | High | None | High |
| Polydimethylsiloxane | 11.1470 | 64.61 | 0.5151 | 0.9612 | None | None | High | High |
| Tetraethoxysilane | 0.5016 | 36.92 | 0.5384 | 0.8754 | None | None | None | High |
| Tetravinyltetramethylcyclotetrasiloxane | 4.5036 | 36.92 | 0.4500 | 0.6582 | None | None | High | High |
| Triethoxysilane | 0.3762 | 27.69 | 0.7000 | 0.8693 | None | None | None | High |
| Trifluoropropyltrimethoxysilane | 1.0443 | 27.69 | 0.5384 | 0.8209 | None | High | None | High |
| Triisopropylsilane | 2.9316 | 0.00 | 0.5000 | 0.7586 | None | None | None | High |
| Trimethoxysilane | −0.8427 | 27.69 | 0.7142 | 0.8355 | None | Low | None | High |
| Vinyltrimethoxysilane | −0.4135 | 27.69 | 0.5555 | 0.8962 | None | High | None | High |
Different compounds display a range of toxicological potential. Several silicones have been identified as strong irritants that can cause irritation when they come into contact with the oral mucosa. The direct contact between dental impression materials and soft oral tissues marks irritation from these materials as a major issue [38], The mutagenic and tumorigenic properties of methacryloxypropyltrimethoxysilane and chlorotrimethylsilane suggestthat these chemicals could lead to genetic damage and cancer, which becomes worrisome in dental useowinge to possible long-term patient exposure. Research has identified polydimethylsiloxane, octamethylcyclotetrasiloxane, and dodecamethylcyclohexasiloxane as highly reproductive toxins that may affect fertility and pregnancy outcomes. As dental impressions are a standard procedure for pregnant patients, their reproductive toxicity should be examined further to confirm their safety for use. Research indicates that methacryloxypropyltrimethoxysilane and chlorotrimethylsilane possess tumor-causing properties that could heighten cancer risks, thereby necessitating careful use of these compounds in dental impression substances because they often remain in contact with oral tissues over long durations.
Discussion
The 3D structural modeling of various silicone compounds via MM2 energy minimization demonstrates significant energetic profile and stability differences, which are essential for evaluating material suitability across various applications, particularly in dental applications. The study’s total energy measurements show a spectrum from deeply negative for substances such as decamethylcyclopentasiloxane (−10.28 kcal/mol) and dodecamethylcyclohexasiloxane (−17.01 kcal/mol) to significantly positive for compounds like aminopropyltriethoxysilane (8.62 kcal/mol). These compounds maintain their stability through a balance of torsional forces alongside VDW forces and dipole–dipole interactions found in specific silane derivatives. Research findings confirm that molecular dynamics and force field-based simulations are critical methods for predicting the behavior of polymeric materials under various conditions [21, 39, 40]. Aminopropyltriethoxysilane demonstrated high dipole–dipole interactions, which led to a rigid structure, whereas butyltrichlorosilane showed lower dipole interactions, resulting in higher flexibility. The energetic insights derived from these studies direct the design process of silicone-based materials that possess specific properties for dental applications [41].
Molecular dynamics simulations generate important mechanical data on the elastic modulus and tensile strength of different silicone materials. The mechanical properties of silicone materials are essential when they are applied in dental prosthetics, impressions, and adhesives. The combination of an elastic modulus of 2.533 GPa and tensile strength of 5.387 MPa allows polydimethylsiloxane to show superior flexibility and rigidity, which qualifies it as the best choice for dental impression materials [13, 42]. Methacryloxypropyltrimethoxysilane (3.248 GPa) and hexaphenylcyclotrisiloxane (3.017 GPa) exhibited enhanced stiffness, suggesting their usefulness in load-bearing scenarios. Butyltrichlorosilane and diethylsilane possess low elastic moduli, which make them more appropriate for flexible use. These research findings agree with earlier studies demonstrating that adjusting the mechanical properties of silicone-based materials is critical to match specific application needs, especially in dental practice, which requires materials to display both strong and flexible characteristics in various situations [43, 44].
The study of polymerization dynamics and dimensional stability demonstrates how silicone-based materials can be used in various dental and medical applications. Methacryloxypropyltrimethoxysilane is suitable for applications requiring swift curing and stable performance because of its high polymerization rates and energy barriers, which make it ideal for dental adhesives and sealants [45]. Materials such as methyltrichlorosilane possess lower energy barriers, which makes them appropriate for applications requiring controlled working times. Dimensional stability is a fundamental property of dental materials that undergo temperature changes and moisture exposure. Hexaphenylcyclotrisiloxane and cyclohexyltrimethoxysilane maintain their dimensions well and, therefore, serve as suitable materials for extended use. polydimethylsiloxane provides both flexibility and moderate stability, making it an appropriate choice for dynamic oral environments. These results emphasize the necessity for specific silicone materials that match the exact requirements of medical and dental applications because stability and polymerization traits affect the performance of the final product [46, 47].
The in silico findings reported here can be explained by the intrinsic molecular structures of the silicone compounds. For example, compounds with highly cross-linked or aromatic substituents (e.g., hexaphenylcyclotrisiloxane) display greater stiffness and dimensional stability due to restricted rotational freedom, whereas linear or less substituted molecules (e.g., diethylsilane) demonstrate higher flexibility. Similarly, the polymerization trends observed in silico correlate with substituent effects on reactivity, with methacryloxy groups accelerating polymerization by stabilizing transition states. To validate these predictions, several in vitro tests may be performed. Mechanical findings (elastic modulus, tensile strength) can be compared with results from universal testing machine experiments. Dimensional stability can be assessed using ISO 4823 protocols for dental elastomeric impression materials [48]. Polymerization rates and curing behavior may be examined via differential scanning calorimetry and Fourier-transform infrared spectroscopy [49]. Finally, degradation and toxicity predictions from the simulations can be confirmed by in vitro cytotoxicity assays such as MTT or Live/Dead tests using human gingival fibroblasts [50]. These complementary in vitro methods would provide robust evidence to either support or reject the computational predictions, thereby strengthening translational relevance.
Limitations and clinical considerations
This study yields important findings about the structure and mechanical properties of silicone-based materials but requires consideration of its limitations. Theoretical data from molecular dynamics simulations and energy minimization studies cannot completely represent the true behavior of compounds in the oral environment. When materials interact with oral tissues, certain biological interactions occur, and in silico methods fail to consider which can affect material performance over time. Simulation calculations provided mechanical property data for these materials, including the elastic modulus and tensile strength, but clinical validation remains crucial to establish their applicability in dental procedures. This study did not examine the effects of long-term aging and degradation due to thermal cycling, moisture absorption, or mechanical stress, which occur in the oral environment. The evaluation of material behavior under prolonged exposure conditions requires clinical testing and in vivo studies. The study’s restriction to a finite number of silicone compounds creates a scope for better material selection through research on additional materials in clinical applications. The examination of promising materials failed to address potential oral cavity reactions such as toxicity and inflammation, which are essential for patient safety and biocompatibility assessments.
Conclusion and future works
This study presents the mechanical features and polymerization capabilities of silicone-based materials, which show their potential for multiple dental applications. The mechanical testing results indicate that polydimethylsiloxane, Methacryloxypropyltrimethoxysilane, and Hexaphenylcyclotrisiloxane meet all the properties required for use in applications requiring both flexibility and strength while maintaining stability. Through energy minimization and molecular dynamics simulations, scientists have obtained improved insights into the energetic profiles and mechanical behaviors of these materials, which will aid future material design and optimization efforts. These limitations highlight the need for further experimental validation and clinical trials to establish the performance of these materials in practical applications. The next phase of research should broaden molecular simulation studies to cover more materials and conditions, while incorporating experimental validation to demonstrate the clinical relevance of the results. Performing long-term in vivo studies together with investigating the toxicity profiles of the materials will result in an enhanced understanding of their appropriateness for oral use. The creation of new composite materials with silicone could lead to improved mechanical properties and extended service life for dental restorations, while delivering better and more resilient solutions for patients.
Supplementary information
Acknowledgements
All the authors thank King Khalid University, Saudi Arabia, for the financial Support.
Author contributions
Conceptualization and Methodology: Ravinder Saini, Doni Dermawan; Data Curation and Formal Analysis: Doni Dermawan, Lujain Aldosari, Rajesh Vyas; Validation and Formal Analysis: Artak Heboyan, Mario Alberto Alarcón-Sanchez; Investigation and Resources: Doni Dermawan, Rayan Binduhayyim, Abdullah Hassan; Original draft preparation: Ravinder Saini, Masroor Kanji, Mario Alberto Alarcón-Sánchez; Writing, Reviewing, and Editing: Doni Dermawan, Abdulkhaliq Alshadidi, Lujain Ibrahim, Artak Heboyan; Supervision and Project Administration: Ravinder Saini, Artak Heboyan.
Funding
The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work through Large Research Project under grant number RGP2/545/46.
Data availability
The data are available to the corresponding author upon request.
Compliance with ethical standards
Conflict of interest
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Mario Alberto Alarcón-Sánchez, Email: marioaasanchez@hotmail.com.
Artak Heboyan, Email: a_heboyan@farabi.tums.ac.ir.
Supplementary information
The online version contains supplementary material available at 10.1007/s10856-025-06944-w.
References
- 1.Chistyakov EM, Kolpinskaya N, Posokhova V, Chuev V. Dental composition modified with aryloxyphosphazene containing carboxyl groups. Polymers. 2020;12:1176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Kireev VV, Chistyakov EM, Filatov SN, Tupikov AS, Panfilova DV, Chetverikova AI. Polymeric dental composites modified with carboxy phosphazene methacrylates. Russ J Appl Chem. 2015;88:866–70. [Google Scholar]
- 3.Madanshetty P, Guttal SS, Meshramkar R, Newaskar PS, Anehosur GV. Addition silicone impressions in fixed prosthodontics: clinical standpoints. Cureus. 2023;15:e44014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Naumovski B, Kapushevska B. Dimensional stability and acuracy of silicone—based impression materials using different impression techniques—a literature review. Pril (Makedon Akad Nauk Umet Odd Med Nauki). 2017;38:131–8. [DOI] [PubMed] [Google Scholar]
- 5.Gradinariu AI, Racles C, Stoica I, Stelea CG, Simionescu AA, Jehac AE, et al. Silicones for maxillofacial prostheses and their modifications in service. Materials. 2024;17:3297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Cramer NB, Stansbury JW, Bowman CN. Recent advances and developments in composite dental restorative materials. J Dent Res. 2011;90:402–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Antonucci JM, Dickens SH, Fowler BO, Xu HH, McDonough WG. Chemistry of silanes: interfaces in dental polymers and composites. J Res Natl Inst Stand Technol. 2005;110:541–58. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Fu S-Y, Feng XQ, Lauke B, Mai YW. Effects of particle size, particle/matrix interface adhesion and particle loading on mechanical properties of particulate–polymer composites. Compos Part B: Eng. 2008;39:933–61. [Google Scholar]
- 9.Gomes de Araújo-Neto V, Sebold M, Fernandes de Castro E, Feitosa VP, Giannini M. Evaluation of physico-mechanical properties and filler particles characterization of conventional, bulk-fill, and bioactive resin-based composites. J Mech Behav Biomed Mater. 2021;115:104288. [DOI] [PubMed] [Google Scholar]
- 10.Fakoori M, Hesaraki S, Nezafati N, Ghiass M. Evaluation of condensation silicone dental impression materials modified with diatomaceous Earth and zinc oxide fillers. Sci Rep. 2025;15:25289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Rohanian A, Ommati Shabestari G, Zeighami S, Samadi MJ, Shamshiri AR. Effect of storage time of extended-pour and conventional alginate impressions on dimensional accuracy of casts. J Dent. 2014;11:655–64. [PMC free article] [PubMed] [Google Scholar]
- 12.Walker MP, Burckhard J, Mitts DA, Williams KB. Dimensional change over time of extended-storage alginate impression materials. Angle Orthod. 2010;80:1110–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Ariati R, Sales F, Souza A, Lima RA, Ribeiro J. Polydimethylsiloxane composites characterization and its applications: a review. Polymers. 2021;13:4258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Miranda I, Souza A, Sousa P, Ribeiro J, Castanheira E, Lima R, et al. Properties and applications of pdms for biomedical engineering: a review. J Funct Biomater. 2021;13:2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Stark FO, Falender JR, Wright AP. Silicones. In: Wilkinson G, Stone FGA, Abel EW, editors. Comprehensive organometallic chemistry. Oxford: Pergamon; 1982. p. 305–63.
- 16.Zakir M, Ashraf U, Tian T, Han A, Qiao W, Jin X, et al. The role of silane coupling agents and universal primers in durable adhesion to dental restorative materials—a review. Curr Oral Health Rep. 2016;3:244–53. [Google Scholar]
- 17.Shahi S, Özcan M, Maleki Dizaj S, Sharifi S, Al-Haj Husain N, Eftekhari A, et al. A review on potential toxicity of dental material and screening their biocompatibility. Toxicol Mech Methods. 2019;29:1–24. [DOI] [PubMed] [Google Scholar]
- 18.Mojsiewicz-Pieńkowska K, Jamrógiewicz M, Szymkowska K, Krenczkowska D. Direct human contact with siloxanes (silicones) – safety or risk part 1. Characteristics of siloxanes (silicones). Front Pharm. 2016;7:132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Saini RS, Binduhayyim R, Gurumurthy V, Alshadidi A, Aldosari L, Okshah A, et al. Dental biomaterials redefined: molecular docking and dynamics-driven dental resin composite optimization. BMC Oral Health. 2024;24:557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Saini RS, Binduhayyim R, Gurumurthy V, Alshadidi A, Bavabeedu SS, Vyas R, et al. In silico assessment of biocompatibility and toxicity: molecular docking and dynamics simulation of PMMA-based dental materials for interim prosthetic restorations. J Mater Sci Mater Med. 2024;35:28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Saini RS, Vaddamanu SK, Dermawan D, Mosaddad SA, Heboyan A. Investigating the role of temperature and moisture on the degradation of 3D-printed polymethyl methacrylate dental materials through molecular dynamics simulations. Sci Rep. 2024;14:26079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Gartner TE III, Jayaraman A. Modeling and simulations of polymers: a roadmap. Macromolecules. 2019;52:755–86. [Google Scholar]
- 23.Anstine D, Strachan A, Colina C. Effects of an atomistic modeling approach on predicted mechanical properties of glassy polymers via molecular dynamics. Model Simul Mater Sci Eng. 2020;28:025006. [Google Scholar]
- 24.Mallakpour S, Barati A. Efficient preparation of hybrid nanocomposite coatings based on poly(vinyl alcohol) and silane coupling agent modified TiO2 nanoparticles. Prog Org Coat. 2011;71:391–8. [Google Scholar]
- 25.Horii Y. Survey of organosilicone compounds, including cyclic and linear siloxanes, in personal-care and household products. Arch Environ Contam Toxicol. 2008;55:701–10. [DOI] [PubMed] [Google Scholar]
- 26.Dermawan D, Prabowo BA, Rakhmadina CA. In silico study of medicinal plants with cyclodextrin inclusion complex as the potential inhibitors against SARS-CoV-2 main protease (Mpro) and spike (S) receptor. Inf Med Unlocked. 2021;25:1–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Fernández B, Ríos MA, Carballeira L. Molecular mechanics (MM2) and conformational analysis of compounds with N–C–O units. Parametrization of the force field and anomeric effect. J Comput Chem. 1991;12:78–90. [Google Scholar]
- 28.Vanommeslaeghe K, Guvench O, MacKerell AD Jr. Molecular mechanics. Curr Pharm Des. 2014;20:3281–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Meunier M, Robertson S. Materials Studio 20th anniversary. Mol Simul. 2021;47:537–9. [Google Scholar]
- 30.Chessells M, Hawker DW, Connell DW. Critical evaluation of the measurement of the 1- octanol/water partition coefficient of hydrophobic compounds. Chemosphere. 1991;22:1175–90. [Google Scholar]
- 31.Alotaiq N, Dermawan D. Advancements in virtual bioequivalence: a systematic review of computational methods and regulatory perspectives in the pharmaceutical industry. Pharmaceutics. 2024;16:1414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Saini RS, Dermawan D, Madhav VNV, Aldosari LIN, Binduhayyim RIH, Vyas R, et al. Computational analysis of protein-like and synthetic polymers in denture base resins: a molecular docking and dynamics study for improving biocompatibility. Polymer. 2025;336:128897. [Google Scholar]
- 33.Prasanna S, Doerksen RJ. Topological polar surface area: a useful descriptor in 2D-QSAR. Curr Med Chem. 2009;16:21–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Kortagere S, Krasowski MD, Ekins S. The importance of discerning shape in molecular pharmacology. Trends Pharm Sci. 2009;30:138–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Braun J, Katzberger P, Landrum GA, Riniker S. Understanding and quantifying molecular flexibility: torsion angular bin strings. J Chem Inf Model. 2024;64:7917–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Sander T, Freyss J, von Korff M, Rufener C. DataWarrior: an open-source program for chemistry aware data visualization and analysis. J Chem Inf Model. 2015;55:460–73. [DOI] [PubMed] [Google Scholar]
- 37.Zhang G, Sun Y, Qian B, Gao H, Zuo D. Experimental study on mechanical performance of polydimethylsiloxane (PDMS) at various temperatures. Polym Test. 2020;90:106670. [Google Scholar]
- 38.Hamalian TA, Nasr E, Chidiac JJ. Impression materials in fixed prosthodontics: influence of choice on clinical procedure. J Prosthodont. 2011;20:153–60. [DOI] [PubMed] [Google Scholar]
- 39.Araújo MC, Martins JP, Mirkhalaf SM, Lanceros-Mendez S, Pires FMA, Simoes R. Predicting the mechanical behavior of amorphous polymeric materials under strain through multi-scale simulation. Appl Surf Sci. 2014;306:37–46. [Google Scholar]
- 40.Garduño-Juárez R, Tovar-Anaya DO, Perez-Aguilar JM, Lozano-Aguirre Beltran LF, Zubillaga RA, Alvarez-Perez MA, et al. Molecular dynamic simulations for biopolymers with biomedical applications. Polymers. 2024;16:1864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Mutahar M, Al Ahmari NM, Alqahtani SM, Gadah TS, Alshehri AH, Adawi HA, et al. Comparative analysis of properties in chairside silicone denture liners versus heat-cured molloplast-b: an in-depth in vitro evaluation. Med Sci Monit. 2023;29:941793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Han C-S, Sanei SHR, Alisafaei F. On the origin of indentation size effect and depth dependent mechanical properties of elastic polymers. J Polym Eng. 2015; 36,103–111.
- 43.Lührs A-K, Geurtsen W. The application of silicon and silicates in dentistry: a review. Prog Mol Subcell Biol. 2009;47:359–80. [DOI] [PubMed] [Google Scholar]
- 44.Mladenovic T, Zivic F, Petrovic N, Njezic S, Pavic J, Kotorcevic N, et al. Application of silicone in ophthalmology: a review. Materials. 2024;17:3454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Tsukeoka T, Suzuki M, Ohtsuki C, Sugino A, Tsuneizumi Y, Miyagi J, et al. Mechanical and histological evaluation of a PMMA-based bone cement modified with g-methacryloxypropyltrimethoxysilane and calcium acetate. Biomaterials. 2006;27:3897–903. [DOI] [PubMed] [Google Scholar]
- 46.Chuchulska B, Dimitrova M, Vlahova A, Hristov I, Tomova Z, Kazakova R. Comparative analysis of the mechanical properties and biocompatibility between CAD/CAM and conventional polymers applied in prosthetic dentistry. Polymers. 2024;16:877. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Islam MA, Mobarak MH, Rimon MIH, Al Mahmud MZ, Ghosh J, Ahmed MMS, et al. Additive manufacturing in polymer research: advances, synthesis, and applications. Polym Test. 2024;132:108364. [Google Scholar]
- 48.Martins F, Branco P, Reis J, Barbero Navarro I, Maurício P. Dimensional stability of two impression materials after a 6-month storage period. Acta Biomater Odontol Scand. 2017;3:84–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Xu M, Zhang J, Li B, Fan Z, He L. Effects of the backbone’s structures on the curing behaviors and properties of phthalonitrile containing benzoxazine rings. Molecules. 2024;29:5637. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Savabi O, Nejatidanesh F, Sharifi M, Shirani M, Valanezhad A, Watanabe I, et al. Cytotoxicity of dental ceramic materials on human gingival fibroblasts. Dent Res J. 2024;21:53. [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data are available to the corresponding author upon request.



