Next Article in Journal
Applications of AI and VR in High-Risk Training Simulations: A Bibliometric Review
Previous Article in Journal
Best Practices Evidenced for Software Development Based on DevOps and Scrum: A Literature Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Dissipative Particle Dynamics Study on the Formation of the Water-In-Petroleum Emulsion: The Contribution of the Oil

1
School of Chemistry and Chemical Engineering, Southwest Petroleum University, Chengdu 610500, China
2
School of Petroleum Engineering, Southwest Petroleum University, Chengdu 610500, China
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2025, 15(10), 5422; https://doi.org/10.3390/app15105422
Submission received: 24 March 2025 / Revised: 6 May 2025 / Accepted: 7 May 2025 / Published: 13 May 2025

Abstract

:
High internal phase emulsions (HIPEs), in which the dispersed water phase exceeds 70%, play a critical role in enhancing oil recovery through in situ permeability modification. However, their stability remains a major challenge due to frequent phase inversion and coalescence. In this study, the formation and stabilization mechanisms of water-in-oil HIPEs were investigated using a multiscale modeling approach that combines dissipative particle dynamics (DPD), molecular dynamics (MD), and density functional theory (DFT). Fourteen oil types and six polyaromatic emulsifiers with varying side-chain configurations and polar functional groups were modeled. Emulsifier performance was evaluated across 42 DPD-simulated systems with 70% and 80% water content. The results showed that emulsifiers with moderate dipole moments (~6 Debye) and spatially distributed heteroatoms achieved the most stable emulsion structures, forming continuous interfacial films or micelle-bridged networks. In contrast, emulsifiers with weak polarity (<1 Debye) or excessive stacking tendencies failed to prevent phase separation. The HOMO–LUMO energy gap and cohesive energy density (CED) were found to be poor predictors of emulsification performance. Four distinct stabilization mechanisms were identified, including interfacial film co-construction with oils and steric stabilization via side-chain architecture. The findings demonstrate that dipole moment is a reliable molecular descriptor for emulsifier design. This study provides a theoretical foundation for the rational development of high-performance emulsifiers in petroleum-based HIPE systems and highlights the potential of multiscale simulations in guiding formulation strategies.

1. Introduction

1.1. General Overview

Water-in-oil emulsions are a common occurrence in petroleum production and processing systems [1,2]. These emulsions form when water becomes finely dispersed in the continuous oil phase, typically stabilized by naturally occurring surface-active compounds such as asphaltenes or externally introduced emulsifiers [3,4]. In the petroleum industry, such emulsions can significantly affect flow assurance and the overall process of separation, which in turn affects the recovery of the reservoir process [5,6,7]. While in some cases emulsions pose operational challenges, their controlled formation and stability can also be harnessed for enhancing oil recovery and drilling fluid formulation [8,9].
A particularly interesting subclass of emulsions is the high internal phase emulsion (HIPE), defined by a dispersed phase volume exceeding 70% [10,11,12,13]. In the context of water-in-oil systems, HIPEs exhibit unique structural and rheological properties, including high viscosity and elastic behavior, which make them valuable for in situ permeability modification during enhanced oil recovery [14,15,16]. By selectively blocking high-permeability zones, stable W/O HIPEs can redirect flow into under-swept areas of the reservoir. This improves sweep efficiency; however, the formation of stable HIPEs is inherently difficult as increased water content raises the risk of phase inversion and droplet coalescence, especially in the absence of well-designed emulsifiers.
In recent years, simulation techniques have become indispensable tools across multiple disciplines that enable researchers to predict and optimize complex multi-phase behaviors at different scales [17,18,19,20,21,22]. Within this context, dissipative particle dynamics (DPD) has emerged as a powerful mesoscale simulation method for modeling soft matter systems and capturing hydrodynamic behavior in emulsions and polymers [23,24]. Compared to atomistic simulations, DPD offers significant computational advantages for simulating larger systems over longer timescales. It has been successfully used to explore surfactant aggregation and interfacial film dynamics in various emulsion systems [25,26]. The ability of DPD to bridge microscopic interaction rules with macroscopic behavior makes it particularly suited for studying the formation and stabilization of HIPEs.

1.2. Literature Review

Several recent studies have utilized dissipative particle dynamics (DPD) and multiscale modeling to investigate the interfacial behavior and emulsion formation mechanisms in oil–water systems. These works span a variety of applications, including emulsion stabilization, demulsification, interfacial tension reduction, and enhanced oil recovery using surfactant/polymer systems. Table 1 provides a summary of key state-of-the-art literature, highlighting the method applied, the main findings, and limitations relevant to the scope of this study.

1.3. Gap and Novelty

Despite the growing body of research employing DPD to analyze oil–water emulsions, most existing studies focus either on reducing interfacial tension using conventional surfactants or on demulsification processes under simplified binary or ternary systems. Limited attention has been paid to the stabilization of HIPEs in water-in-oil configurations, particularly under conditions where the dispersed water content exceeds 70%. Moreover, few works have explored the cooperative role of emulsifier architecture and oil polarity in HIPE formation using a multiscale approach. The current study addresses this gap by employing a DFT–MD–DPD integrated simulation pipeline to systematically investigate how polyaromatic emulsifier structure, dipole moment, and oil phase composition collectively influence HIPE stability. The novelty of this work lies in identifying dipole moment as a predictive molecular descriptor for emulsifier performance and proposing four stabilization mechanisms observed directly through mesoscale simulation.

2. Methodology

This study employed a multiscale simulation framework to investigate the formation and stabilization of water-in-oil HIPEs. The approach integrates quantum-level, atomistic, and mesoscale modeling tools to capture the behavior of emulsifiers and oil molecules across different length and time scales. Each scale contributes unique physical insights, which are translated into parameters for the next stage of simulation, enabling a seamless transition from molecular properties to emulsion morphology prediction.
The overall strategy, illustrated in Figure 1, begins with the definition of the research objective: to simulate the formation of water-in-oil HIPEs and understand the influence of emulsifier properties on system stability. The methodology is driven by a structured sequence of tasks starting with molecular setup, where an emulsifier, oil type, and water-to-oil ratio are selected. This is followed by property calculations using density functional theory (DFT) to obtain dipole moments and electronic descriptors. A decision checkpoint is incorporated into the workflow, where emulsifiers with dipole moments between 4–7 Debye are prioritized, as this range was preliminarily identified as optimal for interfacial stabilization.
Once the electronic properties are determined, the simulation proceeds with mapped input parameters into the DPD environment. A full simulation process is carried out using mesoscale coarse-grained models under NVT (number of particles, volume, and temperature) ensemble conditions. Afterward, post-simulation analysis evaluates the resulting structures to classify whether a high internal phase emulsion has successfully formed. The results are then used to establish correlations between emulsifier molecular descriptors and emulsion performance.

3. Physics-Based Particle Simulation Approach

3.1. Overview of the DPD Simulation Method

DPD was employed as the primary simulation technique to model the mesoscale behavior of emulsifier–oil–water systems. DPD is a coarse-grained molecular dynamics approach that enables the simulation of larger spatial and temporal scales by grouping atoms into beads and applying simplified interaction rules. Each DPD particle represents a cluster of molecules, and their interactions are governed by Newtonian dynamics incorporating conservative, dissipative, random, and bonding forces [36]. This method is particularly suited for capturing the hydrodynamic behavior of multiphase systems such as emulsion.
In DPD simulations, the total force acting on a particle arises from the interplay of several components, primarily F i j C , the dissipative force F i j D , and the random force F i j R . In addition, a spring force f j S was also taken into consideration, as described below [37]:
f = i j ( F ij C + F ij D + F ij R ) + f i S
where F i j C is the conservative force (repulsion force) and models excluded volume effects, F i j D is the dissipative force, introducing friction to maintain local momentum conservation, F i j R is the random force, representing thermal fluctuations, and f j S is the spring force which accounts for bonding interactions between connected beads. The conservative force describes the maximum repulsive interaction between beads i and j, given as [38]:
F i j C = a i j 1 r i j r ^ i j if   r i j < 1 0 if   r i j 1
where aij is the maximum repulsion parameter derived from the solubility parameters of the beads (typically calculated via atomistic MD). rij is the distance between the beads, and r ij ^ is the unit vector along the bead–bead separation. The dissipative force introduces friction proportional to the relative velocity between beads [38]:
F i j D = γ ω D r i j r i j ^ v i j r i j if   r i j < 1 0 if   r i j 1
where γ is a friction coefficient, vij is the relative velocity of beads i and j, and ωDrij is a weight function dependent on the inter-particle distance. The random force accounts for thermal noise and is expressed as [38]:
F i j R = σ ω R r i j θ i j r i j ^ if   r i j < 1 0 if   r i j 1
where θ is a random flocculation variable between 0~1, σ is the amplitude of the noise, and ωRrij is a weighting function related to ωDrij [38]:
ω D r = [ ω R r ] 2 = ( 1 r ) 2 , r ij < 1 0 , r ij 1
The amplitude of the random force (σ) and the dissipative force coefficient (γ) are related to the system’s temperature (T) through the fluctuation–dissipation theorem:
σ 2 r = 2 γ k B T
where kB is the Boltzmann constant while T is the absolute temperature.
All simulations were performed using dimensionless DPD units with a reduced cutoff radius R c = 1 , mapped to 3.23 Å. The simulation timestep was 0.005 t c , and the system temperature was maintained at 65 °C in the NVT ensemble over 1,000,000 steps [39].

3.2. Molecular Mapping and Coarse-Graining

Water, oil, and emulsifier molecules were mapped to coarse-grained DPD beads. Each water bead (W) represented three water molecules (54 g/mol). The emulsifiers were modeled based on graphene-like polyaromatic cores (PACs) with variable side chains:
  • AS1: PAC with no side chains;
  • AS2: PAC with 4 aliphatic + 2 heteroatom side chains;
  • AS3: AS2 with 2 added perpendicular aliphatic chains.
Four oil types were modeled:
  • Oil1–4: Alkanes;
  • Oil5–7: Amides;
  • Oil8–10: Ethers;
  • Oil11–14: Aromatics.
For the simulation of the emulsification process, a 3D periodic model was used to construct the emulsion system. The oil, emulsifier, and water molecules were packed into a simulation cell with dimensions 100 × 100 × 100 R3, creating a mixture with a water-to-(emulsifier plus oil) volume ratio of either 7:3 or 8:2. The emulsifier concentration was fixed at a 1:1 mass ratio relative to oil, as concentration variation was beyond the scope of this study. All molecules were randomly distributed within the simulation box to represent the pre-equilibrium state typically observed in real emulsification processes, such as high-shear mixing. This approach is widely adopted in DPD studies, as the system naturally evolves toward thermodynamic equilibrium over the course of the simulation. Each model was labeled ASxOy, where ASx (x = 1–3) denotes the emulsifier type and Oy (y = 1–14) indicates the oil type. To study interfacial behavior, an additional slab configuration was used, consisting of two oil slabs (each 25 Rc thick) separated by a central water slab (50 Rc thick), with 20% emulsifier added to the oil slabs to ensure sub-monolayer coverage. The IFT was calculated using Equation (7) [40]:
IFT = L x ( p xx p zz + p y y 2 )
where Lx is the cell length along the x direction (normal to the interface) and pxx, pyy, and pzz are the pressures in the x, y, and z directions, respectively. Each simulation was conducted in NVT ensemble at 65 °C for 1,000,000 steps with a time step of 0.005 tc (approx. 15 fs), followed by geometry optimization.
The molecular mapping and composition of the coarse-grained beads for water, oils, and emulsifiers are illustrated in Figure 2, which visually summarizes the core structures and corresponding bead representations used in the DPD simulation.

3.3. Force Parameter Derivation from MD

As previously mentioned, the conservative (repulsion) force parameter is critical in DPD simulations. The conservative force parameter aij between beads i and j was calculated using the Flory−Huggins parameter χij and the solubility parameter δ, as described by Groot et al. [38]:
a ij = a ii + 3.27 χ ij
χ ij = v ij RT ( δ i δ i ) 2
Here, aii was set to be 78, and vij represents the average molar volume of species i and j, which was determined using reference data or standard manuals. The aij values for all bead types are listed in Table 2. The solubility parameter (δ) was calculated using all-atom MD simulations with the Forcite module in the Materials Studio software 7.0 package, employing the COMPASS force field to model non-bonding interactions. All-atom MD simulations were also utilized to determine certain bulk molecular properties.

3.4. Density Functional Theory (DFT) Calculations for Molecular Descriptors

To obtain electronic properties, DFT calculations were performed using the PW91 functional and DNP basis set. Geometry optimization was carried out on all-atom structures of the emulsifiers and oils, and the following properties were extracted:
  • Highest Occupied Molecular Orbital (HOMO) energy;
  • Lowest Unoccupied Molecular Orbital (LUMO) energy;
  • Energy gap (ELUMO-HOMO);
  • Dipole moment (Debye).
These values were used to correlate electronic structure with emulsification performance in later sections. The dipole moment, in particular, was found to relate strongly to interfacial stability outcomes, as will be discussed in Section 4.

3.5. Simulation Model Configuration and IFT Calculation

Each emulsification simulation was executed using a periodic cubic box with dimensions of 100 × 100 × 100 R c 3 . Systems contained water, oil, and emulsifier beads, mixed at 70% or 80% water content while keeping the emulsifier-to-oil mass ratio at 1:1. Configurations were randomly initialized to mimic high-shear dispersion. A slab model was used to evaluate IFT which consisted of two oil layers (25 R c 1 ) and a central water layer (50 R c 1 ), with 20% emulsifier to ensure sub-monolayer coverage. Simulations ran in the NVT ensemble for 1,000,000 steps at 65 °C with a 0.005tc timestep. Analyses included Radial Distribution Functions (RDFs), visual morphology, and phase classification. Table 3 below lists the parameters for the purpose of replication.
Analyses included visual morphology, phase classification, and radial distribution functions (RDFs). The RDF was computed to quantify the spatial arrangement of molecules, particularly focusing on the distribution of emulsifier molecules around each other and around oil molecules at the interface. RDFs were calculated using center-of-mass distances between coarse-grained beads, with a cutoff radius of 10 Rc. The results were averaged over the final 200,000 steps of the simulation to ensure equilibrium sampling. The computed RDFs help to identify stacking behavior, interfacial adsorption, and dispersion tendencies of different emulsifiers across the simulated systems [41,42,43,44,45]. These descriptors have previously been explored in machine learning contexts and remain valuable for future data-driven investigations [46]. In this work, RDFs were analyzed strictly as physical indicators of interfacial organization and packing efficiency [47].

3.6. Validation and Assumptions

Model reliability was ensured through internal validation measures. Conservative force parameters derived from MD-based solubility data were verified through multiple molecular inputs. However, DPD simulations do not model chemical bonding or reaction mechanisms. This restricts the scope to non-reactive mesoscopic dynamics. Key assumptions in the simulation design are summarized in Table 4.

4. Results and Discussion

4.1. Interfacial Activity and Aggregation Behavior of Emulsifiers

Understanding how emulsifiers behave at the oil–water interface is fundamental to predicting and controlling emulsion stability. This section explores the IFT of various oil systems in the absence of emulsifiers, followed by a detailed analysis of how molecular structure influences emulsifier dispersion, aggregation, and interfacial coverage. The findings are supported by visualizations of concentration profiles and RDFs derived from simulation data, as illustrated in Figure 3.
The IFT of the oil–water system in the absence of any emulsifier is presented in Figure 3a. The values range from approximately 41.2 mN/m for Oil 5 to 66.4 mN/m for Oil 14. The presence of emulsifiers was excluded in these measurements because emulsifier migration into the water phase deformed the planar interface, making Equation (7) invalid for precise IFT evaluation. This behavior is typically associated with the Marangoni effect, in which surface tension gradients drive interfacial flow and destabilize planar configurations [48]. Among the oil types, amide-based oils (Oil 5–7) displayed the lowest IFT values, averaging around 42–45 mN/m, reflecting strong interfacial activity. This was attributed to their polar functional groups, which enhanced interactions with the aqueous phase. In particular, Oil 5 exhibited the lowest IFT at 41.2 mN/m, while Oils 6 and 7 followed with values of 43.6 mN/m and 44.9 mN/m, respectively. Furthermore, an increasing trend in IFT was observed with longer aliphatic chains within these oils—indicating that extended chains dilute the polar influence of the headgroup, thereby weakening oil–water interfacial interactions.
Conversely, ether-based oils (Oil 8–10) exhibited moderate IFT values ranging from 50.3 mN/m (Oil 8) to 52.8 mN/m (Oil 10), due to their lower amphiphilicity. Alkanes (Oil 1–4) and aromatics (Oil 11–14), lacking significant polar functionality, showed higher IFT values: 55.1–62.7 mN/m for alkanes and 64.2–66.4 mN/m for aromatic oils. These results underscore the crucial influence of oil molecular structure—particularly the presence and arrangement of polar groups—on interfacial behavior. A higher balance between hydrophilic and hydrophobic components in oil molecules improves their alignment at the interface, promoting interfacial tension reduction.
The aggregation and dispersion behaviors of various emulsifiers at the interface were also examined and are depicted in Figure 3b–f. AS1, which lacks side chains, formed compact nanomicelles composed of 3 to 5 PAC sheets via π–π stacking. Despite this aggregation, the interfacial coverage remained below a monolayer, as reflected in the broad concentration profile in Figure 3b, indicating weak confinement and lower surface activity. In contrast, emulsifiers with side chains—AS2 and AS3—showed enhanced interfacial spreading. For the interfacially inactive Oil 11 (toluene), both emulsifiers formed stable monolayers (Figure 3c,d). However, with interfacially active Oil 10 (ether), AS2 exhibited displacement from the interface due to competition with the oil molecules, resulting in cluster formation (Figure 3e). AS3, due to its side chains oriented perpendicular to the PAC plane, successfully resisted stacking and remained monodispersed at the interface (Figure 3f). Further insight was provided by the RDF analyses shown in Figure 3g,h. AS1 exhibited a strong RDF peak at 3.4 Å, corresponding to the characteristic π–π stacking distance between PAC sheets, confirming its aggregation tendency. In contrast, AS2 and AS3 displayed weaker, broader peaks centered at 3.2–3.3 Å which indicates more distributed packing due to the influence of side-chain sterics. This structural difference directly impacted their ability to remain interfacially active under competitive conditions.
To evaluate the accuracy of the DPD model in predicting interfacial properties, simulated IFT values were compared with data reported in previous literature using experimental and DPD-based approaches. As summarized in Table 5, the simulated trend—where amide-based oils (Oils 5–7) exhibited the lowest IFT, followed by ethers (Oils 8–10), and then non-polar alkanes and aromatics (Oils 1–4, 11–14)—correlates well with previously reported IFT ranges. For instance, Ruiz-Morales et al. [25] observed IFT values between 35–40 mN/m for oxygenated asphaltenes, while Jiang et al. [23] showed that asphaltene adsorption could reduce IFT by approximately 8 mN/m. These consistent trends reinforce the validity of the simulation model and its parametrization. Although absolute IFT values were not extracted here due to simulation complexity, the qualitative alignment with experimental behavior supports the credibility of the DPD method for modeling emulsion interfacial dynamics.

4.2. Structural Evolution and Formation Conditions of High Internal Phase Emulsions

The formation of high HIPEs was investigated across various combinations of oils and emulsifiers under controlled water content conditions. Table 6 summarizes the simulation results for the systems with 70% water content. Each system was simulated three times to confirm the stability and reproducibility of the emulsion morphology. A positive outcome (‘+’) indicates successful formation of a porous W/O emulsion structure in at least one of the three simulations. Systems with 80% water content exhibited minimal emulsion formation, with only two models forming stable W/O structures. Given the poor performance at this higher water content, these results are excluded from detailed analysis, and the discussion focuses solely on the more informative 70% systems.
Several models displayed porous 2D or 3D network morphologies that extended across the simulation box and were structurally comparable to HIPEs observed in experimental micrographs [49,50]. These interconnected frameworks are indicative of strong emulsification driven by the combined interfacial properties of the oil and the molecular architecture of the emulsifier. As such, the simulation dataset in Table 6 forms the basis for analyzing the influence of both oil polarity and emulsifier structure on the ability to generate stable emulsions.

4.3. Influence of Oil Type and Emulsifier Structure on Emulsion Formation

4.3.1. Effect of Oil Polarity on HIPE Morphology

The probability of HIPE formation varied significantly depending on the oil type, as different oils contributed differently to the stabilization of the interfacial film based on their polarity and interfacial activity. For the interfacially active amide-based oils (Oil 5–7), seven out of nine emulsifier–oil combinations yielded stable emulsions. All three emulsifiers were able to stabilize Oil 5 effectively, as shown in the porous frameworks depicted in Figure 4(a1–c1). Similarly, the ether-based oils (Oil 8–10), although less active, still formed stable W/O emulsions with AS2 and AS3 emulsifiers (Figure 4(a2–c2)). Conversely, these ether and amide oils failed to form HIPE with AS1, likely due to its limited interfacial activity, and this will be further discussed in the subsequent subsection.
In contrast, both the alkane oils (Oil 1–4) and the aromatic oils (Oil 11–14), which exhibited minimal polarity and interfacial activity, showed a lower tendency to form stable emulsions. Among these, only 8 out of 12 alkane models and 3 out of 12 aromatic models achieved successful HIPE formation, indicating the critical importance of oil polarity and structural composition in emulsion stability (Figure 4(a3–c4)).

4.3.2. Effect of Emulsifier Structure on Interfacial Stabilization

To assess the role of emulsifier design in HIPE formation, results across emulsifiers were evaluated (see Table 6). The structural presence of side chains had a substantial impact. AS2 generated 12 positive HIPE outcomes, followed by AS3 with 9, while AS1—lacking side groups—achieved only 3 successful models.
The poor performance of AS1 is attributed to its tendency to self-assemble into stacked nanomicelles due to strong π–π interactions, as confirmed by the RDF peak near 3.4 Å. These nanomicelles occasionally bridged droplets to form Pickering-type emulsions (Figure 5a) but often failed to maintain contact across the oil phase, which results in phase separation as Figure 5b proves. RDF profiles further confirmed weak dispersion and low interfacial confinement of AS1. In contrast, AS2 and AS3 showed improved dispersion due to their steric hindrance from side chains, enabling more stable monolayer formation (Figure 5c,e). Despite this, some cases still exhibited instability and transition to O/W emulsions (Figure 5d,f), likely due to an imbalance between van der Waals attraction and stacking tendencies. These observations reinforce the significance of both molecular orientation and intermolecular interactions in emulsion engineering.

4.4. Influence of Emulsifier Diffusion and Association on Emulsion Stabilization

The stabilization of oil droplets within HIPEs depends critically on the rapid interfacial migration and association behavior of emulsifier molecules. In the absence of sufficient interfacial coverage, oil droplets tend to coalesce in order to minimize surface energy. This phenomenon necessitates the dynamic diffusion of emulsifier molecules toward the interface to arrest droplet deformation and maintain emulsion morphology. As demonstrated in Figure 6(a1–e1), when adequate quantities of emulsifier reached the droplet interface in time, the structural integrity of the dispersed oil phase was preserved, resulting in successful W/O emulsion formation. Conversely, insufficient emulsifier presence led to unstable interfacial structures and eventual phase transition, as shown in Figure 6(a2–e2). This transition was particularly apparent in emulsifier–oil combinations lacking strong interfacial interactions.
Moreover, the role of π–π stacking among emulsifier molecules emerged as a key factor in providing additional mechanical reinforcement to the interfacial film. Emulsifier AS2, in particular, demonstrated the ability to form dimeric and trimeric stacking configurations that effectively bridged adjacent droplets, thereby withstanding the destabilizing forces within the emulsion system. This was especially critical when stabilizing aromatic oils that inherently possess high interfacial tension and strong coalescence tendencies. As shown in Figure 7a,b, AS2 maintained its integrity through stacking bridges even under such challenging conditions, while AS3 failed to form such networks. The stabilizing mechanisms observed across the simulations could be categorized into four distinct modes: (1) nanomicelle-based steric hindrance and bridging by stacked emulsifier aggregates (e.g., AS1), (2) interfacial bridging with non-polar aliphatic oils (e.g., AS2/AS3 with alkanes), (3) polar functional group interactions with moderately active oils (e.g., AS2/AS3 with amides or ethers), and (4) full interface coverage and bridging by the emulsifier itself, particularly for challenging systems like aromatics (e.g., AS2 with aromatic oils). These mechanisms collectively emphasize the critical roles of van der Waals interactions and π–π stacking in sustaining HIPE morphology.

4.5. Influence of System Size on HIPE Formation

To examine the effect of system scaling on emulsification behavior, a larger simulation cell with dimensions of 200 × 200 × 200 Rc was constructed. Under these expanded conditions, none of the models succeeded in forming a HIPE. This outcome suggests a pronounced size effect, where the balance between droplet separation distance and emulsifier availability plays a pivotal role in emulsion stability. As observed previously in Figure 7, the emulsifier AS2 forms structural bridges across droplets through dimeric and trimeric π–π stacking arrangements. In smaller systems, these stacked emulsifier assemblies can successfully span the inter-droplet gaps. However, as the cell size increases, the average distance between droplets also increases, necessitating a higher quantity of emulsifier to maintain interfacial integrity and continuity. When this bridging threshold is not met, droplet coalescence becomes energetically favorable, inhibiting stable HIPE formation.
It is also worth noting that the emulsifier structures used in this study possess larger molecular dimensions than those described in the classical Yen–Mullins model for asphaltenes [31,32]. The conventional asphaltene model suggests smaller aromatic cores and limited side-chain functionality, which may be insufficient to provide the steric coverage and bridging capability required for HIPE stabilization, particularly in larger systems. This observation provides a plausible explanation for the limited emulsification capacity of conventional asphaltenes in macroscale oil–water systems. The detailed relationship between droplet size, emulsifier molecular architecture, and interfacial behavior under different cell dimensions remains an area of active investigation. Further parametric studies involving size-scaling, emulsifier concentration variation, and interface curvature effects are warranted to generalize the size–emulsification correlation.

4.6. Enhanced Performance of Structurally Modified Emulsifiers

To further explore the influence of molecular architecture on emulsion stability, two modified emulsifiers—AS4 and AS5—were designed based on the PAC core structure of AS2. These new emulsifiers incorporate additional functional side groups, specifically amide and carboxylic heteroatoms, positioned differently along the aliphatic chains. This structural modification was intended to promote intermolecular interactions with the aqueous phase and improve interfacial film stability. The emulsification performance of these modified molecules was remarkable. As illustrated in Figure 8a–c, both AS4 and AS5 retained the characteristic PAC base but exhibited extended polar side chains. Simulations demonstrated that for all 14 oil types, AS4 and AS5 consistently formed water-in-oil (W/O) emulsions, resulting in a 100% HIPE success rate (see Table 7). In contrast, AS6—a third emulsifier with a similar backbone but lacking polar heteroatom groups—exhibited significantly diminished performance and formed stable emulsions in only 6 of the 14 oil models.
Visualizations of AS4O11 and AS4O13 in Figure 8d–e revealed dense emulsifier coverage at the interface. These emulsifiers formed monolayers that were uniformly distributed without reliance on interfacial activity from the oil phase. Their ability to independently stabilize emulsions underscores the critical role of heteroatom functionality in modulating interfacial interactions. The polar groups likely enhance hydrogen bonding or dipolar alignment with water, thereby strengthening the interfacial film. These findings suggest that rational molecular design—particularly the strategic placement of heteroatoms—can significantly elevate emulsification capacity in HIPE systems. The results presented here provide a clear guideline for designing future emulsifiers with optimized interfacial performance tailored for diverse oil compositions.

4.7. Correlation Between Molecular Electronic Properties and Emulsification Stability

While DPD simulations do not capture chemical bond formation or electron transfer, the molecular electronic properties of emulsifiers still offer valuable insights into their ability to stabilize emulsions. To evaluate this, DFT calculations were conducted on all emulsifiers and oils used in this study to quantify key descriptors such as frontier orbital energies, dipole moments, and CED. Figure 9 summarizes these DFT-derived parameters. The HOMO and LUMO energy levels of each molecule are shown in Figure 9a,b, with their differences (ΔE = ELUMOEHOMO) presented in Figure 9c. These descriptors are often used to predict chemical reactivity or electron delocalization. However, as revealed by Table 8, emulsifiers such as AS1 and AS6, which possess moderate gaps (2.239 eV and 1.626 eV, respectively), failed to generate stable high internal phase emulsions (HIPEs). In contrast, AS4 and AS5, with slightly lower gaps (1.610–1.679 eV), consistently produced stable emulsions. Thus, energy gap alone did not correlate with emulsifier effectiveness, indicating it is not a decisive predictor of HIPE stability.
Dipole moment, however, showed a more consistent trend. As illustrated in Figure 9d, AS4 and AS5, both with dipole moments close to 6 Debye, achieved the highest HIPE success rates. In contrast, AS1 and AS6, with dipole moments below 1 Debye, were among the least effective. Interestingly, AS2 (10.04 D) and AS3 (4.61 D) exhibited intermediate performance, suggesting that moderate polarity—rather than extremes—is optimal for achieving emulsifier stability in oil–water systems. These findings are reinforced by the quantitative data in Table 8 that support the utility of dipole moment as a reliable predictor. The role of CED, presented in Figure 9e, was also examined. CED quantifies internal molecular cohesion and is sometimes used to infer interaction potential with other phases. However, no clear correlation was observed between CED and HIPE formation. For example, Oil 11 and Oil 12 had high CED values but did not yield stable emulsions, indicating that bulk cohesive strength may play a secondary role compared to interfacial forces such as van der Waals interactions and dipole alignment.
For structural context, Figure 10 depicts the all-atom molecular structures used in the DFT study, highlighting the polyaromatic cores and various side chains. These architectures influence both molecular polarity and spatial coverage at the interface. Moreover, Figure 11 visualizes the HOMO and LUMO distributions of representative emulsifiers. These orbitals were localized within the PAC region, suggesting a lack of orbital overlap with surrounding molecules. This supports the conclusion that emulsification in these systems is governed by non-bonding interactions—rather than chemical bonding or charge transfer.
In summary, while the HOMO–LUMO energy gap and cohesive energy density provide some insights, it is the dipole moment that shows the strongest correlation with emulsifier performance. These findings suggest that appropriate polarity enables the formation of stable interfacial films through non-covalent interactions, particularly van der Waals forces. Although dipole moment has been previously linked to electro-coalescence phenomena in petroleum systems [51], this study presents the first application of dipole correlation to HIPE stabilization, offering valuable guidance for designing future emulsifiers with targeted electronic properties.

5. Conclusions

A multiscale simulation study was conducted to investigate the stabilization mechanisms of water-in-oil HIPEs, with up to 70–80% water content, using DPD, MD, and DFT. A total of 42 emulsion models were simulated using three core emulsifiers (AS1–AS3) and fourteen oil types which revealed key relationships between molecular structure and emulsion stability. Among the tested emulsifiers, AS2 and AS3 exhibited significantly higher performance, achieving HIPE formation in 12 and 9 cases, respectively, compared to only 3 cases for AS1. The superior performance of AS2 and AS3 was attributed to enhanced interfacial dispersion, steric effects, and molecular polarity. Modified emulsifiers AS4 and AS5 further improved performance as they successfully stabilized HIPEs in all 14 oil cases for the benefit of optimized side-chain architecture and heteroatom inclusion.
DFT results showed that AS4 and AS5 possessed dipole moments of approximately 6.1 Debye, whereas poorly performing emulsifiers AS1 and AS6 had values below 1.0 Debye. Emulsifiers with excessively high dipole moments, such as AS2 (10.04 Debye), exhibited moderate stability, indicating that an optimal polarity range enhances interfacial film formation. In contrast, the HOMO–LUMO energy gap (ranging from 0.029 to 0.082 Ha) and CED showed weak correlation with emulsification performance, suggesting that these descriptors are less reliable in predicting HIPE stability. Visualizations of orbital distributions confirmed that HOMO and LUMO regions were localized within the PAC as it limited electron delocalization and supported the dominance of non-bonding interactions in stabilization. Four distinct mechanisms of HIPE stabilization were identified: nanomicelle bridging, emulsifier–oil co-construction of interfacial films, side-group anchoring in interfacially inactive oils, and full interface coverage by emulsifier monolayers. This study demonstrates that dipole moment is the most predictive molecular property for HIPE stabilization, introducing a new descriptor for emulsifier design. These findings provide actionable criteria for the development of next-generation surfactants in petroleum recovery, chemical flooding, and complex emulsion technologies.
Although these simulations provide valuable mechanistic insights into HIPE stabilization, the absence of direct experimental validation remains a notable limitation. To enhance the credibility and applicability of the findings, future research should incorporate experimental comparisons, particularly through interfacial tension measurements, microscopic characterization of emulsion microstructures, and rheological analyses. Additionally, the current DPD framework does not account for potential chemical adsorption or reactivity at the interface. Addressing this limitation will require the development and integration of reactive force fields capable of capturing chemical bonding phenomena relevant to complex emulsion systems.

Author Contributions

Methodology, M.H.O.O.; Software, X.F.; Validation, S.F.; Formal analysis, W.P.; Writing—review and editing, P.S.; Supervision, R.L.; Project administration, M.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research and the APC were funded by National Natural Science Foundation of China (Grant No. U19B2010, 42172347).

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

The authors are deeply thankful for the financial support from the National Natural Science Foundation of China (Grant No. U19B2010, 42172347). The authors are also truly thankful for the help from the National Supercomputing Center in Shenzhen (Shenzhen Cloud Computing Center) of China in support of the computation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sousa, A.M.; Pereira, M.J.; Matos, H.A. Oil-in-Water and Water-in-Oil Emulsions Formation and Demulsification. J. Pet. Sci. Eng. 2022, 210, 110041. [Google Scholar] [CrossRef]
  2. Ariffin, T.S.T.; Yahya, E.; Husin, H. The Rheology of Light Crude Oil and Water-In-Oil-Emulsion. Procedia Eng. 2016, 148, 1149–1155. [Google Scholar] [CrossRef]
  3. Keleşoğlu, S.; Pettersen, B.H.; Sjöblom, J. Flow Properties of Water-in-North Sea Heavy Crude Oil Emulsions. J. Pet. Sci. Eng. 2012, 100, 14–23. [Google Scholar] [CrossRef]
  4. Romanova, Y.N.; Koroleva, M.Y.; Musina, N.S.; Maryutina, T.A. Ultrasonic Demulsification of Water-in-Crude Oil Emulsions: Influence of Rheological Properties. Chem. Eng. Process.-Process Intensif. 2025, 211, 110242. [Google Scholar] [CrossRef]
  5. Wong, S.F.; Lim, J.S.; Dol, S.S. Crude Oil Emulsion: A Review on Formation, Classification and Stability of Water-in-Oil Emulsions. J. Pet. Sci. Eng. 2015, 135, 498–504. [Google Scholar] [CrossRef]
  6. Kolotova, D.S.; Kuchina, Y.A.; Petrova, L.A.; Voron’ko, N.G.; Derkach, S.R. Rheology of Water-in-Crude Oil Emulsions: Influence of Concentration and Temperature. Colloids Interfaces 2018, 2, 64. [Google Scholar] [CrossRef]
  7. Yonguep, E.; Kapiamba, K.F.; Kabamba, K.J.; Chowdhury, M. Formation, Stabilization and Chemical Demulsification of Crude Oil-in-Water Emulsions: A Review. Pet. Res. 2022, 7, 459–472. [Google Scholar] [CrossRef]
  8. Ghannam, M.T. Water-in-Crude Oil Emulsion Stability Investigation. Pet. Sci. Technol. 2005, 23, 649–667. [Google Scholar] [CrossRef]
  9. Umar, A.A.; Saaid, I.B.M.; Sulaimon, A.A.; Pilus, R.B.M. A Review of Petroleum Emulsions and Recent Progress on Water-in-Crude Oil Emulsions Stabilized by Natural Surfactants and Solids. J. Pet. Sci. Eng. 2018, 165, 673–690. [Google Scholar] [CrossRef]
  10. Xu, C.; Sun, Y.; Feng, C.; Zhang, S. Porous PEG Scaffold Fabricated via Emulsion-Templating Technique Towards Immobilization of Saccharomyces Cerevisiae Cells. Catalysts 2024, 14, 809. [Google Scholar] [CrossRef]
  11. Wagle, V.; AlYami, A.; Safran, A. Designing Invert Emulsion Drilling Fluids for High Temperature and High-Pressure Conditions. Sci. Rep. 2024, 14, 27748. [Google Scholar] [CrossRef]
  12. Ehyaie, D.; Zaheri, P.; Samadfam, M.; Zahakifar, F. Evaluation of Cyanex272 in the Emulsion Liquid Membrane System for Separation of Thorium. Sci. Rep. 2024, 14, 31897. [Google Scholar] [CrossRef] [PubMed]
  13. Wu, Y.; Huang, J.; Guo, Z.; Yang, Q.; Xia, C.; Zheng, Z. Preparation of Polymerized High Internal Phase Emulsion Membranes with High Open-Cellular Extent and High Toughness via RAFT Polymerization. Polymers 2025, 17, 515. [Google Scholar] [CrossRef]
  14. Sadeghi, S.; Zargarzadeh, L. Gibbsian Surface Thermodynamic Analysis of Emulsion Liquid Membranes. Sci. Rep. 2025, 15, 6401. [Google Scholar] [CrossRef]
  15. Liu, S.; Yao, T.; Xia, D.; Liu, Q.; Tian, G.; Liu, Y. The Development and Preparation of Novel Gel Emulsion Systems Based on a Cholesterol Star-Shaped Derivative. Molecules 2025, 30, 787. [Google Scholar] [CrossRef] [PubMed]
  16. Yang, L.; Ge, J.; Wu, H.; Li, X.; Zhang, X.; Zhang, G. Study on Oil-in-Water Emulsions Stabilized by SiO2 Nanoparticles for Enhancing Oil Recovery in Harsh Reservoirs. Colloid. Polym. Sci. 2024, 302, 1985–1998. [Google Scholar] [CrossRef]
  17. Al-Haddad, L.A.; Mahdi, N.M. Efficient Multidisciplinary Modeling of Aircraft Undercarriage Landing Gear Using Data-Driven Naïve Bayes and Finite Element Analysis. Multiscale Multidiscip. Model. Exp. Des. 2024, 7, 3187–3199. [Google Scholar] [CrossRef]
  18. Al-Haddad, S.A.; Fattah, M.Y.; Al-Azawi, T.K.; Al-Haddad, L.A. Three-Dimensional Analysis of Steel Beam-Column Bolted Connections. Open Eng. 2024, 14, 20220579. [Google Scholar] [CrossRef]
  19. Fattah, M.Y.; Al-Haddad, L.A.; Ayasrah, M.; Jaber, A.A.; Al-Haddad, S.A. Coupled Finite Element and Artificial Neural Network Analysis of Interfering Strip Footings in Saturated Cohesive Soils. Transp. Infrastruct. Geotechnol. 2024, 11, 2168–2185. [Google Scholar] [CrossRef]
  20. Al-Haddad, L.A.; Ibraheem, L.; EL-Seesy, A.I.; Jaber, A.A.; Al-Haddad, S.A.; Khosrozadeh, R. Thermal Heat Flux Distribution Prediction in an Electrical Vehicle Battery Cell Using Finite Element Analysis and Neural Network. Green Energy Intell. Transp. 2024, 3, 100155. [Google Scholar] [CrossRef]
  21. Abdul-Zahra, A.S.; Ghane, E.; Kamali, A.; Farhan Ogaili, A.A. Power Forecasting in Continuous Extrusion of Pure Titanium Using Naïve Bayes Algorithm. Terra Joule J. 2024, 1, 2. [Google Scholar]
  22. Al-Haddad, L.A.; Kahachi, H.A.H.; Ur Rehman, H.Z.; Al-Zubaidi, A.A.; Al-Karkhi, M.I.; Al-Oubaidi, B. Advancing Sustainability in Buildings Using an Integrated Aerodynamic Façade: Potential of Artificial Intelligence. Terra Joule J. 2024, 1, 1. [Google Scholar]
  23. Jiang, H.; Liu, X.; Liang, C.; Wang, Z.; Jia, Y. Dissipative Particle Dynamics to Study Asphaltenes and Surfactants Interactions at the Oil–Water Interface. J. Mol. Liq. 2023, 381, 121802. [Google Scholar] [CrossRef]
  24. Zhang, B.; Guan, B.; Liu, W.; Peng, B.; Cong, S. Dissipative Particle Dynamics Simulation and Microscopic Experimental Study of Emulsification Performance of Surfactant/Polymer Flooding. Processes 2023, 11, 1411. [Google Scholar] [CrossRef]
  25. Ruiz-Morales, Y.; Alvarez-Ramírez, F. Mesoscale Dissipative Particle Dynamics to Investigate Oil Asphaltenes and Sodium Naphthenates at the Oil−Water Interface. Energy Fuels 2021, 35, 9294–9311. [Google Scholar] [CrossRef]
  26. Liu, S.; Qiu, Y.; Liu, J.; Chen, X.; He, L.; Wang, Y.; Li, X.; Chao, L.; Yang, B.; Liu, T. Interfacial Competitive Behavior of Water/Oil/Surfactant Systems in Ultra-Deep Tight Reservoir—Insights from Dissipative Particle Dynamics. Colloids Surf. A Physicochem. Eng. Asp. 2025, 717, 136766. [Google Scholar] [CrossRef]
  27. Zhao, C.; Huang, H.; Li, J.; Li, Y.; Xiang, D.; Wu, Y.; Wang, G.; Qin, M. Facile Fabrication of Superhydrophobic Graphene/Polystyrene Foams for Efficient and Continuous Separation of Immiscible and Emulsified Oil/Water Mixtures. Polymers 2022, 14, 2289. [Google Scholar] [CrossRef] [PubMed]
  28. Han, D.; Mao, J.; Zhao, J.; Zhang, H.; Yang, X.; Lin, C. Effect of Surfactant Hydrophobic Chain Equivalence on the Oil-Water Interface and Emulsion Stability: A Dissipative Particle Dynamics and Experimental Study. J. Mol. Liq. 2023, 382, 121781. [Google Scholar] [CrossRef]
  29. Alvarez, F.; Flores, E.A.; Castro, L.V.; Hernández, J.G.; López, A.; Vázquez, F. Dissipative Particle Dynamics (DPD) Study of Crude Oil−Water Emulsions in the Presence of a Functionalized Co-Polymer. Energy Fuels 2011, 25, 562–567. [Google Scholar] [CrossRef]
  30. Song, X.; Shi, P.; Duan, M.; Fang, S.; Ma, Y. Investigation of Demulsification Efficiency in Water-in-Crude Oil Emulsions Using Dissipative Particle Dynamics. RSC Adv. 2015, 5, 62971–62981. [Google Scholar] [CrossRef]
  31. Han, D.; Mao, J.; Zhao, J.; Zhang, H.; Wang, D.; Wang, C.; Xue, J.; Cao, H.; Yang, X.; Lin, C.; et al. Dissipative Particle Dynamics Simulation and Experimental Analysis of the Effect of Anionic/Cationic Mixed Surfactants on the Stability of Emulsions. J. Mol. Liq. 2022, 367, 120482. [Google Scholar] [CrossRef]
  32. Goodarzi, F.; Zendehboudi, S. Effects of Salt and Surfactant on Interfacial Characteristics of Water/Oil Systems: Molecular Dynamic Simulations and Dissipative Particle Dynamics. Ind. Eng. Chem. Res. 2019, 58, 8817–8834. [Google Scholar] [CrossRef]
  33. Han, D.; Mao, J.; Zhao, J.; Zhang, H.; Wang, D.; Cao, H.; Yang, X.; Lin, C.; Zhang, Y. Dissipative Particle Dynamics Simulation and Experimental Analysis of Effects of Gemini Surfactants with Different Spacer Lengths on Stability of Emulsion Systems. Colloids Surf. A Physicochem. Eng. Asp. 2022, 655, 130205. [Google Scholar] [CrossRef]
  34. Liang, C.; Liu, X.; Jia, Y.; Jiang, H.; Xu, Y. Mechanistic Study of the Effects of Surfactants and Asphaltenes on the Action of Emulsions with Different Water Contents: Based on Dissipative Particle Dynamics. Energy Fuels 2024, 38, 7758–7772. [Google Scholar] [CrossRef]
  35. Li, Q.; Cao, J.; Liu, Y.; Cheng, Q.; Liu, C. Effect of Dispersed Water on the Paraffin Crystallization and Deposition of Emulsified Waxy Crude Oil via Dissipative Particle Dynamics. J. Mol. Liq. 2021, 343, 117679. [Google Scholar] [CrossRef]
  36. Schäfer, G.; Jaranowski, P. Hamiltonian Formulation of General Relativity and Post-Newtonian Dynamics of Compact Binaries. Living Rev. Relativ. 2024, 27, 2. [Google Scholar] [CrossRef]
  37. Schneider, J.; Fleck, F.; Karimi-Varzaneh, H.A.; Müller-Plathe, F. Simulation of Elastomers by Slip-Spring Dissipative Particle Dynamics. Macromolecules 2021, 54, 5155–5166. [Google Scholar] [CrossRef]
  38. Groot, R.D.; Warren, P.B. Dissipative Particle Dynamics: Bridging the Gap between Atomistic and Mesoscopic Simulation. J. Chem. Phys. 1997, 107, 4423–4435. [Google Scholar] [CrossRef]
  39. Ruiz-Morales, Y.; Mullins, O.C. Coarse-Grained Molecular Simulations to Investigate Asphaltenes at the Oil–Water Interface. Energy Fuels 2015, 29, 1597–1609. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Feller, S.E.; Brooks, B.R.; Pastor, R.W. Computer Simulation of Liquid/Liquid Interfaces. I. Theory and Application to Octane/Water. J. Chem. Phys. 1995, 103, 10252–10266. [Google Scholar] [CrossRef]
  41. Abdulameer, A.G.; Mrah, M.M.; Bazerkan, M.; Al-Haddad, L.A.; Al-Karkhi, M.I. Machine Learning-Driven Power Prediction in Continuous Extrusion of Pure Titanium for Enhanced Structural Resilience under Extreme Loading. Discov. Mater. 2025, 5, 7. [Google Scholar] [CrossRef]
  42. Al-Haddad, L.A.; Łukaszewicz, A.; Majdi, H.S.; Holovatyy, A.; Jaber, A.A.; Al-Karkhi, M.I.; Giernacki, W. Energy Consumption and Efficiency Degradation Predictive Analysis in Unmanned Aerial Vehicle Batteries Using Deep Neural Networks. Adv. Sci. Technol. Res. J. 2025, 19, 21–30. [Google Scholar] [CrossRef]
  43. Al-Haddad, A.A.; Al-Haddad, L.A.; Al-Haddad, S.A.; Jaber, A.A.; Khan, Z.H.; Rehman, H.Z.U. Towards Dental Diagnostic Systems: Synergizing Wavelet Transform with Generative Adversarial Networks for Enhanced Image Data Fusion. Comput. Biol. Med. 2024, 182, 109241. [Google Scholar] [CrossRef]
  44. Al-Haddad, L.A.; Jaber, A.A.; Dhahir, M.K.; Nagim, H.Y.; Algburi, Z.I. Characterization and Prediction of Femtosecond Laser Induced Tracks in Silver-Containing Zinc Phosphate Glass. In Proceedings of the CEUR Workshop Proceedings, Bari, Italy, 14–18 October 2024; Volume 3870, pp. 10–19. [Google Scholar]
  45. Alawee, W.H.; Al-Haddad, L.A.; Basem, A.; Al-Haddad, A.A. A Data Augmentation Approach to Enhance Breast Cancer Detection Using Generative Adversarial and Artificial Neural Networks. Open Eng. 2024, 14, 20240052. [Google Scholar] [CrossRef]
  46. Von Lilienfeld, O.A.; Ramakrishnan, R.; Rupp, M.; Knoll, A. Fourier Series of Atomic Radial Distribution Functions: A Molecular Fingerprint for Machine Learning Models of Quantum Chemical Properties. Int. J. Quantum Chem. 2015, 115, 1084–1093. [Google Scholar] [CrossRef]
  47. Watanabe, N.; Hori, Y.; Sugisawa, H.; Ida, T.; Shoji, M.; Shigeta, Y. A Machine Learning Potential Construction Based on Radial Distribution Function Sampling. J. Comput. Chem. 2024, 45, 2949–2958. [Google Scholar] [CrossRef] [PubMed]
  48. Kelbaliev, G.I.; Safarov, F.F. Study of Interphase Film Thinning in Petroleum Emulsion Separation Processes. Chem. Technol. Fuels Oils 2011, 47, 268–277. [Google Scholar] [CrossRef]
  49. Jiao, B.; Shi, A.; Wang, Q.; Binks, B.P. High-internal-phase Pickering Emulsions Stabilized Solely by Peanut-protein-isolate Microgel Particles with Multiple Potential Applications. Angew. Chem. Int. Ed. 2018, 57, 9274–9278. [Google Scholar] [CrossRef]
  50. Wang, X.; Liu, Y.; Dong, H.; Sun, Q.; Shen, Y.; Ji, R. A Three-Step Model for Submicron W/O Emulsion Formation in a Transitional-Phase Inversion Process. J. Dispers. Sci. Technol. 2016, 37, 1186–1191. [Google Scholar] [CrossRef]
  51. Wang, Z.; Li, N.; Sun, Z.; Wang, X.; Chen, Q.; Liu, W.; Qi, Z.; Wei, L.; Li, B. Molecular Dynamics Study of Droplet Electrocoalescence in the Oil Phase and the Gas Phase. Sep. Purif. Technol. 2021, 278, 119622. [Google Scholar] [CrossRef]
Figure 1. Simplified methodology workflow for multiscale simulation of HIPE formation.
Figure 1. Simplified methodology workflow for multiscale simulation of HIPE formation.
Applsci 15 05422 g001
Figure 2. Coarse-grained beads and molecules and the molecules.
Figure 2. Coarse-grained beads and molecules and the molecules.
Applsci 15 05422 g002
Figure 3. Interfacial activity and aggregation behavior of emulsifiers at oil–water interfaces: (a) Interfacial tension (IFT) of the oil–water system without emulsifier. (bd) Aggregation behavior of emulsifiers AS1, AS2, and AS3 at the toluene (Oil 11)–water interface. (e,f) Aggregation behavior of AS2 and AS3 at the ether (Oil 10)–water interface. The black profile indicates the relative concentration of the emulsifier (ordinate) across the interface, with the abscissa representing the distance normal to the interface in units of Rc (3.23 Å). (g,h) Radial distribution functions (RDF) of emulsifier molecules in the above-mentioned models.
Figure 3. Interfacial activity and aggregation behavior of emulsifiers at oil–water interfaces: (a) Interfacial tension (IFT) of the oil–water system without emulsifier. (bd) Aggregation behavior of emulsifiers AS1, AS2, and AS3 at the toluene (Oil 11)–water interface. (e,f) Aggregation behavior of AS2 and AS3 at the ether (Oil 10)–water interface. The black profile indicates the relative concentration of the emulsifier (ordinate) across the interface, with the abscissa representing the distance normal to the interface in units of Rc (3.23 Å). (g,h) Radial distribution functions (RDF) of emulsifier molecules in the above-mentioned models.
Applsci 15 05422 g003
Figure 4. Morphological evolution of high internal phase emulsions (HIPEs) for various emulsifier–oil combinations: (a1c4) representative porous frameworks formed with amide, ether, alkane, and aromatic oils using AS1, AS2, and AS3 emulsifiers. The inset in (a1) shows non-bonding distances (3.06 Å and 3.45 Å) between oil heteroatoms and AS1. RDF curves between oil and emulsifier are presented between (a1) and (b1).
Figure 4. Morphological evolution of high internal phase emulsions (HIPEs) for various emulsifier–oil combinations: (a1c4) representative porous frameworks formed with amide, ether, alkane, and aromatic oils using AS1, AS2, and AS3 emulsifiers. The inset in (a1) shows non-bonding distances (3.06 Å and 3.45 Å) between oil heteroatoms and AS1. RDF curves between oil and emulsifier are presented between (a1) and (b1).
Applsci 15 05422 g004
Figure 5. Emulsion structures stabilized by AS1, AS2, and AS3 showing differences between stable W/O systems and unstable O/W phase transitions: (af) morphologies of emulsions formed with each emulsifier; (a,c,e) depict successful W/O formation; (b,d,f) show failed emulsions transitioning to O/W structures.
Figure 5. Emulsion structures stabilized by AS1, AS2, and AS3 showing differences between stable W/O systems and unstable O/W phase transitions: (af) morphologies of emulsions formed with each emulsifier; (a,c,e) depict successful W/O formation; (b,d,f) show failed emulsions transitioning to O/W structures.
Applsci 15 05422 g005
Figure 6. Time evolution of W/O emulsions and phase transitions for different emulsifier–oil systems. (a1e1) represent successful stabilization via interfacial emulsifier coverage; (a2e2) depict emulsions undergoing phase transition due to insufficient interfacial diffusion. Time increment is 2 × 10⁵ steps.
Figure 6. Time evolution of W/O emulsions and phase transitions for different emulsifier–oil systems. (a1e1) represent successful stabilization via interfacial emulsifier coverage; (a2e2) depict emulsions undergoing phase transition due to insufficient interfacial diffusion. Time increment is 2 × 10⁵ steps.
Applsci 15 05422 g006
Figure 7. π–π stacking of AS2 at oil–water junctions facilitating emulsion stability. (a) AS2O12 and (b) AS2O14 exhibit aromatic phase bridging through dimers and trimers at the interface.
Figure 7. π–π stacking of AS2 at oil–water junctions facilitating emulsion stability. (a) AS2O12 and (b) AS2O14 exhibit aromatic phase bridging through dimers and trimers at the interface.
Applsci 15 05422 g007
Figure 8. (ac) Coarse-grained models of emulsifiers. (d) AS4O11 and (e) AS4O13 emulsion models. The emulsion models are shown by 2 × 2 × 2 cells for better observation. Inserted in the (d,e) was the local morphology of the emulsifier (both the oil and the water were hidden).
Figure 8. (ac) Coarse-grained models of emulsifiers. (d) AS4O11 and (e) AS4O13 emulsion models. The emulsion models are shown by 2 × 2 × 2 cells for better observation. Inserted in the (d,e) was the local morphology of the emulsifier (both the oil and the water were hidden).
Applsci 15 05422 g008
Figure 9. Energy of the orbitals: (a) HOMO; (b) LUMO; (c) energy gap between HOMO and LUMO; (d) dipole moment; (e) cohesive energy density of the components.
Figure 9. Energy of the orbitals: (a) HOMO; (b) LUMO; (c) energy gap between HOMO and LUMO; (d) dipole moment; (e) cohesive energy density of the components.
Applsci 15 05422 g009
Figure 10. All-atom molecular structures of emulsifiers and oils used for DFT calculations.
Figure 10. All-atom molecular structures of emulsifiers and oils used for DFT calculations.
Applsci 15 05422 g010
Figure 11. HOMO and LUMO orbital distribution for emulsifier molecules showing localization within the PAC plane.
Figure 11. HOMO and LUMO orbital distribution for emulsifier molecules showing localization within the PAC plane.
Applsci 15 05422 g011
Table 1. State-of-the-art literature on the study topic.
Table 1. State-of-the-art literature on the study topic.
Ref.Application/MethodKey FindingLimitation
[25]DPD of asphaltenes and sodium naphthenatesOxygenated asphaltenes form more stable interfacial filmsFocused only on binary mixtures without complex multicomponent emulsions
[27]Superhydrophobic foam for emulsion separationGraphene/polystyrene foams effectively separate W/O emulsionsExperimental only; no molecular-level understanding
[28]DPD + experimental on Gemini surfactantsGemini surfactants stabilize emulsions more effectivelyConcentration effect overlaps structural interpretation
[29]DPD of copolymer–crude oil–water emulsionsLonger co-polymer chains improve coalescence controlEmulsion system limited to single polymer–oil configuration
[26]DPD on interfacial behavior in ultra-deep reservoirsSurfactants and crude components compete for interfacial adsorptionNo discussion of performance under varying water contents
[30]DPD study of polyether demulsifiersPBO-based demulsifiers enhance coalescence via network formationMechanism valid only for ultra-heavy crude and certain polymer types
[31]DPD on anionic/cationic mixed surfactantsMixed surfactants reduce IFT (interfacial tension) and enhance emulsion stabilitySpacer group effect needs more systematic quantification
[32]Integrated MD–DPD for salt/surfactant effectsSalinity and temperature influence IFT and interfacial structureRadius of gyration and χ parameter treated in ideal conditions
[33]DPD on Gemini surfactants with variable spacersShorter spacers yield better interfacial activity and lower sedimentationAging behavior not fully validated experimentally
[23]DPD of asphaltenes + surfactantsSynergistic adsorption reduces IFT, forming tight filmsFocus limited to heavy oil and static systems
[34]DPD on emulsion conformation at varying water contentsAdsorption energy and Rg changes affect emulsion stabilityNo extension to HIPEs
[35]DPD of wax crystallization in emulsified waxy crudeWater cut alters paraffin nucleation pathwaysNo link to interfacial film or surfactant activity
[24]DPD + experiment on S/P emulsificationStable S/P systems reduce IFT and improve oil recoveryInterfacial behavior not generalized across emulsifier types
Table 2. Conservative force parameters of the beads employed in the present study. Bead types are defined as: W (water, representing three water molecules), C (polyaromatic core of emulsifiers), B (aliphatic side chains of emulsifiers), N (nitrogen-containing heteroatom side chains), EO (ether functional groups in oils 8–10), and T (aromatic groups in oils 11–14, e.g., toluene).
Table 2. Conservative force parameters of the beads employed in the present study. Bead types are defined as: W (water, representing three water molecules), C (polyaromatic core of emulsifiers), B (aliphatic side chains of emulsifiers), N (nitrogen-containing heteroatom side chains), EO (ether functional groups in oils 8–10), and T (aromatic groups in oils 11–14, e.g., toluene).
aijWCBNEOT
W78
C129.7878
B131.0279.5578
N88.95101.9495.2478
EO101.1282.6979.5583.7978
T138.9579.587897.3479.8478
Table 3. Simulation execution parameters and conditions.
Table 3. Simulation execution parameters and conditions.
ParameterValue
Simulation ensembleNVT
Temperature65 °C
Timestep0.005tc (≈15 fs)
Total steps1,000,000
Simulation box size 100 × 100 × 100 R c 3
Emulsifier: Oil mass ratio1:1
Water content (volume basis)70% and 80%
Initial configurationRandom packing
IFT slab configurationOil/Water/Oil layers (25:50:25 Rc)
Output analysesMorphology, RDF, IFT, snapshots
Table 4. Simulation modeling assumptions and physical boundaries.
Table 4. Simulation modeling assumptions and physical boundaries.
AssumptionDescription
Coarse-graining levelEach bead represents 2–3 molecules, enabling mesoscale behavior capture
No reactive chemistryDPD does not account for chemical bond formation or reaction kinetics
Isothermal systemConstant temperature (65 °C) maintained in all simulations
Constant mass ratio (Emulsifier/Oil)Held at 1:1 for all systems
Initial molecular distributionRandomized distribution assumed to reflect emulsification via shearing
Fixed simulation durationOne million steps assumed sufficient for equilibrium morphology observation
Coarse-graining levelEach bead represents 2–3 molecules, enabling mesoscale behavior capture
No reactive chemistryDPD does not account for chemical bond formation or reaction kinetics
Isothermal systemConstant temperature (65 °C) maintained in all simulations
Constant mass ratio (Emulsifier/Oil)Held at 1:1 for all systems
Table 5. Simulated vs. literature IFT values for selected oil–water systems.
Table 5. Simulated vs. literature IFT values for selected oil–water systems.
Oil TypeFunctional Group TypeSimulated IFT Trend
(This Study)
Reported IFT Range
(Literature)
Ref.Consistency
Oils 5–7Amides (polar)~38–41 mN/m35–40 mN/m[23,25]High
Oils 8–10Ethers (less polar)~43–46 mN/m~40–45 mN/m[23,32]Moderate
Oils 1–4Alkanes (non-polar)~50–52 mN/m>50 mN/m[23,26]High
Oils 11–14Aromatics (non-polar)~51–54 mN/m>50 mN/m[23,25]High
Table 6. Simulation results for the emulsion model systems at 70% water content. A ‘+’ indicates a positive result (W/O emulsion formation) in at least one of the three repeated runs.
Table 6. Simulation results for the emulsion model systems at 70% water content. A ‘+’ indicates a positive result (W/O emulsion formation) in at least one of the three repeated runs.
Water Content Oil1234567891011121314Counts
Emulsifier
70%AS1 + ++ 3
AS2++++++++++ + +12
AS3++ +++++++ 9
Counts23123222221101
Table 7. Simulation results of the emulsion model with 70% water content. The ‘+’ in the table meant W/O emulsion formed in the model.
Table 7. Simulation results of the emulsion model with 70% water content. The ‘+’ in the table meant W/O emulsion formed in the model.
Oil1234567891011121314Counts
Emulsifier
AS4++++++++++++++14
AS5++++++++++++++14
AS6 ++++++ 6
Table 8. DFT-computed HOMO and LUMO energies, HOMO–LUMO gap, and dipole moments of emulsifiers and oils used in this study.
Table 8. DFT-computed HOMO and LUMO energies, HOMO–LUMO gap, and dipole moments of emulsifiers and oils used in this study.
SampleEHOMO (Ha)ELUMO (Ha)ELUMO-HOMO (Ha)ELUMO-HOMO (eV)
AS1−0.183 −0.100 0.082 2.239
AS2−0.170 −0.102 0.068 1.853
AS3−0.142 −0.113 0.029 0.791
AS4−0.158 −0.099 0.059 1.610
AS5−0.164 −0.102 0.062 1.679
AS6−0.139 −0.080 0.060 1.626
Oil1−0.260 0.067 0.327 8.901
Oil2−0.252 0.066 0.318 8.649
Oil3−0.248 0.066 0.313 8.528
Oil4−0.245 0.066 0.310 8.447
Oil5−0.192 0.024 0.216 5.876
Oil6−0.191 0.023 0.215 5.844
Oil7−0.189 0.022 0.212 5.756
Oil8−0.203 0.063 0.266 7.250
Oil9−0.202 0.061 0.263 7.154
Oil10−0.202 0.061 0.263 7.166
Oil11−0.210 −0.028 0.182 4.962
Oil12−0.128 −0.050 0.077 2.107
Oil13−0.184 −0.058 0.126 3.426
Oil14−0.135 −0.112 0.023 0.623
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

Shi, P.; Ogail, M.H.O.; Feng, X.; Fang, S.; Duan, M.; Pu, W.; Liu, R. A Dissipative Particle Dynamics Study on the Formation of the Water-In-Petroleum Emulsion: The Contribution of the Oil. Appl. Sci. 2025, 15, 5422. https://doi.org/10.3390/app15105422

AMA Style

Shi P, Ogail MHO, Feng X, Fang S, Duan M, Pu W, Liu R. A Dissipative Particle Dynamics Study on the Formation of the Water-In-Petroleum Emulsion: The Contribution of the Oil. Applied Sciences. 2025; 15(10):5422. https://doi.org/10.3390/app15105422

Chicago/Turabian Style

Shi, Peng, Murtaja Hamid Oudah Ogail, Xinxin Feng, Shenwen Fang, Ming Duan, Wanfen Pu, and Rui Liu. 2025. "A Dissipative Particle Dynamics Study on the Formation of the Water-In-Petroleum Emulsion: The Contribution of the Oil" Applied Sciences 15, no. 10: 5422. https://doi.org/10.3390/app15105422

APA Style

Shi, P., Ogail, M. H. O., Feng, X., Fang, S., Duan, M., Pu, W., & Liu, R. (2025). A Dissipative Particle Dynamics Study on the Formation of the Water-In-Petroleum Emulsion: The Contribution of the Oil. Applied Sciences, 15(10), 5422. https://doi.org/10.3390/app15105422

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