Donggang Caoa and
Gilad Yossifon*ab
aSchool of Mechanical Engineering, Tel-Aviv University, Tel-Aviv, 69978, Israel. E-mail: gyossifon@tauex.tau.ac.il
bDepartment of Biomedical Engineering, Tel-Aviv University, Tel-Aviv, 69978, Israel
First published on 24th July 2025
Active (self-propelling) particles have emerged as innovative microscale tools in the field of single cell analysis with the advantages of being untethered, remotely controlled, hybrid powered, with sub-cellular precision. This study investigates the dielectrophoretic (DEP) response and electro-mechanical deformation of cell nuclei interacting with active metallo-dielectric Janus Particles (JPs) under an externally applied electric field. An “equivalent droplet” two-phase model is employed to simulate the bioparticle, coupling the Navier–Stokes equations with the phase field model to capture fluid motion and interface dynamics. Good qualitative agreement is obtained among experimental, analytical, and numerical results. The findings reveal a nonlinear relationship between nucleus deformation and its surface coverage of the JP with respect to the applied voltage. The overall coverage ratio of the JP's dielectric hemisphere increases with voltage as the positive DEP force on the dielectric side strengthens, exhibiting a maximum at a certain voltage. The strong correlation between nucleus flexibility and JP surface coverage suggests that the JP coverage ratio could serve as a biomechanical marker for nucleus deformability, providing a novel method for in situ evaluation of nucleus mechanics.
The asymmetric design of JPs allows them to function as mobile microelectrodes under an applied electric field, generating localized field gradients that drive DEP forces.26 These forces can trap and deform soft bioparticles, providing a dual mechanism for both manipulation and mechanical assessment.27 The electro-deformation of trapped cells and nuclei under DEP forces offers valuable insights into cellular mechanics, which are critical for understanding cellular health, disease mechanisms, and mechanical phenotyping—that is, the assessment of structural integrity and mechanical properties in both healthy and diseased cells.28,29 Notably, the nucleus, as the largest and mechanically dominant organelle in most cells, plays a central role in this context. Its deformability has been shown to correlate with metastatic potential, chromatin organization, and a range of disease states such as laminopathies and cancers.30,31 These findings highlight the potential of electric-field-based probing to reveal biologically meaningful information about the mechanical behavior of soft bioparticles such as nuclei and whole cells.
While experimental studies have provided valuable preliminary insights into JP-bioparticle interactions, intriguing phenomena such as JP insertion and engulfment at cell membranes have been observed.32,33 Our recent work has demonstrated that active, polarizable metallo-dielectric JPs effectively function as mobile microelectrodes, generating the necessary field gradients for efficient DEP manipulation.34 Furthermore, we have highlighted the strong correlation between DEP-induced electro-deformation of cell nuclei and applied field strength, underscoring the potential of JPs as active carriers for mechanical probing of subcellular components.35 However, to the best of our knowledge, no systematic simulation studies have been conducted to investigate these complex JP-bioparticle interactions, particularly in DEP-based applications. This gap hinders the comprehensive interpretation of experimental data and the optimization of JP-bioparticle interactions for practical biomedical use. To address this gap, the present study employs numerical simulations using COMSOL Multiphysics 6.236 to analyse the DEP-driven interactions between JPs and deformable particles, with a particular focus on cell nuclei. Both the Janus particle and nucleus are modeled as spheres with 10 μm diameters. To simulate nucleus deformation under varying electric fields, we adopt an ‘equivalent droplet’ two-phase model that combines the properties of the cytoplasm and membrane (thickness ∼10 nm)37–41 (see Section 2). This approach couples the Navier–Stokes equations with a phase-field method42–44 to capture fluid dynamics and interface behavior, accounting for the interplay of hydrodynamic and electric forces driving nucleus motion and deformation.
The present study systematically investigates how particle composition, applied voltage, and nucleus stiffness influence dielectrophoretic response and electro-deformation. By providing a comprehensive framework for understanding these interactions, this research enhances fundamental physical insights, predictive modelling, and the design of DEP-based biomanipulation strategies, with broad implications for active particle-based biomedical applications.
![]() | (1) |
To track the interface during the transient evolution of two-phase laminar flow, the phase-field method is employed for its efficiency in accurately capturing interface deformation under complex force interactions:45,46
![]() | (2) |
The density and viscosity are automatically calculated from the phase-field variable ϕ:
![]() | (3) |
The surface tension force is calculated as:
![]() | (4) |
The electric force is given by the divergence of the Maxwell stress tensor by the equation below:47,48
![]() | (5) |
![]() | (6) |
Furthermore, the current conservation equation can be solved to determine E:
![]() | (7) |
Directly solving these equations in the physical time domain to simulate nucleus deformation at 1 MHz requires prohibitively small time steps and incurs high computational costs. To balance efficiency and accuracy, the electrical problem is solved in the frequency domain, and a cycle-averaged electric force is applied in the time domain.49 This is an appropriate approximation given the significant disparity in timescales between fluid motion (∼ms) and electrical excitation (∼MHz). Consequently, the governing equation for the electric field in the frequency domain becomes:
![]() | (8) |
![]() | (9) |
The cell nucleus is modelled as a homogenized “equivalent droplet” with effective electrical properties that capture the combined influence of the nucleus membrane and nucleus cytoplasm, as illustrated in Fig. S1 (ESI†). The characteristic dielectric properties of the nucleus membrane and cytoplasm, obtained from the literature, are summarized in Table S1 (ESI†). The DEP response across a broad frequency range, evaluated using the Clausius–Mossotti factor, is presented in Fig. S2 (ESI†). Table S2 (ESI†) provides the physical properties of the nucleus and the surrounding fluid, selected to satisfy the experimental condition that a positive DEP (pDEP) force arises at the experimentally investigated frequency of 1 MHz.
The governing equations are solved with the following boundary conditions: the upper and lower walls act as electrodes with applied voltages, while the left and right channel edges are set to electric insulation. Since our primary focus is on simulating the high frequency domain (∼1 MHz), consistent with the preliminary experiments,35 we neglect induced-charge electro-osmosis (ICEO) effects and the effective Helmholtz-Smoluchowski slip boundary condition on the metal hemisphere. This simplification is justified as the operating frequency is significantly higher than the RC frequency associated with the formation of the induced electrical double layer (EDL) (fRC = 1/2πτ = 1.3 kHz refers to the characteristic charging frequency, where τ = λR/D is the induced charge relaxation time; R = 5 μm is the radius of JP; is the Debye length; σ = 6 × 10−4 S m−1 is the solution conductivity; εs = 78ε0 is the solution permittivity and ε0 = 8.854187817 × 10−12 F m−1 is the vacuum permittivity; D = 2 × 10−9 m2 s−1 is the diffusion coefficient of the ionic species51). Accordingly, the JP is modeled by assigning distinct boundary conditions to its two hemispheres: a floating electrode condition for the metallic (gold) hemisphere and an insulating condition for the dielectric (polystyrene) hemisphere.34 Validation results in the ESI† confirm that the applied boundary conditions effectively capture JP–nucleus interactions at 1 MHz (see Fig. S3 and S4, ESI†). Additionally, Fig. S5 (ESI†) presents the Clausius–Mossotti factor and the corresponding JP–nucleus interaction, derived from the parameters in Table S2 (ESI†) and simulations using the floating boundary condition. These results indicate that, as frequency increases beyond 1 MHz—where positive dielectrophoresis (pDEP) is observed experimentally—a transition to negative DEP (nDEP) occurs, leading the nucleus to migrate toward regions of lower electric field strength. The strong agreement between analytical predictions and numerical simulations further supports the validity and robustness of the proposed modeling approach.
In addition, no-slip boundary conditions are imposed on all solid walls, which is justified at the examined frequency (∼1 MHz), well above the RC frequency fRC = 1/2πτ = 1.3 kHz. A time-dependent solver is used to capture the dynamic JP–nucleus interaction, with the focus placed on the quasi-steady-state results, as illustrated in Fig. S6 (ESI†). To accurately resolve the complex interactions in the system, second-order numerical schemes are employed in conjunction with a locally refined mesh (see Fig. S7, ESI†), concentrating computational effort in regions with steep field gradients. Four computational grids with varying cell counts (see Table S3, ESI†) are used to assess numerical errors arising from limited spatial resolution. The results confirm that a fine mesh with approximately 30000 cells provides accurate and grid-independent solutions. Further details on the grid-independence validation are provided in Fig. S8 of the ESI.†
![]() | ||
Fig. 1 Experimental and numerical system setups and overview of JP-nucleus interactions: (A) schematic of the experimental setup35 (Copyright 2020 John Wiley and Sons). (B) Schematic representation of the simulation model. (C) Time evolution of JP-nucleus interaction observed experimentally, showing nucleus attraction and deformation onto the JP (inset: side-view schematic)35 (Copyright 2020 John Wiley and Sons). (D) Corresponding numerical simulation snapshots. (E) Normalized electric field intensity distributions at different normalized (scaled by the characteristic DEP migration time ts = 3.1 ms at Vpp = 10 V, f = 1 MHz, see ESI,† for more details) time points (white line indicates the nucleus surface contour). |
When exposed to an electric field, the JP attracts and deforms the nucleus onto its surface, enabling it to carry and transport the loaded nucleus35 (Fig. 1C, the reduced contrast of the nucleus in the final experimental frame is due to a slight displacement in the z-direction during interaction, causing the nucleus to move partially out of the focal plane). Fig. 1D shows numerical simulation snapshots (a cross-sectional view of plane A–A in Fig. 3A) illustrating the JP-nucleus interaction over time. The nucleus is attracted to the JP by DEP forces, gradually deforming and migrating until it reaches a steady-state shape, settling on the JP's surface and covering parts of both the gold and polystyrene hemispheres. The initial spacing between the JP and the nucleus was set arbitrarily at 2.5 μm. This behaviour aligns qualitatively with experimental observations (see Movie S1 in the ESI†) though there is an orientation mismatch.
Fig. 1E presents the time-evolving normalized electric field distribution, further supporting the experimental findings. Both nucleus and JP undergo polarization in the electric field, with the JP inducing perturbations that generate distinct dielectrophoretic (DEP) potential wells, labelled ①–⑥ in the first panel of Fig. 1E. Local maxima in the electric field at locations ① and ⑤ indicate regions where positive DEP (pDEP) trapping of bioparticles is likely, while a weaker pDEP well appears on the polystyrene side at location ④. Conversely, locations ②, ③, and ⑥ correspond to negative DEP (nDEP) trapping sites due to local electric field minima. The electric field asymmetry arises from the Janus particle's intrinsic surface asymmetric dielectric properties (gold vs. polystyrene, where the former is ideally polarizable while the latter is insulating) in contrast with a homogenous particle where symmetry does exist. In response to a positive dielectrophoresis (pDEP) force,35 the nucleus, behaving as a homogeneous dielectric droplet, moves toward the high-intensity potential well at location ①. Upon contact with the JP, the electric field intensity at the JP-nucleus interface (location ①) gradually decreases until the system reaches a steady-state deformed nucleus configuration.
![]() | ||
Fig. 2 Normalized electric field distributions (scaled by 106 V m−1) are shown for nuclei positioned at various locations around (A) a conductive particle, (B) a dielectric particle, and (C) a Janus particle at different normalized time points (scaled by the characteristic DEP migration time ts = 3.1 ms at Vpp = 10 V, f = 1 MHz, see ESI,† for more details). |
The electric field distributions in Fig. 2A and B exhibit bilateral (mirror) symmetry about the vertical axis due to the uniform material properties of the conductive (Au) and dielectric (PS) particles, respectively. This symmetry is preserved from the initial state to the quasi-steady state, resulting in symmetric field lines and DEP forces acting radially and evenly around the particle. In contrast, the Janus particle in Fig. 2C introduces asymmetry due to the distinct material properties of its Au and PS hemispheres. This breaks the mirror symmetry of the field distribution, concentrating electric field gradients near the metallo-dielectric interface. As a result, the nucleus experiences a directional DEP force, leading to asymmetric migration and deformation. The comparison underscores the role of particle symmetry in governing field gradients and the resulting mechanical response of nearby soft particles.
![]() | ||
Fig. 3 JP-Nucleus interaction under varying electric fields at 1 MHz: (A) Comparison of 3D and 2D simulation results. (B) Experimental images of the nucleus (red) and Janus particle (JP, dotted yellow outline) reconstructed using Imaris 5.052 at the central plane. The green region represents the polystyrene (PS) hemisphere of the JP, while the shaded circular sector denotes the PS portion covered by the nucleus. The red color of the nucleus in the experimental image highlights its shape, as extracted from fluorescence microscopy35 (Copyright 2020 John Wiley and Sons). (C) Numerical images of the nucleus (white) and JP, with the dashed circular sector indicating the PS hemisphere region covered by the nucleus. (D) Simulated electric field distributions during JP-nucleus interactions at different applied voltages. (E) Experimental and simulated dielectrophoretic coverage ratio of the PS hemisphere versus applied voltage. (F) Total coverage ratio (including both PS and Au hemispheres) versus applied voltage. (G) Deformation ratio of the nucleus versus applied voltage. |
Fig. 3D illustrates the electric field magnitude distribution around the JP and nucleus at different voltages, demonstrating how the field evolves as the applied voltage increases. As described in Section 2.1, the nucleus experiences pDEP forces, migrating toward regions of higher electric field intensity. Initially, the strongest field intensity at location ① serves as the primary trapping site. However, as the nucleus is captured, this pDEP well weakens, and other wells, particularly on the PS side (location ④), become more influential. The increasing field intensity at ④ at higher voltages enhances the nucleus's attraction to the polystyrene hemisphere.
The quantitative analysis of the dielectrophoretic coverage ratio (Fig. 3E), defined as the fraction of the JP's PS hemisphere covered by the nucleus (η1 = (θ1/π) × 100%, where θ1 is the angular coverage), further supports the validity of the 2D simulations. Despite the limited number of 3D data points due to computational constraints, both 2D and 3D simulations show good agreement with experimental results, using surface tension as the sole fitting parameter to the differing deformability of MDA-MB-231 and MCF-7 nuclei.
Generally, the coverage ratio increases with the applied voltage up to approximately 14 V, after which a slight decline is observed. This non-monotonic behaviour may result from competing mechanisms: on one hand, increased electric field intensity enhances pDEP-driven attraction toward location ④, promoting greater nucleus coverage of the dielectric side; on the other hand, elongation of the nucleus along the field lines reduces its effective contact area with the JP, thereby decreasing θ1. This elongation also influences the total coverage ratio (η2 = (θ2/2π) × 100%), which declines with increased voltages (Fig. 3F). Although the 2D simulations qualitatively match the 3D results on the plane of interest in terms of coverage ratios, they do not fully replicate the deformation ratio of the 3D system (Fig. 3G). The deformation ratio, defined as the percentage increase in the nucleus's surface area (φ = (ΔS/S) × 100%), quantifies how much the nucleus stretches under the electric field. Since 2D simulations inherently capture deformation in only two directions, the deformation ratios are different from those from 3D simulations. Hence, while the 2D simulations have quantitative limitations in predicting deformation ratios, they provide valuable insights into the key mechanisms of JP-nucleus interaction.
To quantify nucleus deformability, the transmigration-based deformability assay (TDA) using Transwell inserts with 5 μm and 8 μm pores was employed in ref. 35. In this assay, MDA-MB-231 cells – known for their high deformability - exhibited increased transmigration, with over 180 cells passing through both pore sizes. This high rate of passage suggests that their nuclei can easily deform and squeeze through narrow openings. In contrast, MCF-7 cells, with lower deformability, depict limited transmigration, with only about 50 cells passing through the 8 μm pores and very few through the 5 μm pores, highlighting the stiffness of their nuclei (Fig. 4A).
![]() | ||
Fig. 4 Influence of nucleus stiffness on JP-nucleus interaction at 1 MHz: (A) schematic of the transmigration-based deformability assay (TDA) and results from ref. 35 (Copyright 2020 John Wiley and Sons) demonstrating the deformability of MDA-MB-231 and MCF-7 cells. (B) Comparison of experimental35 (Copyright 2020 John Wiley and Sons) and numerical results for JP-nucleus interactions with varying nucleus stiffness (MDA-MB-231 and MCF-7) at 4 V and 18 V. (C) Simulated dielectrophoretic coverage ratio as a function of applied voltage for different nucleus stiffnesses (represented by equivalent surface tension, Γ); (D) simulated dielectrophoretic coverage ratio as a function of equivalent surface tension (Γ) at the applied voltage of 18 V. |
Correspondingly, simulations have been conducted by modelling the nucleus deformability through an effective surface tension parameter (γ). Fig. 4B presents experimental images and simulation snapshots comparing interactions of the JP with nuclei of varying deformability (represented by the normalized equivalent surface tension, Γ = γ/γ0, with γ0 = 0.0002 N m−1 (ref. 54)) at 4 V and 18 V. For the highly deformable MDA-MB-231 nuclei, simulations with lower Γ values (indicating higher deformability) show increased coverage of the JP's polystyrene side. Conversely, stiffer MCF-7 nuclei exhibit reduced coverage, in qualitative agreement with experimental observations. Fig. 4C quantifies the dielectrophoretic coverage ratio (η1) as a function of applied voltage for nuclei with varying surface tension values. Across all deformability levels, η1 increases with voltage, but the rate of increase is strongly influenced by nucleus stiffness. Softer nuclei (lower Γ values, e.g., MDA-MB-231) exhibit a steep rise in η1, indicating their greater ability to conform to the JP's surface under stronger electric fields. Conversely, stiffer nuclei (higher Γ values, e.g., MCF-7) show a more gradual and limited increase in η1, reflecting their resistance to deformation. All curves follow a similar trend: an initial rapid rise in η1, followed by a reduced growth rate beyond a critical transition voltage. This transition voltage shifts higher for stiffer nuclei, highlighting the interplay between deformability and electric field strength.
Fig. 4D shows an inverse correlation between η1 and Γ at 18 V and 1 MHz, indicating that stiffer nuclei (higher Γ) exhibit lower dielectrophoretic coverage due to reduced deformability. This suggests that JP coverage ratio may serve as a biomechanical marker of nucleus stiffness. Unlike the population-based TDA assay, which uses mechanical squeezing for indirect, high-throughput screening, our electro-deformation approach employs electrostatic forces for direct, single cell resolution. While TDA is simple and scalable, EDA offers quantitative, real-time insight into single-cell mechanics. Together, they provide complementary perspectives on nuclear deformability.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5sm00462d |
This journal is © The Royal Society of Chemistry 2025 |