Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Probability and kinetics of rupture and electrofusion in giant unilamellar vesicles under various frequencies of direct current pulses

  • Md. Tariqul Islam Bhuiyan,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Physics, Bangladesh University of Engineering and Technology, Dhaka, Bangladesh

  • Mohammad Abu Sayem Karal ,

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    asayem221@phy.buet.ac.bd

    Affiliation Department of Physics, Bangladesh University of Engineering and Technology, Dhaka, Bangladesh

  • Urbi Shyamolima Orchi,

    Roles Data curation, Formal analysis, Investigation, Methodology, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Physics, Bangladesh University of Engineering and Technology, Dhaka, Bangladesh

  • Nazia Ahmed,

    Roles Data curation, Formal analysis, Investigation, Methodology, Validation, Writing – original draft, Writing – review & editing

    Affiliation Department of Physics, Bangladesh University of Engineering and Technology, Dhaka, Bangladesh

  • Md. Moniruzzaman,

    Roles Investigation, Validation, Writing – original draft, Writing – review & editing

    Affiliation Department of Physics, Bangladesh University of Engineering and Technology, Dhaka, Bangladesh

  • Md. Kabir Ahamed,

    Roles Conceptualization, Data curation, Investigation, Validation, Writing – original draft, Writing – review & editing

    Affiliation Radiation, Transport and Waste Safety Division, Bangladesh Atomic Energy Regulatory Authority, Agargaon, Dhaka, Bangladesh

  • Md. Masum Billah

    Roles Validation, Writing – original draft, Writing – review & editing

    Affiliation Department of Physics, Jashore University of Science and Technology, Jashore, Bangladesh

Abstract

Irreversible electroporation induces permanent permeabilization of lipid membranes of vesicles, resulting in vesicle rupture upon the application of a pulsed electric field. Electrofusion is a phenomenon wherein neighboring vesicles can be induced to fuse by exposing them to a pulsed electric field. We focus how the frequency of direct current (DC) pulses of electric field impacts rupture and electrofusion in cell-sized giant unilamellar vesicles (GUVs) prepared in a physiological buffer. The average time, probability, and kinetics of rupture and electrofusion in GUVs have been explored at frequency 500, 800, 1050, and 1250 Hz. The average time of rupture of many ‘single GUVs’ decreases with the increase in frequency, whereas electrofusion shows the opposite trend. At 500 Hz, the rupture probability stands at 0.45 ± 0.02, while the electrofusion probability is 0.71 ± 0.01. However, at 1250 Hz, the rupture probability increases to 0.69 ± 0.03, whereas the electrofusion probability decreases to 0.46 ± 0.03. Furthermore, when considering kinetics, at 500 Hz, the rate constant of rupture is (0.8 ± 0.1)×10−2 s-1, and the rate constant of fusion is (2.4 ± 0.1)×10−2 s-1. In contrast, at 1250 Hz, the rate constant of rupture is (2.3 ± 0.8)×10−2 s-1, and the rate constant of electrofusion is (1.0 ± 0.1)×10−2 s-1. These results are discussed by considering the electrical model of the lipid bilayer and the energy barrier of a prepore.

1. Introduction

Biomembranes exhibit distinct electrical properties that can be effectively manipulated through the application of electric field. At weak direct current (DC) pulses of electric field, the membrane can deform under the influence of electric stresses [1]. At strong pulses, transient pores form in the membranes (i.e., electroporation), which dramatically increases the membrane permeability. Electroporation technology is utilized for the transmembrane transport of drugs, gene, proteins, and other molecules in the fields of medicine, food processing, and specific environmental contexts [25]. In the irreversible electroporation (IRE) technique, a series of high electric field (in V/cm) DC pulses are applied to ablate cancer cells/tumors and rupture vesicles in a non-thermal process [6, 7]. A related phenomenon of electroporation called electrofusion can be employed to merge two cells or vesicles that are in close proximity, enabling the creation of hybrid cell-cell, vesicle-vesicle, or cell-vesicle fusion products [8, 9]. In a clinical study, it was observed that elevating the frequency of AC field exerts an inhibitory effect on the growth of cancer cells (e.g., HeLa), attributed to the alteration of membrane potential at distinct frequencies [10]. The effect of frequency of the AC field on the distribution of electrical stresses on the cell surface has been investigated and found that frequency has a significant effect on cell deformation and membrane permeability [11]. Very recently, the dynamics of cell death (e.g., Chinese hamster ovary cells (CHO), mouse melanoma cells (B16F1), and rat heart myoblast cells (H9c2)) have been observed by changes in metabolic activity and membrane integrity at high electroporation intensities [12]. This suggests that cell death resulting from electroporation is not an immediate all-or-nothing response but rather a dynamic process that occurs over an extended period of time.

Most of the electroporation and electrofusion in lipid vesicles have been performed by varying the electric field of DC and some cases AC with a constant frequency [8, 13, 14]. Previously, we have investigated the IRE in giant unilamellar vesicles (GUVs) by varying the surface charge and salt concentration in a physiological buffer [15], cholesterol concentration in the lipid membranes [16], concentration gradient between the exterior and interior of GUVs (e.g., osmotic effect), and sugar concentration [17, 18]. In addition, the electrotension (electric field induces electrotension) dependent kinetics of vesicle rupture has been investigated in which the rate constant of rupture formation increases with the applied tension [19]. Basically, the probability of rupture and kinetics of rupture are two important parameters found from the statistical analysis of many ‘single GUVs’ which helps to understand the elementary process of pore formation in the lipid bilayer due to the electric and mechanical tension [2023]. The theoretical fitting to the experimental data on the tension (electric or mechanical) dependent rate constant of pore formation gives the opportunity to find the most significant intrinsic parameter of the lipid membrane called pore edge tension, which is responsible for closing the pore [22, 2426]. Moreover, the direct evidence of transient pores was visualized in the bilayer [27] and quantified the potential energy barrier associated with lipid bilayer electroporation [28, 29].

Although there are several reports on the DC pulses induced poration and fusion in GUVs, almost all the investigations are performed by changing the electric field with a single frequency. Moreover, the probability and the kinetics of electroporation and electrofusion is rarely investigated under various frequencies of DC pulses. As the vesicle/cell membranes behave as a parallel resistor-capacitor circuit, it is indispensable to investigate the responses of GUVs under a range of frequencies. Through the investigation of the impacts of frequency on GUVs, the optimum frequency that is effective in either electroporation or facilitating the electrofusion has been obtained. The results of several independent experiments comprising many ‘single GUVs’ are analyzed statistically for determining the average time, probability, and the rate constant of rupture and electrofusion.

2. Materials and methods

2.1 Chemicals and reagents

1,2-dioleoyl-sn-glycero-3-phospho-(1′-rac-glycerol) (sodium salt) (DOPG) and 1, 2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) were purchased from Avanti Polar Lipids Inc. (Alabaster, AL). 1,4-Piperazinediethanesulfonic acid (PIPES), bovine serum albumin (BSA), O,O´-Bis (2-aminoethyl) ethyleneglycol-N,N,,-tetraacetic acid (EGTA), sodium chloride, glucose, and sucrose were purchased from Sigma-Aldrich (Germany).

2.2 Preparation of GUVs

DOPG and DOPC lipids were used to prepare the DOPG/DOPC (40/60)-GUVs [here 40/60 indicates the molar ratio] by the natural swelling method [30] in a physiological buffer (10 mM PIPES, 150 mM NaCl, pH 7.0, 1mM EGTA). At first a mixture of 1 mM DOPG and DOPC of total amount 200 μL was taken into a glass vial and dried with a gentle flow of nitrogen gas for producing a thin, homogeneous lipid film. Then the vial was placed in a vacuum desiccator for 12 h for removing the remaining chloroform. The lipid film was pre-hydrated for 8 min at 45°C by adding an amount of 20 μL MilliQ water into the glass vial. After that the sample was incubated for 3 h at 37°C with 1 mL buffer containing 0.10 M sucrose for producing the suspension of GUVs. To separate the suspension of GUVs from lipid aggregates and multilamellar vesicles, the samples were subjected to centrifugation using 13000×g (here g is the acceleration due to gravity) at 20°C for 20 min using a centrifuge machine (NF 800R Centrifuge, Nuve, Turkey). Subsequently, the supernatant of the suspension was purified using the membrane filtering method in which the purified GUVs contained 0.10 M glucose in buffer as an external solution [31, 32]. An amount of 300 μL purified suspension of GUVs was transferred into a U-shaped hand-made microchamber (see Fig 1B) which was coated with 0.10% (w/v) BSA, dissolved in 0.10 M glucose containing buffer. An inverted phase contrast microscope (Olympus IX-73, Japan) with a 20× objective (numerical aperture: 0.45) was used to observe the GUVs at 25 ± 1°C (Tokai Hit Thermo Plate, Japan). The recording of GUVs images was done using a charged coupled camera of 25 frames per second (fps) (Model: DP22, Olympus) connected to the microscope.

thumbnail
Fig 1. Experimental setup to apply the electric field on the GUVs.

(A) Experimental set up of a microchamber at the stage of an inverted phase contrast microscope. The microchamber comprised a glass slide, a silicon rubber spacer, and a cover slip, forming a U-shaped structure. (B) Schematic representation of the U-shaped microchamber contained GUVs suspension and gold-coated electrode with proper configuration. (C) Schematic of a ‘single GUV’ exposed to the DC pulses of electric field. The direction of electric field (E) is shown in the figure. (D) The equivalent circuit of the lipid membrane, containing a resistor (Rm) and a capacitor (Cm) in parallel. (E) A DC pulse of electric field. (F) Illustration of DC pulses (rectangular pulses) with ON time (i.e., 200 μs) and OFF time (variable of time, i.e., 1800, 1050, 752, and 600 μs) along the x-axis. The y-axis indicates the applied electric field. The frequencies of the DC pulses were 500, 800, 1050, and 1250 Hz.

https://doi.org/10.1371/journal.pone.0304345.g001

2.3 Experimental setup for applying electric field in the microchamber interior

We have investigated the effects of frequency of DC pulses of electric field on a ‘single GUV’. A ‘single GUV’ refers specifically to the targeted GUV used for investigation upon the application of an external electric field. A field of view of raw image of GUVs at different time points is provided in the supporting information (S) (see S1). In addition, the ‘single GUV’ method [20] is described briefly for understanding the statistical analysis of the experimental data in the S2. The experimental setup of a microchamber at the stage of an inverted phase contrast microscope is depicted in Fig 1(A). The microchamber consists of a glass slide, a silicon rubber spacer, and a cover slip, arranged to create a U-shaped structure. The size of the microchamber and electrode configuration are shown in Fig 1(B). Fig 1(C) shows an illustration of a ‘single GUV’ between the two gold coated electrodes. Usually, vesicle suspension contains different sizes of GUVs. The electrical model of a membrane is characterized by a capacitor Cm and a resistor Rm as shown in Fig 1(D). A rectangular DC pulse is used in the investigations (Fig 1E). The detailed experimental technique has been reported previously [33]. Fig 1(F) shows an illustration of the rectangular DC pulses of electric field with fixed ON time of 200 μs and varying OFF times of 1800, 1050, 752, and 600 μs, generates corresponding frequency 500, 800, 1050, and 1250 Hz.

2.4 Rupture and electrofusion of GUVs

At first, we explain how DC pulses of electric field induces lateral electric tension in the membranes of GUVs. Let us consider a spherical ‘single GUV’ of radius R (Fig 1C). The electric field propels the inside and outside charged ions towards the vesicle membrane, causing the membrane to become charged, similar to the behavior of a capacitor. The accumulation of charges along the membrane generates transmembrane voltage (Vm) according to the Schwan’s equation [34]: (1) where, the characteristic charging time of the membrane is τ, the angle, θ, between the direction of the field E and the normal to the bilayer surface vary from 0 to 90°. Vm = 1.5RE = Vc at θ = 90°, which is the ‘critical membrane voltage for breakdown of GUV [35]. Generally, if R = 16 μm and E = 332 V/cm, Vm = 0.8 V. The transmembrane voltage leads to induce the lateral electric tension σe in the membrane as follows [35]: (2) where, εm is relative membrane permittivity (~4.5), ε0 is the permittivity of free space, h is the thickness of membrane (~4 nm) and he is the membrane dielectric thickness (~2.8 nm) [13, 36, 37]. After simplification, the following equation can be obtained for a spherical shaped GUV [19]: (3)

For simplicity, we considered uniform tension in the membrane induced by electric stress, although recent numerical results show non-uniform membrane tension for non-spherical (i.e., elliptical) shaped vesicles with different conductivity ratios between the inside and outside of the vesicle [38].

We examined a ‘single GUV’ in a microchamber. The size range of GUVs was 26−30 μm. At first, we measured the radius of each selected ‘single GUV’, and then determined the required electric tension, i.e., σe = 5.2 mN/m using Eq (3) for a specific frequency (say 500 Hz) by changing E. To induce the required membrane tension (σe) to rupture a ‘single GUV’ or to fuse multiple GUVs of typical diameter, the corresponding range of electric field (E) was 320–340 V/cm to monitor the stochastic phenomena of the GUVs for a maximum time of 60 s. The time of rupture is defined as the moment when a ‘single GUV’ initiates the formation of rupture. Similarly, the time of electrofusion is defined as the moment when two or multiple GUVs commence the process of fusing into a single GUV. In both cases, the initial time point (t = 0) is defined as the moment when electric tension is just applied to the membranes. The recorded video was employed to identify the initiation of rupture or fusion formation. We conducted 3 to 4 independent experiments (e.g., N = 3–4), examining 13 to 16 GUVs (e.g., n = 13–16) in each individual experiment, for each specified frequency. In the case of electrofusion of ‘two-connected GUVs’ or ‘multiple-connected GUVs’, we measured the radius of bigger size GUV among the connected GUVs and then calculated σe = 5.2 mN/m using Eq (3). By statistically analyzing the data, we determined both the probability and rate constant of rupture and electrofusion. As larger GUV requires higher membrane tension to induce pore formation, we chose a larger size GUV for the electrofusion. It is noteworthy that we examine a single event such as rupture or electrofusion in each microchamber. Due to technical constraints, individual GUVs were examined for a maximum duration of 60 s. GUVs that did not undergo rupture/fusion within this 60 s timeframe were classified as intact GUVs for a particular chamber. We have calculated the average time given the distribution of rupture/fusion event that have occurred within 60 s.

3. Results

We initially present the qualitative description of the experimental results on the rupture and electrofusion of GUVs at different frequencies of direct current (DC) pulses of electric field. Then the results are analyzed statistically for the quantification of different parameters, e.g., average time, probability, and rate constant of rupture and electrofusion.

3.1 Rupture of DOPG/DOPC (40/60)-GUVs under different frequencies

We investigated the rupture in DOPC/DOPG (40/60)-GUVs at different frequencies. A representative result of one independent experiment is presented in Fig 2. At first, we consider the case of 500 Hz. Before applying the tension, σe = 5.2 mN/m in the membranes, the ‘single GUV’ remains perfectly spherical in shape at 0 s as shown in Fig 2(A). After applying the tension, the GUV remains intact with the same shape at observed at 0 s until time 41 s. At time 42 s, the shape of the GUV becomes distorted and started to rupture (i.e., rupture occurs). The structure of the GUV vanishes at time 43 s (Fig 2A). We have performed the same experiment and obtained the similar phenomena of GUVs at 800, 1050, and 1250 Hz. The corresponding times for rupture of GUVs are observed at 34 s, 27 s, and 22 s. Through optical microscopy, the nano-sized pores were not directly visible at the time of rupture, but there was a noticeable decrease in the contrast of the GUVs. In certain instances, the observation of internal content leakage was noted. The initiation of electropore formation in the lipid bilayer is detected within a few nano to microseconds [39]. Therefore, the utilization of a high-resolution camera, along with other instrumental setups, is necessary. In this paper, our focus is on rupture formation, which can be readily observed through optical microscopy. It is well accepted that when the nanopores underwent rapid enlargement over time, GUVs experienced complete rupture while the electric tension persisted. This process is akin to the aspiration of a ‘single GUV’ using the micropipette technique [25].

thumbnail
Fig 2. Rupture of DOPG/DOPC (40/60)-GUVs at different frequencies of DC pulses under σe = 5.2 mN/m.

The phase contrast images of the rupture of GUVs at (A) 500, (B) 800, (C) 1050, and (D) 1250 Hz. The electric field (E) direction is shown with an arrow on the left side of each figure. The numbers in each image show the time in seconds after applying E.

https://doi.org/10.1371/journal.pone.0304345.g002

In some cases, at 1250 Hz, we have observed the change in size of GUVs before rupture. A representative result of a ‘single GUV’ is shown in Fig 3. Before applying the tension at time t = 0 s, the size of GUV was 30 μm. After applying the tension, the size remains same until 5 s. At time 16 s, the size reduces to 20.7 μm, and at 18 s it becomes to 17.7 μm. The GUV becomes complete rupture at 19 s. The size reduction can be attributed to the expel of lipids from the bilayer due to the submicron pore formation at 16 s in the membranes [40, 41].

thumbnail
Fig 3. Size change before rupture of DOPG/DOPC (40/60)-GUV at 1250 Hz of DC pulses under σe = 5.2 mN/m.

The phase contrast images of the progressively decrease of GUV and resulting rupture of GUV. The electric field (E) direction is shown with an arrow on the left side of each figure. The numbers in each image show the time in seconds after applying E.

https://doi.org/10.1371/journal.pone.0304345.g003

Next, we have investigated the rupture of several ‘single GUVs’ in each frequency. In Fig 4(A), the first, second, and fifth GUVs were ruptured at 47, 30, and 48 s, respectively. The third and fourth GUVs were not ruptured within 60 s. At 500 Hz, among the 13 GUVs observed, there were 6 instances of GUV rupture (as depicted in Fig 4A). Additionally, at 800, 1050, and 1250 Hz, the numbers of ruptured GUVs were 7, 8, and 9, respectively (as presented in Fig 4B–4D, respectively). Thus, the time of rupture of several ‘single GUVs’ at a specified frequency is not same showing a stochastic nature as reported in the literature [22, 42, 43]. This is clearly indicating that the number of ruptured GUVs becomes higher at higher frequencies. The rupture time (on average) decreases as the frequency increases. The quantitative analysis of stochastic ruptures of several ‘single GUVs’ is considered in 3.3.

thumbnail
Fig 4. Rupture time (trup) of DOPG/DOPC (40/60)-GUVs at different frequencies of DC pulses under σe = 5.2 mN/m.

Stochastic rupture of several ‘single GUVs’ in one independent experiment at (A) 500, (B) 800, (C) 1050, and (D) 1250 Hz. The x-axis shows the GUV label number (m). The cross mark (×) above the bars indicates the intact GUVs until time 60 s.

https://doi.org/10.1371/journal.pone.0304345.g004

3.2 Electrofusion in two ‘single DOPG/DOPC (40/60)- GUVs’ under different frequencies

In this section, we have presented the results of electrofusion of ‘two-connected GUVs’ at different frequencies under σe = 5.2 mN/m. A representative result of electrofusion of two GUVs is presented in Fig 5 for several frequencies. At first, we consider the case of 500 Hz. Before applying the tension, the two connected GUVs remain perfectly spherical in shape at 0 s as shown in Fig 5(A). In the presence of tension, the GUVs remain intact with the same shape and size at observed at 0 s until time 15 s. In this stage the fusion neck is almost zero (see Fig 11 for clear understanding of fusion neck). At time 18 s, the connected GUVs merge slightly (i.e., fusion neck moderate) and at time 21 s the GUVs merges a greater degree (i.e., fusion neck enlarged), and the GUVs merged completely at 22 s (Fig 5A). Thus, the ‘two-connected GUVs’ transformed to a single GUV. It means the two GUVs looks like a dumbbell shaped single GUV at 21 s and this dumbbell shape disappeared within ~ 1 s and converted to a spherical shaped GUV. The electrofusion of GUVs is observed at 21 s for 500 Hz (Fig 5A), 27 s for 800 Hz (Fig 5B), 30 s for 1050 Hz (Fig 5C), and 36 s for 1250 Hz (Fig 5D). It is assumed that the electrofusion occurs due to the simultaneous pore formation in the membranes at contact point [44]. The phenomena of GUVs become similar for several independent experiments.

thumbnail
Fig 5. Fusion of DOPG/DOPC (40/60)-GUVs at different frequencies of DC electric pulses unser σe = 5.2 mN/m.

The phase contrast images of the fusion of ‘two-connected GUVs’ at (A) 500, (B) 800, (C) 1050, and (D) 1250 Hz. The electric field (E) direction is shown with an arrow on the left side of each figure. The numbers in each image show the time in seconds after applying E.

https://doi.org/10.1371/journal.pone.0304345.g005

Next, we have investigated the electrofusion of several ‘two connected GUVs’. Here, we have presented the results of one independent experiment comprising 13 ‘two connected GUVs’ in each frequency. A number of 9 GUVs are fused out of 13 GUVs at 500 Hz (Fig 6A). The same procedure was applied in several ‘two-connected GUVs’ for 800, 1050, and 1250 Hz, and observed the stochastic fusion nature of vesicles [22, 42, 43]. Out of 13 GUVs, the numbers of fused GUVs were 8 for 800 Hz (Fig 6B), 7 for 1050 Hz (Fig 6C), and 6 for 1250 Hz (Fig 6D). The experimental results shown in Fig 6 indicate that the number of fused GUVs reduces following an increase in frequency. The quantitative analysis of electrofusion is presented in 3.3.

thumbnail
Fig 6. Electrofusion time (tfus) of DOPG/DOPC (40/60)-GUVs at different frequencies of DC pulses under σe = 5.2 mN/m.

Stochastic fusion of several ‘two-connected GUVs’ in one independent experiment at (A) 500, (B) 800, (C) 1050, and (D) 1250 Hz. The x-axis shows the GUV label number (m). The cross mark (×) above the bars indicates the intact GUVs until time 60 s.

https://doi.org/10.1371/journal.pone.0304345.g006

3.3 Frequency dependent probability of rupture and electrofusion in DOPG/DOPC (40/60)-GUVs

In this section, the frequency dependent rupture and the electrofusion are analyzed statistically. Firstly, we have calculated the quantitative data of average time of rupture and electrofusion and also the probability of rupture and electrofusion. Fig 7(A) shows that the average rupture time decreases while average fusion time increases with increasing frequency. The average rupture time decreases 39.2% whereas the electrofusion time increases 37.08% for increasing the frequency from 500 to 1250 Hz. The tendency of decreasing the average time for rupture and the increasing the average time for electrofusion follows almost a linear function with frequency. Now the results of the probability of rupture (Prup) and probability of electrofusion (Pfus) are presented. Fig 7(B) shows that the probability of rupture increases while probability of fusion decreases with increasing frequency. The probability of rupture increases 53.3% whereas the probability of electrofusion decreases 35.2% for increasing the frequency from 500 to 1250 Hz. Hence, the frequency enhances the possibility of electroporation whereas the opposite trend is shown for electrofusion. The data on the frequency dependent average time of rupture, average time of electrofusion, probability of rupture, and probability of electrofusion is shown in Table 1.

thumbnail
Fig 7. Frequency dependent average rupture time, average fusion time, rupture probability, and fusion probability of DOPG/DOPC (40/60)-GUVs under σe = 5.2 mN/m.

Average values and standard deviations of each data were determined from several independent experiments.

https://doi.org/10.1371/journal.pone.0304345.g007

thumbnail
Table 1. Frequency dependent average time, probability, and rate constant of rupture and electrofusion of DOPG/DOPG (40/60)-GUVs.

https://doi.org/10.1371/journal.pone.0304345.t001

3.3 Rate constant of rupture and electrofusion in DOPG/DOPC (40/60)- GUVs under different frequencies

Now, the kinetics such as the rate constant of rupture (kp_rup) and rate constant of electrofusion (kp_fus) of the GUVs have been considered. The rate constant governing the rupture/fusion of GUVs stands as a pivotal parameter that elucidates the pore formation/electrofusion process within a collection of GUVs, subject to external influences. Precise insight into the kinetics governing the rupture of GUVs holds significant importance in comprehending the underlying mechanism driving the rupture/fusion phenomenon. To explain the stochastic phenomenon of rupture/electrofusion quantitatively, we considered a parameter Pint (t), meaning the probability of GUVs being in intact state at the instant of time, t, among all the examined GUVs. The value of Pint (t) considering rupture/electrofusion events as the transition from intact state to ruptured/fused state, is given by Eq (4) [19, 25, 45], (4) where, kp is the rate constant of rupture/fusion of GUVs and t is the duration of time to apply the tension in GUVs. We used Pint_rup for rupture and Pint_fus for electrofusion as for convenience.

To determine the rate constant of rupture/electrofusion, the values of Pint (t) have been determined from the experiment data. It can be defined as Pint (t) = 1- Prup or Pint (t) = 1- Pfus [46, 47]. Fig 8(A) shows the time course of Pint_rup (t) for different frequencies. It shows that the decrement of experimental data for Pint_rup versus time is a factor when applied frequency changes from 500 to 1250 Hz. The decrement is smaller at 500 Hz and higher at 1250 Hz. The experimental data on the Pint_rup versus time graph is well fitted by a single exponential decay function of Eq (4). From the fitted curves, the values of kp_rup are obtained 0.8×10−2, 1.3×10−2, 1.7×10−2, and 2.3×10−2 s-1 for 500, 800, 1050 and 1200 Hz, respectively. The same experiments were carried out for several independent experiment (N = 3−4). The values of average kp_rup are obtained (0.8 ± 0.1)×10−2, (1.3 ± 0.1)×10−2, (1.7 ± 0.5)×10−2, and (2.3 ± 0.8)×10−2 s-1 for 500, 800, 1050 and 1200 Hz, respectively, at 5.2 mN/m. Hence, the kp_rup increases with frequency (Fig 8B). The result for 1250 Hz in Fig 8(A) represents an independent experiment, and the data exhibits slight fluctuation. However, the average value of kp_rup (± SD) indicates an increasing trend with frequency. Next, we calculated the kp_fus as described above. Fig 8(C) shows the time course of Pint_fus (t) for different frequencies. It shows that the decrement of experimental data for Pint_fus versus time is a factor when applied frequency changes from 500 to 1250 Hz. The decrement is faster at 500 Hz and slower at 1250 Hz. The experimental data on the Pint_fus versus time graph is well fitted by a single-exponential decay function of Eq (4). From the fitted curves, the values of kp_fus are obtained 2.4×10−2, 2.3×10−2, 1.2×10−2, and 1.0×10−2 s-1 for 500, 800, 1050, and 1250 Hz, respectively. The same experiments were carried out for several times (N = 3−4) and the values of average kp_fus were obtained (2.4 ± 0.1)×10−2, (2.3 ± 0.4)×10−2, (1.2 ± 0.1)×10−2, and (1.0 ± 0.1)×10−2 s-1 for 500, 800, 1050 and 1200 Hz, respectively, at 5.2 mN/m. Hence, the kp_fus decreases with frequency (Fig 8D). The data on the frequency dependent rate constant of rupture and electrofusion is shown in Table 1.

thumbnail
Fig 8. The time course of the fraction of intact DOPG/DOPC (40/60)-GUVs and the rate constant of rupture and electrofusion at different frequencies of DC pulses under σe = 5.2 mN/m.

(A) The time course of the fraction of intact GUVs at 500, 800, 1050, and 1250 Hz in the case of rupture of vesicles. (B) The frequency-dependent rate constant of rupture of GUVs. (C) The time course of the fraction of intact GUVs at 500, 800, 1050, and 1250 Hz in the case of electrofusion of vesicles. (D) The frequency-dependent rate constant of electrofusion of ‘two-connected GUVs’. The experimental data in (A) and (C) is fitted using a single exponential decay function of Eq 4. Average values and standard deviations of each data in (C) and (D) were determined from several independent experiments comprising many ‘single GUVs’.

https://doi.org/10.1371/journal.pone.0304345.g008

As the time course of the fraction of intact GUVs (see Fig 8) follows a single exponential decay function, it confirms the stochastic nature of rupture/fusion events. Essentially, the fraction of intact GUVs has been derived from the probability of rupture/fusion of GUVs. Exponential decay results directly from events occurring randomly over time, where there exists a meaningful average rate for such occurrences.

3.4 Electrofusion in multiple DOPG/DOPC (40/60)- GUVs at 500 Hz

In 3.2, we have described the results of electrofusion of ‘two-connected GUVs’ at different frequencies under σe = 5.2 mN/m. Here we focus the electrofusion of ‘multiple-connected GUVs’ at a fixed frequency, e.g., 500 Hz. A representative result of electrofusion of ‘three-connected GUVs’, ‘four-connected GUVs’, and ‘five-connected GUVs’ is shown in Fig 9. The phenomena to fuse the multiple GUVs is similar to that obtained for fusing of two GUVs as described in 3.2. The similar behavior of electrofusion is also observed for other frequencies. We have summarized the volume of each ‘single GUV’ as presented in Figs 5(A) and 9(A–9C) at 0 s (initial stage) and compared with the volume of fused GUV at final stage. Actually, we have considered to verify the equation, , where ri is the radius of GUVs at 0 s (the GUVs contributing to electrofusion), R is the radius of GUV at final stage (the fused GUV), and Nt is the number of electro-fused GUVs. In Fig 5(A) at 0 s, r1 = 15.48 μm, r2 = 8.14 μm, and thus volume V1 = 4π/3 (r13 + r23) = 17811 μm3. At final stage (i.e., 46 s), volume V2 = 4π/3 (16.24)3 = 17951 μm3. Hence, V2/V1 = 1.008. We have calculated V2/V1 for the case of GUVs shown in Fig 9(A)–9C, and found the values as 1.002, 1.004, and 1.002, respectively. Thus, the average value with SD as V2/V1 = 1.004 ± 0.003 ≈ 1.0.

thumbnail
Fig 9. Electrofusion of multiple DOPG/DOPC (40/60)-GUVs at 500 Hz of DC pulses under σe = 5.2 mN/m.

The phase contrast images of electrofusion of (A) ‘three-connected GUVs’, (B) ‘four-connected GUVs’, and (C) ‘five-connected GUVs’. The electric field (E) direction is shown with an arrow on the left side of each figure. The numbers in each image show the time in seconds after applying E.

https://doi.org/10.1371/journal.pone.0304345.g009

4. Discussion

In this research, we have investigated the rupture and electrofusion of cell-mimetic GUVs under various frequencies of DC pulses. Both the phenomena depend on the applied frequency. As the frequency increases the probability of rupture and the rate constant of rupture increases, whereas electrofusion shows the opposite nature. The mechanism of these phenomena is discussed in detail in 4.1 and 4.2.

4.1 Vesicle rupture and its mechanism

We have conducted the experiments on vesicle rupture using different frequencies. It shows two notable phenomena: vesicle rupture started without alterations in size (Fig 2) and vesicle rupture subsequent to changes in size (Fig 3). A stochastic rupture is observed at each frequency (Fig 4). A rise in frequency leads to a higher occurrence of ruptures (Figs 7 and 8A). The free energy of a prepore in the lipid vesicle is expressed as follows [15, 46]: (5) where Γ represents the free energy per unit length of a prepore (i.e., pore edge tension or line tension), σe is the lateral electric tension, r is the radius of a prepore, and the parameter B (= 1.76 mNm-1) represents the electrostatic interaction arising from the surface charge of membrane [48]. The term -πr2σe favors prepore expansion and the term 2πrΓ favors prepore closure. We specifically considered the toroidal structure of the prepore as illustrated in Fig 10.

The energy barrier for pore formation at critical radius rc = Γ/ (σe + B) is expressed [49]: (6)

According to the Arrhenius equation the rate constant is expressed as follows [50]: (7) where, A is a constant (in s-1), kB is the Boltzmann constant, and T is absolute temperature.

thumbnail
Fig 10. Schematic illustration of the possible steps for the rupture of a ‘single GUV’.

(A) A ‘single GUV’ with (i) intact (ii) single pore, and (iii) ruptured conditions. (B) Bilayers with (i) intact (ii) transverse view of a hydrophilic toroidal pore, and (iii) segregated conditions. In (ii) σe is the lateral tension, Γ is the pore edge tension, and r is the pore radius. The electric field (E) direction is shown with an arrow.

https://doi.org/10.1371/journal.pone.0304345.g010

The equivalent circuit of lipid bilayers takes the form of a capacitor (capacitance Cm) in parallel with a resistor (Rm) as shown in Fig 1(D). Basically, Rm comes from combination of an extracellular resistance (Re) in parallel with the series combination of the membrane capacitance and an intracellular resistance (Ri). The total impedance of the circuit shown in Fig 1(D) [51]:

, where Rm is frequency (f)-independent and the capacitive reactance Xc (= −1/jf Cm) depends on f, where j indicates the imaginary part. The amplitude of the impedance of the lipid bilayers, (8)

With the increase of frequency of DC pulses, the value of Z decreases (i.e., conductance increases). It is to be noted that electroporation intensity can be increased by increasing the pulse amplitude, pulse width (i.e., frequency) and/or pulse number. In our investigations, we changed the frequency of the signal while keeping the membrane tension constant.

The time evolution of pores (n) of radius, r, is represented as a diffusion process in the potential function ΔWpWp has the same physical concept of U(r, σe)), which can be expressed by the Einstein–Smoluchowski equation [51]: (9) where, D is the diffusion coefficient of pores in the space of pore radius, and the source term S accounts for the rate of pore creation and annihilation. The parameter ΔWp governs the evolution of the pore population, representing the energy required for pore formation in the membrane and can be considered equivalent to the free energy associated with pore formation. The increase in pore formation enhances the conductance in the lipid bilayer, which further supports the discussion mentioned above based on the parallel RmCm circuit. The increase in conductance effectively decreases the energy barrier (i.e., Ub) for electroporation (i.e., rupture) and hence increase the rate constant of rupture according to Eq (7).

Again, the probability of rupture Pp_rup within 60 s duration of observation is presented as follows [25]: (10)

As described earlier in the possible mechanism of vesicle rupture induced by DC pulses of various frequencies (see 4.1), this size reduction of the GUVs can be attributed to the expulsion of lipid molecules from the membrane in the form of small vesicles and/or tubules [40]. The quantity of ejected lipids is directly proportional to the permeabilized membrane area. A microscopy study confirmed a similar lipid ejection phenomenon, thereby affirming that lipid loss is not an artifact of membrane labeling [41]. The observed size change of GUVs at 1250 Hz before rupture (Fig 3) can be elucidated based on the above discussion. Hence, the increase in membrane conductance and the possibility of pore formation at higher frequency results in the higher probability and rate constant of rupture, as obtained in our experiments (see Figs 2, 3, 4, 7 and 8A).

Recently, the deformation of GUVs has been explored at different electric field [1, 52]. The authors introduced a slightly different conductivity between the inside and outside of the GUVs. In addition, they utilized different lipids compared to our study. Thus, they observed a clear change in shape in response to the electric field. In contrast, our research focused on the investigation of GUV rupture and fusion, along with an examination of their kinetics under various frequencies of DC pulses. We maintained identical conductivity both inside and outside the GUVs. It is worthy to mention that the electroporation models are sometimes unable or at least show limitations to explain the experimental data reported recently [53]. Hence the inferences made from the electroporation models should be taken with some care.

Recently, a Boundary Integral Method (BIM) was employed to study an elliptical vesicle placed in a uniform electric field [38]. The investigation primarily focused on the deformation of non-spherical shaped vesicles in the presence of DC electric pulses with the aim of understanding cell electroporation. The computational methodology did not allow for the simulation of actual pore formation due to topological changes in the membrane; rather, it examined tension evolution and distribution in the membranes. Simulation of vesicle dynamics in DC pulses revealed the critical role played by the conductivity ratio, Λ, and field strength in shape evolution. While a vesicle with Λ > 1 remained prolate throughout, with the highest tension at the poles (and hence most likely to porate there), a vesicle undergoing prolate-oblate transition (Λ < 1 and sufficiently high field strength) exhibited more complex tension evolution and was likely to porate in a region between the pole and the equator. These results underscored the importance of shape changes of non-spherical (i.e., elliptical) vesicles during the electroporation process.

In our study, we investigated electroporation (i.e., rupture) and electrofusion of spherical vesicles with a conductivity ratio of Λ = 1. Additionally, the electric field applied to the vesicles was relatively low, ranging from 320 to 340 V/cm. As for simplicity, we considered uniform tension in the membrane induced by electric stress.

Commonly, the pure lipid membrane is considered as an insulator separating two electrically conductive compartments. The equivalent circuit of lipid bilayers takes the form of a capacitor in parallel with a resistor, see Fig 1A.

4.2 Electrofusion mechanism

In this experiment, we observed that electrofusion is more pronounced at lower frequencies and decreases at higher frequencies (Figs 58B). At lower frequencies, the presence of pores in the contact region of the GUVs promotes their electrofusion. However, at higher frequencies, electrofusion decreases because the GUVs tend to rupture rather than fuse. The precise molecular mechanisms underlying how membrane electroporation facilitates fusion are not fully understood. Previously, a model was proposed suggesting that the electric field induces pores that span across both adjacent membranes in the contact zone [44]. Essentially, the formation of a pore in one of the bilayers could locally increase the electric field, promoting the nucleation of another pore in the adjacent bilayer. If numerous such double-membrane pores are nucleated, they could coalesce into larger loop-like and tongue-like cracks. High-speed optical imaging with a time resolution of ~ 50 μs during the electrofusion process between two GUVs revealed that several double-membrane pores (fusion necks) typically form in the contact zone during the pulse [54, 55]. The expansion and subsequent coalescence of these fusion necks result in the formation of small contact-zone vesicles, which remain trapped within the interior of the fused GUV. This same mechanism is also applicable to the electrofusion of multiple GUVs (Fig 9).

The most accepted model of electrofusion of lipid vesicles is the ‘The stalk mechanism’ [56]. A schematic illustration of several steps of the electrofusion of stalk mechanism is depicted in Fig 11.

thumbnail
Fig 11. Schematic illustration of the possible steps of ‘stalk mechanism’ of the electrofusion of two ‘single GUVs’.

(A) Stepwise changing of (i) two closely approached bilayers (ii) form a stalk following close apposition of the interfaces (iii) the monolayers undergo a continuous deformation toward a trans-monolayer contact (TMS), (iv) the strongly curved dimple ruptures under stress to form a fusion pore, and (v) merged completely. (B) Stepwise changing of two ‘single GUVs’ which are (i) approaching closely, (ii) at contact position, (iii) merged slightly, (iv) merged greatly, and (v) fused completely. The length of fusion neck (L) progresses with time. The electric field (E) direction is shown with an arrow.

https://doi.org/10.1371/journal.pone.0304345.g011

In this mechanism, the membrane fusion is thought to occur through intermediate structure known as ‘stalk’. This stalk is form due to the bending energy (related to bending modulus) of the lipid bilayer [57]. Theoretical analysis of the stalk has allowed to determine the elastic energy associated with this structure and make predictions regarding the possible evolution of the overall process by assuming that the stalk has a shape of revolution of a circular arc. The energy of the stalk is expressed as follows: (11) where, κ is bending rigidity, cm and cp are principal curvatures along the meridian and parallel to the body of revolution representing the stalk, and c0 is the spontaneous curvature. The first integral represents the bending energy of the stalk membrane and the second integral is equal to the bending energy of the initial membrane. Later the stalk structure was revisited [58] by considering the correct shape of the stalk. In the revised form, the stalk is considered completely stress free and the calculated energy become negative instead of high positive value, facilitating the whole fusion process.

In electrofusion experiments, vesicles or cells are typically brought into contact through the application of a low-intensity AC electric field. When the frequency and intensity of the AC field are appropriately chosen, attractive electrostatic interaction forces come into play between individual GUVs, causing them to align linearly in response to the direction of the applied electric field [59]. In our investigations, we found a high population of GUVs during the synthesis process, resulting in many vesicles being interconnected. This outcome was achieved by drying lipids in 3–4 vials and subsequently purifying them together.

5. Conclusions

The cell mimetic vesicle rupture and electrofusion were studied under different frequencies of DC pulses, revealing interesting findings. Both the rupture and electrofusion followed stochastic phenomena for many GUVs. Understanding stochastic phenomena is crucial in various scientific disciplines as it allows researchers to make probabilistic predictions. The probability and the kinetics of rupture increase as the frequency increases. The frequency dependent probability and kinetics of electrofusion show the opposite trend. The investigation enables to identify the optimal frequency for effective electroporation or facilitation of electrofusion. These findings can inspire to develop novel liposome approaches for practical biomedical applications.

References

  1. 1. Riske KA, Dimova R. Electro-deformation and poration of giant vesicles viewed with high temporal resolution. Biophys J. 2005;88: 1143–1155. pmid:15596488
  2. 2. Yarmush ML, Golberg A, Serša G, Kotnik T, Miklavčič D. Electroporation-based technologies for medicine: principles, applications, and challenges. Annu Rev Biomed Eng. 2014;16: 295–320. pmid:24905876
  3. 3. Haberl Meglič S, Kotnik T. Electroporation-based applications in biotechnology. In: Miklavčič D, editor. Handbook of Electroporation. Cham: Springer International Publishing; 2017. pp. 2153–2169. https://doi.org/10.1007/978-3-319-32886-7_33
  4. 4. Muralidharan A, Boukany PE. Electrotransfer for nucleic acid and protein delivery. Trends Biotechnol. 2023. pmid:38102019
  5. 5. Campelo SN, Huang P-H, Buie CR, Davalos RV. Recent Advancements in electroporation technologies: from bench to clinic. Annu Rev Biomed Eng. 2023;25: 77–100. pmid:36854260
  6. 6. Miller L, Leor J, Rubinsky B. Cancer cells ablation with irreversible electroporation. Technol Cancer Res Treat. 2005;4: 699–705. pmid:16292891
  7. 7. Rubinsky B, Onik G, Mikus P. Irreversible Electroporation: A new ablation modality—clinical implications. Technol Cancer Res Treat. 2007;6: 37–48. pmid:17241099
  8. 8. Dimitrov DS. Chapter 18—Electroporation and electrofusion of membranes. In: Lipowsky R, Sackmann E, editors. Handbook of Biological Physics. North-Holland; 1995. pp. 851–901. https://doi.org/10.1016/S1383-8121(06)80011-4
  9. 9. Strömberg A, Karlsson A, Ryttsén F, Davidson M, Chiu DT, Orwar O. Microfluidic device for combinatorial fusion of liposomes and cells. Anal Chem. 2001;73: 126–130. pmid:11195496
  10. 10. Li X, Yang F, Rubinsky B. A Correlation between electric fields that target the cell membrane potential and dividing HeLa cancer cell growth inhibition. IEEE Trans Biomed Eng. 2021;68: 1951–1956. pmid:33275576
  11. 11. Dastani K, Moghimi Zand M, Kavand H, Javidi R, Hadi A, Valadkhani Z, et al. Effect of input voltage frequency on the distribution of electrical stresses on the cell surface based on single-cell dielectrophoresis analysis. Sci Rep. 2020;10: 68. pmid:31919394
  12. 12. Peng W, Polajžer T, Yao C, Miklavčič D. Dynamics of Cell Death Due to Electroporation using different pulse parameters as revealed by different viability assays. Ann Biomed Eng. 2023. pmid:37704904
  13. 13. Dimova R, Bezlyepkina N, Jordö MD, Knorr RL, Riske KA, Staykova M, et al. Vesicles in electric fields: Some novel aspects of membrane behavior. Soft Matter. 2009;5: 3201.
  14. 14. Politano TJ, Froude VE, Jing B, Zhu Y. AC-electric field dependent electroformation of giant lipid vesicles. Coll Surf B: Biointerf. 2010;79: 75–82. pmid:20413284
  15. 15. Karal MAS, Orchi US, Towhiduzzaman M, Ahamed MK, Ahmed M, Ahammed S, et al. Electrostatic effects on the electrical tension-induced irreversible pore formation in giant unilamellar vesicles. Chem Phys Lipids. 2020;231: 104935. pmid:32569600
  16. 16. Karal MAS, Ahamed MK, Mokta NA, Ahmed M, Ahammed S. Influence of cholesterol on electroporation in lipid membranes of giant vesicles. Eur Biophys J. 2020;49: 361–370. pmid:32535676
  17. 17. Sarkar MK, Karal MAS, Ahmed M, Ahamed MK, Ahammed S, Sharmin S, et al. Effects of osmotic pressure on the irreversible electroporation in giant lipid vesicles. PLOS ONE. 2021;16: e0251690. pmid:33989363
  18. 18. Sarkar MK, Karal MAS, Levadny V, Belaya M, Ahmed M, Ahamed MK, et al. Effects of sugar concentration on the electroporation, size distribution and average size of charged giant unilamellar vesicles. Eur Biophys J. 2022;51: 401–412. pmid:35716178
  19. 19. Ahamed MK, Karal MAS, Ahmed M, Ahammed S. Kinetics of irreversible pore formation under constant electrical tension in giant unilamellar vesicles. Eur Biophys J. 2020;49: 371–381. pmid:32494845
  20. 20. Yamazaki M. Chapter 5 The single GUV method to reveal elementary processes of leakage of internal contents from liposomes induced by antimicrobial substances. Advances in Planar Lipid Bilayers and Liposomes. Academic Press; 2008. pp. 121–142. https://doi.org/10.1016/S1554-4516(08)00005-7
  21. 21. Böckmann RA, de Groot BL, Kakorin S, Neumann E, Grubmüller H. Kinetics, statistics, and energetics of lipid membrane electroporation studied by molecular dynamics simulations. Biophys J. 2008;95: 1837–1850. pmid:18469089
  22. 22. Karal MAS, Ahamed MK, Ahmed M, Mahbub ZB. Recent developments in the kinetics of ruptures of giant vesicles under constant tension. RSC Adv. 2021;11: 29598–29619. pmid:35479542
  23. 23. Billah MM, Saha SK, Rashid MMO, Hossain F, Yamazaki M. Effect of osmotic pressure on pore formation in lipid bilayers by the antimicrobial peptide magainin 2. Phys Chem Chem Phys. 2022;24: 6716–6731. pmid:35234764
  24. 24. Evans E, Smith BA. Kinetics of hole nucleation in biomembrane rupture. New J Phys. 2011;13: 095010. pmid:21966242
  25. 25. Levadny V, Tsuboi T, Belaya M, Yamazaki M. Rate Constant of tension-induced pore formation in lipid membranes. Langmuir. 2013;29: 3848–3852. pmid:23472875
  26. 26. Billah MM, Rashid MMO, Ahmed M, Yamazaki M. Antimicrobial peptide magainin 2-induced rupture of single giant unilamellar vesicles comprising E. coli polar lipids. Biochim Biophys Acta (BBA)—Biomembr. 2023;1865: 184112. pmid:36567034
  27. 27. Sengel JT, Wallace MI. Imaging the dynamics of individual electropores. Proc Natl Acad Sci USA. 2016;113: 5281–5286. pmid:27114528
  28. 28. Sengel JT, Wallace MI. Measuring the potential energy barrier to lipid bilayer electroporation. Phil Trans R Soc B: Biol Sci. 2017;372: 20160227. pmid:28630163
  29. 29. Muralidharan A, Rems L, Kreutzer MT, Boukany PE. Actin networks regulate the cell membrane permeability during electroporation. Biochim Biophys Acta (BBA)—Biomembr. 2021;1863: 183468. pmid:32882211
  30. 30. Reeves JP, Dowben RM. Formation and properties of thin-walled phospholipid vesicles. J Cell Physiol. 1969;73: 49–60. pmid:5765779
  31. 31. Tamba Y, Terashima H, Yamazaki M. A membrane filtering method for the purification of giant unilamellar vesicles. Chem Phys Lipids. 2011;164: 351–358. pmid:21524642
  32. 32. Karal MAS, Rahman M, Ahamed MK, Shibly SUA, Ahmed M, Shakil MM. Low cost non-electromechanical technique for the purification of giant unilamellar vesicles. Eur Biophys J. 2019;48: 349–359. pmid:30918998
  33. 33. Karal MAS, Ahamed MK, Rahman M, Ahmed M, Shakil MM, Rabbani KS. Effects of electrically-induced constant tension on giant unilamellar vesicles using irreversible electroporation. Eur Biophys J. 2019;48: 731–741. pmid:31552440
  34. 34. Grosse C, Schwan HP. Cellular membrane potentials induced by alternating fields. Biophys J. 1992;63: 1632–1642. pmid:19431866
  35. 35. Needham D, Hochmuth RM. Electro-mechanical permeabilization of lipid vesicles. Role of membrane tension and compressibility. Biophys J. 1989;55: 1001–1009. pmid:2720075
  36. 36. Lisin R, Zion Ginzburg B, Schlesinger M, Feldman Y. Time domain dielectric spectroscopy study of human cells. I. Erythrocytes and ghosts. Biochim Biophys Acta (BBA)—Biomembr. 1996;1280: 34–40. pmid:8634314
  37. 37. Tanizaki S, Feig M. A generalized Born formalism for heterogeneous dielectric environments: Application to the implicit modeling of biological membranes. J Chem Phys. 2005;122: 124706. pmid:15836408
  38. 38. McConnell LC, Miksis MJ, Vlahovska PM. Continuum modeling of the electric-field-induced tension in deforming lipid vesicles. J Chem Phys. 2015;143: 243132. pmid:26723617
  39. 39. Dimova R, Riske KA, Aranda S, Bezlyepkina N, Knorr RL, Lipowsky R. Giant vesicles in electric fields. Soft Matter. 2007;3: 817. pmid:32900072
  40. 40. Portet T, Camps i Febrer F, Escoffre J-M, Favard C, Rols M-P, Dean DS. Visualization of membrane loss during the shrinkage of giant vesicles under electropulsation. Biophys J. 2009;96: 4109–4121. pmid:19450482
  41. 41. Mauroy C, Portet T, Winterhalder M, Bellard E, Blache M-C, Teissié J, et al. Giant lipid vesicles under electric field pulses assessed by non invasive imaging. Bioelectrochemistry. 2012;87: 253–259. pmid:22560131
  42. 42. Islam MZ, Alam JM, Tamba Y, Karal MAS, Yamazaki M. The single GUV method for revealing the functions of antimicrobial, pore-forming toxin, and cell-penetrating peptides or proteins. Phys Chem Chem Phys. 2014;16: 15752–15767. pmid:24965206
  43. 43. Moghal MMR, Hossain F, Yamazaki M. Action of antimicrobial peptides and cell-penetrating peptides on membrane potential revealed by the single GUV method. Biophys Rev. 2020;12: 339–348. pmid:32152921
  44. 44. Sugar IP, Förster W, Neumann E. Model of cell electrofusion: Membrane electroporation, pore coalescence and percolation. Biophys Chem. 1987;26: 321–335. pmid:3607233
  45. 45. Karal MAS, Billah MM, Ahamed MK. Determination of pore edge tension from the kinetics of rupture of giant unilamellar vesicles using the Arrhenius equation: effects of sugar concentration, surface charge and cholesterol. Phys Chem Chem Phys. 2024;26: 6107–6117. pmid:38299672
  46. 46. Litster JD. Stability of lipid bilayers and red blood cell membranes. Phys Lett A. 1975;53: 193–194.
  47. 47. Tamba Y, Yamazaki M. Single Giant Unilamellar Vesicle Method Reveals Effect of Antimicrobial Peptide Magainin 2 on Membrane Permeability. Biochemistry. 2005;44: 15823–15833. pmid:16313185
  48. 48. Karal MAS, Levadnyy V, Tsuboi T, Belaya M, Yamazaki M. Electrostatic interaction effects on tension-induced pore formation in lipid membranes. Phys Rev E. 2015;92: 012708. pmid:26274204
  49. 49. Karal MAS, Levadnyy V, Yamazaki M. Analysis of constant tension-induced rupture of lipid membranes using activation energy. Phys Chem Chem Phys. 2016;18: 13487–13495. pmid:27125194
  50. 50. Kulczycki A, Kajdas C, Kulczycki A, Kajdas C. A new attempt to better understand Arrehnius equation and its activation energy. Tribol Eng. IntechOpen; 2013.
  51. 51. Miklavčič D, Serša G. Handbook of electroporation. Springer International Publishing; 2020.
  52. 52. Kummrow M, Helfrich W. Deformation of giant lipid vesicles by electric fields. Phys Rev A. 1991;44: 8356–8360. pmid:9905991
  53. 53. Scuderi M, Dermol-Černe J, Amaral da Silva C, Muralidharan A, Boukany PE, Rems L. Models of electroporation and the associated transmembrane molecular transport should be revisited. Bioelectrochemistry. 2022;147: 108216. pmid:35932533
  54. 54. Riske KarinA, Bezlyepkina N, Lipowsky R, Dimova R. Electrofusion of model lipid membranes viewed with high temporal resolution. Biophys Rev Lett. 2006;01: 387–400.
  55. 55. Haluska CK, Riske KA, Marchi-Artzner V, Lehn J-M, Lipowsky R, Dimova R. Time scales of membrane fusion revealed by direct imaging of vesicle fusion with high temporal resolution. Proc Natl Acad Sci USA. 2006;103: 15841–15846. pmid:17043227
  56. 56. Markin VS, Kozlov MM, Borovjagin VL. On the theory of membrane fusion. The stalk mechanism. Gen Physiol Biophys. 1984;3: 361–377. pmid:6510702
  57. 57. Karal MAS, Billah MM, Ahmed M, Ahamed MK. A review on the measurement of the bending rigidity of lipid membranes. Soft Matter. 2023;19: 8285–8304. pmid:37873600
  58. 58. Markin VS, Albanesi JP. Membrane fusion: stalk model revisited. Biophys J. 2002;82: 693–712. pmid:11806912
  59. 59. Zimmermann U. Electric field-mediated fusion and related electrical phenomena. Biochim Biophys Acta (BBA)—Rev Biomembr. 1982;694: 227–277. pmid:6758848