Next Article in Journal
Microbial Fuel Cells as Effective Tools for Energy Recovery and Antibiotic Detection in Water and Food
Previous Article in Journal
Optical Fiber Sensing-Aided 3D-Printed Replacement Parts for Enhancing the Sensing Ability of Architectural Heritage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dispersive Modeling of Normal and Cancerous Cervical Cell Responses to Nanosecond Electric Fields in Reversible Electroporation Using a Drift-Step Rectifier Diode Generator

1
Technical Research Analyst (TRA), Electronics/Biomedical Engineering, Aranca, Mumbai 400076, Maharastra, India
2
Department of Electronics and Communication Engineering, Galgotias College of Engineering and Technology, Greater Noida 201310, Uttar Pradesh, India
3
School of Computer Science Engineering and Applications, D Y Patil International University, Pune 411044, Maharastra, India
4
School of Data Science, Indian Institute of Science Education and Research Thiruvananthapuram (IISER TVM), Vithura, Thiruvananthapuram 695551, Kerala, India
5
Department of Information Technology, College of Computer and Information Sciences, Princess Nourah bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
6
University Centre for Research and Development, Chandigarh University, Gharuan, Mohali 140413, Punjab, India
*
Authors to whom correspondence should be addressed.
Micromachines 2023, 14(12), 2136; https://doi.org/10.3390/mi14122136
Submission received: 22 August 2023 / Revised: 16 November 2023 / Accepted: 18 November 2023 / Published: 22 November 2023
(This article belongs to the Section E:Engineering and Technology)

Abstract

:
This paper creates an approximate three-dimensional model for normal and cancerous cervical cells using image processing and computer-aided design (CAD) tools. The model is then exposed to low-frequency electric pulses to verify the work with experimental data. The transmembrane potential, pore density, and pore radius evolution are analyzed. This work adds a study of the electrodeformation of cells under an electric field to investigate cytoskeleton integrity. The Maxwell stress tensor is calculated for the dispersive bi-lipid layer plasma membrane. The solid displacement is calculated under electric stress to observe cytoskeleton integrity. After verifying the results with previous experiments, the cells are exposed to a nanosecond pulsed electric field. The nanosecond pulse is applied using a drift-step rectifier diode (DSRD)-based generator circuit. The cells’ transmembrane voltage (TMV), pore density, pore radius evolution, displacement of the membrane under electric stress, and strain energy are calculated. A thermal analysis of the cells under a nanosecond pulse is also carried out to prove that it constitutes a non-thermal process. The results showed differences in normal and cancerous cell responses to electric pulses due to changes in morphology and differences in the cells’ electrical and mechanical properties. This work is a model-driven microdosimetry method that could be used for diagnostic and therapeutic purposes.

1. Introduction

Reversible electroporation (RE) occurs when a pulsed electric field creates pores in the plasma membranes of biological cells, which then reseal after some time, without inducing apoptosis or cell death. This method is used to deliver drugs, dyes, and genes into cells. Nanosecond pulsed electric fields have previously been applied to biological cells in reversible electroporation applications [1,2,3]. Many works have used numerical models to understand the electroporation process and the effects of nanosecond pulse strength on biological cells [3,4,5]. These numerical models have considered the biological cells as dispersive models using Debye’s second-order dispersive expression to show the change in dielectric strength with respect to the high-frequency electric pulse [6,7,8,9]. The use of dispersive models has enabled reductions in the strength of the electric field and more accurate predictions of the strength and duration of the electric field required to electroporate cells [10,11,12,13]. In order to make the models more realistic, the two-dimensional morphology of the original biological cells extracted using image processing algorithms has been used to demonstrate the model of electroporation [14,15,16]. These studies supported the reversible electroporation method but advocated using three-dimensional models for more accurate predictions [17,18,19]. The problem of infinite long columns was found to be a limitation when numerically studying all two-dimensional biological cells, whether simple geometries or contours extracted via image processing methods [14]. A realistic three-dimensional model provided more accurate predictions, as demonstrated in [14]. In a previous study, reversible electroporation of a two-dimensional numerical model for all types of cervical cells (superficial, intermediate, parabasal, and basal) was studied [20]. Cervical cells have also been experimentally exposed to electric pulses to achieve electroporation. Cancerous cervical cells have been exposed to a range of pulses with different durations and repetition rates. Electric fields ranging from 0.5 to 5 kV/cm and durations of 25–100 μ s rectangular pulses have been used for the electroporation of cervical cells [21]. Another study used Gaussian pulses with a strength of 25 kV/cm and a duration of 200 ns delivered 25 times at 1 Hz repetition rates to achieve irreversible electroporation for cellular apoptosis [22,23,24]. A computational model of the cervical cell, which helps to estimate the strength and duration of the electric field in RE, was not utilized in the above studies.
Cervical cells are squamous epithelial cells. An electric impedance study was conducted on normal and cancerous cervical cells, in which a numerical study was carried out using a cuboid and the dimensions of the cervical cells were accurately measured. The precancerous stages are known as cervical intraepithelial neoplasia (CIN). The three stages of CIN are CIN1, CIN2, and CIN3 [25,26]. In another study, a squamous epithelium cell model was built using a scutoid, prism, and frustum [27,28,29]. Thus, through these studies, it can be stated that cervical cells are flat in nature. The dimensions of cervical cells play an important role in identifying cancerous cervical cells. The dimensional changes in these cells are also an important marker of cancer progression. Another remarkable change observed during cervical cancer stages is the change in the conductivity of the cytoplasm, which becomes more conductive by a factor of 1.8 due to the increased concentration of NaCl [30,31,32]. The cervical cells’ mechanical properties also change as the membrane becomes weak and the elasticity of the cells decreases [33,34,35]. The study and research of electroporation with current electroporators are costly, and the desired high-frequency range electric fields are not available for supra-electroporation [36,37,38]. Nanosecond Gaussian pulses can be used to electroporate cervical cells, as stated in [22,29]. Many pulse generators can be explored and built to deliver these Gaussian pulses. For example, in a previous study, a Marx generator was used to generate Gaussian pulses that were exposed to biological cells [31,33]. A DSRD diode with fast switching abilities has also been used to generate high-voltage and nanosecond Gaussian pulses [32,34,35].
The electrodeformation of cells during electroporation is studied to examine the cytoskeleton integrity of the cells, which is an important condition in RE [37,38,39]. The cells undergo cytolysis due to exposure to high electric pulses. The electrodeformation of biological cells has previously been studied, and multiphysics has been used to inspect cytolysis. Two-dimensional models representing the double-layer membrane as a viscoelastic layer have been studied for CHO cells, yielding accurate results [40]. In previous works, cervical cells underwent deformations during the application of electric fields of low and high frequencies, but a spherical-shaped cell was used in the computational work [41,42,43]. Cervical cells are squamous epithelial cells, and most works have used cuboids and scutoids to demonstrate the cell dimensions. In [36], cuboids were used to represent cells. The authors also studied the dimensions of the normal and cancerous cervical cells. In [41], the authors used a sphere as a 3D model of cervical cells to study electrodeformation. The work carried out in [36] used a frustum and scutoids to represent cell geometry.
Our model uses the original 2D morphology, extrudes the plasma membrane boundaries, and constructs a three-dimensional image. The extrusion depends on the thickness measurement provided by earlier works [25,26]. The nucleus is considered an ellipsoid, and the dimensions are created according to the literature. This model could be used in an initial electroporation study of cervical cells in model-driven microdosimetry. The method involves first extracting the image contour using image processing methods. The images of the normal and cancerous cervical cells are taken from the Herlev database [44]. The cells are then converted into three-dimensional solid images using CAD tools. We first attempt to verify the model with experimental work [21]. A 2.5 kV/cm electric field rectangular pulse of 100 μ s is applied to the computational domain, as used in the literature. The cervical cells are then exposed to electric fields with low- and high-frequency single pulses. The TMV, pore density, pore radius evaluation, strain energy, and solid displacement are calculated. The DSRD diode-based electrical circuit is used to generate high-voltage and high-frequency Gaussian pulses. The Pspice tool is then used to simulate the circuit, and the output pulse is then imported into COMSOL 5.5a. The Laplace equation and the Smoluchowiski equation are used to calculate the TMV and pore density. The Debye second-order dispersive equation is used during high-frequency pulses to demonstrate the dispersive nature of cervical cells. The Maxwell stress tensor (MST) is calculated for the cells. The solid displacement is calculated for the viscoelastic plasma membrane. The Young’s modulus of the cancerous cell lowers with cancer progression, thus causing weakening of the plasma membrane and cytoplasm and providing more displacement with applied electric force. The conductivity of cancerous cells is higher, thus causing higher and faster generation of the TMV and pores in the cell. This work, therefore, provides a computational model of normal and cancerous cervical cells and also provides a testing platform for the electroporator before the experimental work.

2. Materials and Methods

2.1. Three-Dimensional Construction of Normal and Cancerous Cervical Cells

The method described here involves three stages for obtaining the desired results in electroporation. The first stage consists of extracting the contours from the cervical cell images [44]. Figure 1 shows a flowchart of the method used to extract the cell contours and introduce them into COMSOL 5.5. The cervical cells used to demonstrate our work are intermediate (standard cells) and CIN2 (cancer cells). A filter and K-means clustering are used to denoise the images and separate the background from the cytoplasm and nucleus regions. Canny edge detection is used to extract the contours. CAD tools are used to reshape the cell into the size of an intermediate cervical cell. Then, a 3D model is created using extrusion, and an ellipsoid nucleus is placed at the center, using the dimensions presented in the literature [25,26]. Table 1 shows the dimensions of normal (intermediate) cervical cells and the dimensions of CIN1, CIN2, and CIN3 intermediate cells. The image of the CIN2-stage intermediate cells was obtained from the Herlev database, which contains models of cancer cells, and an intermediate normal cell was chosen for our work. Figure 2 shows the orientation of the applied electric field, along with the cancer cell model (Figure 2a) and the normal cell (Figure 2b). The computational domain has dimensions of 200 μ m × 100 μ m × 100 μ m. Here, P indicates the point where the analysis of electroporation was performed.

2.2. Electroporation Model of Cervical Cells

2.2.1. Modeling of Cervical Cell Permittivity

The dielectric properties of cervical cells are modeled using a multi-relaxation Debye-based relationship. The dielectric properties of the cervical cells use a dispersive medium, which is modeled using a second-order equation [8].
ϵ ( ω ) ϵ = Δ · ϵ s + j ω ( τ 1 Δ · ϵ 2 + τ 2 Δ · ϵ 1 ) 1 ω 2 τ p + j ω τ s ,
Δ ϵ s = Δ · ϵ 1 + Δ · ϵ 2 , τ p = τ 1 τ 2 , τ s = τ 1 + τ 2 ,
where ϵ denotes high-frequency permittivity, τ 1 and τ 2 are the relaxation times, and Δ ϵ 1 and Δ ϵ 2 are the relaxation amplitudes. The polarization P is expressed as a time-varying electric field and is given by:
τ p d 2 P d t 2 + τ s d P d t + P = k 2 d 2 E d t 2 + k 1 d E d t + k 0 E ,
k 0 = Δ · ϵ s + ϵ ϵ 0 , k 1 = τ 1 ( k 0 Δ · ϵ 1 ) + τ 2 ( k 0 Δ · ϵ 2 ) , k 2 = τ p ( k 0 Δ · ϵ 2 )

2.2.2. Cervical Cell Pore Formation during Electroporation

Pores with a radius of r p = 0.8 nm are established across the bi-lipid membrane layer due to the nanosecond pulsed electric field. The creation of pores is modeled using an asymptotic Smoluchowski equation:
N t = α e ( t m v V e p ) 2 ) [ 1 N N 0 e ( q ( t m v V e p ) 2 ) ] ,
where N is the pore density on the plasma membrane, V e p is the characteristic voltage of electroporation, α and q are the electroporation parameters, and N 0 is the initial pore density. The transmembrane potential increases as the average conductivity of the cell membrane increases with time, which is updated in the Smoluchowski equation at every step. It leads to an increase in the number of pores. The average conductivity at different pore formation regions is given by
σ ( x , y , t ) σ 0 = π σ p r p N ( x , y , t ) K c ,
where K c is a constant and can be calculated using Equation (6).
K c = e m υ 1 w 0 . e w 0 η . υ m w 0 η . υ m w 0 . e w 0 + η . υ m w 0 + η . υ m ,
where w 0 is the pore energy barrier, η is the pore relative entrance length, and υ m = q e K T . v m is the compensated transmembrane voltage.

2.2.3. Electromagnetic Modeling of Cervical Cells

The electric potential is determined by:
· ( σ + ϵ 0 . ϕ t ) · P t = 0 ,
E = · ϕ ,
The transmembrane voltage (TMV) is determined using a special boundary condition, and the cell membrane bi-lipid layer is replaced with a distributed impedance layer, which reduces computational time and space.
T M V = ϕ o u t e r ( x , y , t ) ϕ i n n e r ( x , y , t ) ,
n . J = σ m d m . ( V V r e f ) + ϵ m d m . ( V t V r e f t ) ,
where V r e f is the exterior surface potential, and V is the interior surface potential of the cell boundary. d m , σ m , and ϵ m are the thickness, conductivity, and permittivity of the cervical cell membrane. The transmembrane potential is updated at each time step. Smoluchowski’s equation is solved using the weak form of the PDE, and the polarization vector equation is solved using the coefficient form of the PDE interface. At each time step, P and σ m are updated. The 3D model is solved in an iterative manner. The number of generated finite elements is fewer in number. The development of the pore radius with respect to the time is given by Equations (11) and (12).
d r j d t = D k b T V m 2 F m a x 1 + r h / ( r + r t ) + 4 β r * r 4 1 r 2 π γ + 2 π N δ e f f ,
where V m is the generated transmembrane potential. The number of hydrophilic pores developed (r > r*) corresponds to j = 1, 2, 3, …, k p , where k b represents the Boltzmann constant. δ e f f is calculated as:
δ e f f = 2 δ ( 2 δ δ 0 ) ( 1 A p A ) 2 ,
where A is the area of the plasma membrane, and the total perforation area is given by A p , which is calculated as [7]:
A p = S N ( t ) π r j 2 d S ,
where S is the plasma membrane surface.

2.2.4. Electrodeformation Model of Cervical Cells

The electroporation electrical force is determined using the Maxwell stress tensor (MST):
M S T ϵ ( t ) = ( E i . E j δ i j E 2 2 ) ,
where E is the electric field in the x-, y-, and z-directions, ϵ ( t ) is the dielectric permittivity, and δ i j is the Knoreckner delta. We have assumed the generalized Maxwell model as the viscoelastic model. The electrical stress is coupled with the mechanical stress on the cervical cell. The structural model membrane displacement u is calculated. The membrane is modeled as a viscoelastic model and the remainder of the cell as an elastic model. The strain energy is calculated by first solving
ρ d 2 u d t 2 = δ × S ,
where S denotes the stress. After obtaining S, the strain energy of the deformable cell membrane is calculated as
E m e m = S × γ d Ω / 2 ,
where d Ω is the volume of the element and γ represents the strain.

2.2.5. Temperature Development during RE

The changes in the electroporation temperature are measured using the Pennes bioheat equation [45,46,47], defined as:
· ( k T ) + σ | ψ | 2 + q W b c b T = ρ c p T t ,
where T is the temperature (310.15 K); W b c b T and q are the heat generated in the bloodstream and metabolism, respectively; c p is the heat capacity; k is the thermal conductivity; ψ is the potential developed; and σ is the conductivity. This equation is used to observe whether the developed thermal stresses exceed the safety threshold. Table 2 contains the parameter values used in the computational study.

2.2.6. High-Frequency Electric Field Using DSRD Generator Design

The electric field is applied to the copper plates on the top, and the lower part is grounded. The electric field is generated using the simulation of the pulse generator using a drift-step rectifier diode (DSRD), as shown in the circuit diagram in Figure 3. Drift-step rectifier diodes are fast-switching high-voltage devices. In the initial stage, the MOSFET is turned on, allowing a current to flow through coil L 1 . The duration of the current flow is represented by Δ t , and its magnitude is determined by the input voltage V e e . The charging current passing through coil L 1 depends on both Δ t and V e e . Additionally, in the forward direction, the charge passing through the DSRD is influenced by Δ t , a voltage V f f , and a resistor R 1 . The MOSFET is subsequently switched off, causing the current to reverse its direction through coil L 1 and flow through the DSRD. The DSRD stays in a reverse conductive mode, allowing the current to flow in the reverse direction. Once the charge from the DSRD is depleted, it practically switches off, allowing the current to flow to the load ( R L ). The load resistance R L determines the magnitude of the current flowing through it. Here, V e e = 18 V and V f f = 11 V in the circuit.
The MOSFET (SPW20N60C3) is activated by the trigger pulse, remaining in an on state through the application of the pulse for a duration of Δ t = 50 ns. V e e is applied to determine the charging current passing through L 1 = 75 nH. The amount of charge passing through the DSRD is determined by Δ t, V f f , and R 1 = 20 Ω . When the MOSFET switches off, the resulting current flows in the reverse direction of the DSRD diode. The current flows until the DSRD is in reverse conducting mode. The removal of the charge causes the DSRD to turn off, allowing the current to flow through R L = 50 Ω . R 1 = 20 Ω plays an important role in balancing the charging of the DSRD in the forward direction. Diode D provides the fastest path to discharge C 1 = 100 pF after the MOSFET switches off. C 2 = 1 μ F and L 3 = 45 nH are used for impedance matching to the load [34]. The circuit has been simulated in the Pspice environment to obtain the resulting output, as shown in Figure 4A. Figure 4B shows the output response from the pulse generator captured in the digital oscilloscope. The second step is to take this output to the COMSOL 5.5 multiphysics.

3. Results

The cervical cells are imported into the multiphysics domain with dimensions of 210 μ m × 100 μ m × 100 μ m. The cells are first exposed to a 2.5 kV/cm electric field for a duration of 100 μ s. A 1 μ s rise and fall time is seen in the rectangular pulse. The copper electrodes are attached to the upper and lower cuboid volumes, as seen in Figure 2a,b. The electric field is applied from the top to the bottom plate, and the lower electrode is grounded. The temporal evolution of the TMV and pore density agrees with our computational work. The temporal pore radius evolution and strain energy density are obtained for a low-frequency pulse. The results agree with the experimental evaluation by reaching the threshold of electroporation. The required TMV (above 1 V) and pore density in log scale (above 14) are generated in normal and cancerous cervical cells, as shown in Figure 5a,b. The pore radius regains its original state after removing the rectangular pulse of 0.8 nm, as shown in Figure 5c. The strain energy ( E m e m ) that developed in the cell membrane due to the induced pulses is plotted in Figure 5d. The displacement of the cells is compared to the experimental study, and the deformation can be seen in Figure 5e.
The cervical cells are retested under a high-frequency electric field. A Gaussian pulse with a magnitude of 11.5 kV/cm for a duration of 40 ns generated using the above-mentioned method is transferred to COMSOL 5.5a. The temporal evolution of the TMV and pore density is calculated and plotted in Figure 6a,b. The TMV calculation demonstrates that a higher potential developed in cancerous cells compared to normal cells. Similar results can be seen for the temporal development of the pore density and the evolution of the pore radius in cancerous cells (Figure 6c). The development of the Maxwell stress tensor is greater in magnitude and faster in a cancerous cervical cell compared to a normal cervical cell (Figure 6d. Another important observation is that the cervical cells return to their normal state after removing electrical stress, confirming that the membrane integrity is intact. The solid displacement demonstrates this in the cervical cells during electroporation, as shown in Figure 6e. The strain energy ( E m e m ) is also calculated for the cervical cells, with cancerous cells developing a greater strain energy. Figure 6f shows the temporal development of the strain energy in the cervical cells. We also checked the increase in the model temperature to prove that RE is a non-thermal process. In Figure 6g, a negligible increase in the temperature can be seen as the initial temperature was maintained at 310 K. Figure 7a depicts a three-dimensional plot of the TMV generated on the surface of a normal cell at 20 ns by the Gaussian pulse when it reaches its peak value. Figure 7b shows the pore density developed at 20 ns. The pore density and TMV reach their maximum values at this time. Similarly, Figure 7b,d depict surface plots of the TMV and pore density generated in the cancerous cells at 20 ns.

4. Discussion

The TMV and pore density necessary for RE are generated. The pore radius returns to 0.8 nm, which proves that the model agrees with the results of previous works. Our study adds three-dimensional models of normal and cancerous cervical cells, and the conductivity changes are shown for all the cells. The cancerous cells show a faster generation of the TMV and pore density, as shown in Figure 6a,b, due to the decrease in their sizes, causing more electric fields to develop. This study uses the Pennes bioheat equation to prove that the process is non-thermal. Hence, factors such as the TMV (>1 V), pore density in log scale (>14), and pore radius return to their original size of 0.8 nm, and the non-thermal process is satisfied. Earlier RE numerical models studied the transmembrane potential, pore density, and evolution of the pore radius, which were calculated to explain the physics behind RE. Our work adds the checking of the cytoskeleton integrity of the cervical cells to the study of RE. The application of electric pulses caused a change in the dimensions of the cervical cells. Figure 6e also shows the cervical cells returning to their initial positions after nanosecond electric pulses, from which we observed that no cytolysis was found in the normal or cancerous cervical cells. This study also revealed that, due to differences in the size of the cells, the electrical stress on the cancerous cells was higher compared to the normal cells. Hence, cell dimensions play an important role in electroporation studies.
Our simulation revealed a maximum displacement of 72 nm in normal cervical cells, accompanied by an aspect ratio or deformation ratio change (displacement along the y-axis/displacement along the x-axis) of 1.0053. Figure 8a illustrates a comparison between our model and experimental findings, with a graph relating the applied electric field to the electrodeformation in cervical cells. Notably, our model closely aligns with the experimental results. Figure 8b represents the maximum displacement for the plasma membrane under a 1 kV/cm electric field applied for a duration of 1 μ s.
This work can be used for indicative purposes to determine the required strength and duration of electric fields that can be applied to cells. The results could be used in the future to determine whether it is possible to transfer drugs, dyes, and genes into cervical cells before resorting to trial-and-error methods and wasting cell samples.

5. Conclusions

This work extracts a two-dimensional image of a cervical cell using simple image processing methods and constructs a three-dimensional geometry of normal and cancerous cervical cells using CAD tools. When exposed to an electric pulse, the three-dimensional cervical cells generated different transmembrane potentials and pore densities due to changes in the dimensions and conductivities of normal and cancerous cells. The cancerous cells showed faster and greater development of the TMV and pore density. This study also examined the viscoelastic movement of the membranes of cervical cells. The elastic properties changed with cancer progression, causing greater solid displacement in cancerous cervical cells. This study provides more insight into cytoskeleton integrity. Thus, in the future, our results could be used to carry out RE experimental studies on cancerous cervical cells at high frequencies. This study could be helpful for diagnostic purposes by detecting cell morphological changes due to cancer progression at an early stage through the influx of dyes into cervical cells. The proposed method could also be used to test drugs and deliver genes into cells. The DSRD generator can potentially be used in the future to deliver dyes into cervical cells to observe morphological changes in the cells due to cancer progression.

Author Contributions

Conceptualization, M.K., S.K. and S.C.; formal analysis, M.K. and B.G.; investigation, M.K., A.P. and H.M.; writing—original draft preparation, M.K.; writing—review and editing, S.K., S.C. and A.P.; supervision, S.K. and S.C. All authors have read and agreed to the published version of the manuscript.

Funding

Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2024R137), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
TMVTransmembrane voltage
MSTMaxwell stress tensor
REReversible electroporation
DSRDDrift-step rectifier diode

References

  1. Miklavcic, D.; Puc, M. Electroporation, Wiley Encyclopedia of Biomedical Engineering; John Wiley: New York, NY, USA, 2006. [Google Scholar]
  2. Kotnik, T.; Miklavčič, D. Theoretical evaluation of voltage inducement on internal membranes of biological cells exposed to electric fields. Biophys. J. 2006, 90, 480–491. [Google Scholar] [CrossRef]
  3. Kotnik, T.; Rems, L.; Tarek, M.; Miklavčič, D. Membrane electroporation and electropermeabilization: Mechanisms and models. Ann. Rev. Biophys. 2019, 48, 63–91. [Google Scholar] [CrossRef]
  4. Yarmush, M.L.; Golberg, A.; Sersa, G.; Kotnik, T.; Miklavcic, D. Electroporation-based technologies for medicine: Princi- ples, applications and challenges. Ann. Rev. Biomed. Eng. 2014, 16, 295–320. [Google Scholar] [CrossRef]
  5. De Angelis, A.; Denzi, A.; Merla, C.; Andre, F.M.; Mir, L.M.; Apollonio, F.; Liberti, M. Confocal Microscopy Improves 3D Microdosimetry Applied to Nanoporation Experiments Targeting Endoplasmic Reticulum. Front. Bioeng. Biotechnol. 2020, 8, 552261. [Google Scholar] [CrossRef]
  6. Merla, C.; Paffi, A.; Apollonio, F.; Leveque, P.; d’Inzeo, G.; Liberti, M. Microdosimetry for Nanosecond Pulsed Electric Field Applications: A Parametric Study for a Single Cell. IEEE Trans. Biomed. Eng. 2011, 58, 1294–1302. [Google Scholar] [CrossRef]
  7. Krassowska, W.; Filev, P.D. Modeling electroporation in a single cell. Biophys. J. 2007, 92, 404–417. [Google Scholar] [CrossRef]
  8. Salimi, E.; Thomson, D.J.; Bridges, G.E. Membrane dielectric dispersion in nanosecond pulsed electroporation of biological cells. IEEE Trans. Dielectr. Electr. Insul. 2013, 20, 1256–1265. [Google Scholar] [CrossRef]
  9. Pucihar, G.; Miklavcic, D.; Kotnik, T. A time-dependent numerical model of transmembrane voltage inducement and electroporation of irregularly shaped cells. IEEE Trans. Biomed. Eng. 2009, 56, 1491–1501. [Google Scholar] [CrossRef]
  10. Mescia, L.; Chiapperino, M.A.; Bia, P.; Gielis, J.; Caratelli, D. Modeling of Electroporation Induced by Pulsed Electric Fields in Irregularly Shaped Cells. IEEE Trans. Biomed. Eng. 2018, 65, 414–423. [Google Scholar] [CrossRef]
  11. Chiapperino, M.A.; Bia, P.; Caratelli, D.; Gielis, J.; Mescia, L.; Dermol-Černe, J.; Miklavčič, D. Nonlinear Dispersive Model of Electroporation for Irregular Nucleated Cells. Bioelectromagnetics 2019, 40, 331–342. [Google Scholar] [CrossRef]
  12. Tang, Y.; Liu, S.; Deng, Y.; Zhang, Y.; Yin, L.; Zheng, W. An improved method for soft tissue modeling. Biomed. Signal Process. Control 2021, 65, 102367. [Google Scholar] [CrossRef]
  13. Zhang, Z.; Wang, L.; Zheng, W.; Yin, L.; Hu, R.; Yang, B. Endoscope image mosaic based on pyramid ORB. Biomed. Signal Process. Control 2022, 71, 103261. [Google Scholar] [CrossRef]
  14. Chiapperino, M.A.; Mescia, L.; Bia, P.; Starešinič, B.; Čemažar, M.; Novickij, V.; Tabašnikov, A.; Smith, S.; Dermol-Černe, J.; Miklavčič, D. Experimental and Numerical Study of Electroporation Induced by Long Monopolar and Short Bipolar Pulses on Realistic 3D Irregularly Shaped Cells. IEEE Trans. Biomed. Eng. 2020, 67, 2781–2788. [Google Scholar] [CrossRef]
  15. Lu, S.; Yang, B.; Xiao, Y.; Liu, S.; Liu, M.; Yin, L.; Zheng, W. Iterative reconstruction of low-dose CT based on differential sparse. Biomed. Signal Process. Control 2023, 79, 104204. [Google Scholar] [CrossRef]
  16. Lu, S.; Liu, S.; Hou, P.; Yang, B.; Liu, M.; Yin, L.; Zheng, W. Soft Tissue Feature Tracking Based on Deep Matching Network. Comput. Model. Eng. Sci. 2023, 136, 363–379. [Google Scholar] [CrossRef]
  17. Denzi, A.; Camera, F.; Merla, C.; Benassi, B.; Consales, C.; Paffi, A.; Apollonio, F.; Liberti, M. A Microdosimetric Study of Electropulsation on Multiple Realistically Shaped Cells: Effect of Neighbours. J. Membr. Biol. 2016, 249, 691–701. [Google Scholar] [CrossRef]
  18. Hu, B.; Das, P.; Lv, X.; Shi, M.; Aa, J.; Wang, K.; Duan, L.; Gilbert, J.A.; Nie, Y.; Wu, X.L. Effects of ’Healthy’ Fecal Microbiota Transplantation against the Deterioration of Depression in Fawn-Hooded Rats. mSystems 2022, 7, e21822. [Google Scholar] [CrossRef]
  19. Lu, S.; Yang, J.; Yang, B.; Yin, Z.; Liu, M.; Yin, L.; Zheng, W. Analysis and Design of Surgical Instrument Localization Algorithm. Comput. Model. Eng. Sci. 2023, 137, 669–685. [Google Scholar] [CrossRef]
  20. Kumar, M.; Mishra, A. Reversible electroporation study of realistic normal and cancerous cervical cells model using avalanche transistor-based nano pulse generator. IOP Biomed. Phys. Eng. Express 2021, 7, 065011. Available online: http://iopscience.iop.org/article/10.1088/2057-1976/ac240b (accessed on 12 March 2023). [CrossRef]
  21. Yang, Y.; Moser, M.A.; Zhang, E.; Zhang, W.; Zhang, B. Development of a statistical model for cervical cancer cell death with irreversible electroporation in vitro. PLoS ONE 2018, 13, e0195561. [Google Scholar] [CrossRef]
  22. Mitsutake, K.; Satoh, A.; Mine, S.; Abe, K.; Katsuki, S.; Akiyama, H. Effect of pulsing sequence of nanosecond pulsed electric fields on viability of HeLa S3 cells. IEEE Trans. Dielectr. Electr. Insul. 2012, 19, 337–342. [Google Scholar] [CrossRef]
  23. Zhu, Y.; Huang, R.; Wu, Z.; Song, S.; Cheng, L.; Zhu, R. Deep learning-based predictive identification of neural stem cell differentiation. Nat. Commun. 2021, 12, 2614. [Google Scholar] [CrossRef]
  24. Li, L.; Wang, S.; Zhou, W. Balance Cell Apoptosis and Pyroptosis of Caspase-3-Activating Chemotherapy for Better Antitumor Therapy. Cancers 2023, 15, 26. [Google Scholar] [CrossRef]
  25. Walker, D.C.; Brown, B.H.; Hose, D.R.; Smallwood, R.H. Modelling the electrical impedivity of normal and premalignant cervical tissue. Electron. Lett. 2000, 36, 1603–1604. [Google Scholar] [CrossRef]
  26. 1, D.C.W.; Brown, B.H.; Smallwood, R.H.; Hose, D.R.; Jones, D.M. Modelled current distribution in cervical squamous tissue. Physiol. Meas. 2002, 23, 159–168. [Google Scholar] [CrossRef]
  27. Gómez-Gálvez, P.; Vicente-Munuera, P.; Tagua, A.; Forja, C.; Castro, A.M.; Letrán, M.; Valencia-Expósito, A.; Grima, C.; Bermúdez-Gallardo, M.; Serrano-Pérez-Higueras, Ó.; et al. Scutoids are a geometrical solution to three-dimensional packing of epithelia. Nat. Commun. 2018, 9, 2960. [Google Scholar] [CrossRef]
  28. Chang, Q.Q.; Chen, C.Y.; Chen, Z.; Chang, S. LncRNA PVT1 promotes proliferation and invasion through enhancing Smad3 expression by sponging miR-140-5p in cervical cancer. Radiol. Oncol. 2019, 53, 443–452. [Google Scholar] [CrossRef]
  29. Chen, J.; Li, X.; Liu, H.; Zhong, D.; Yin, K.; Li, Y.; Zhu, L.; Xu, C.; Li, M.; Wang, C. Bone marrow stromal cell-derived exosomal circular RNA improves diabetic foot ulcer wound healing by activating the nuclear factor erythroid 2-related factor 2 pathway and inhibiting ferroptosis. Diabet. Med. 2023, 40, e15031. [Google Scholar] [CrossRef]
  30. Balidemaj, E.; De Boer, P.; Van Lier, A.L.; Remis, R.F.; Stalpers, L.J.; Westerveld, G.H.; Nederveen, A.J.; van den Berg, C.A.; Crezee, J. In vivo electric conductivity of cervical cancer patients based on B maps at 3T MRI. Phys. Med. Biol. 2016, 61, 1596–1607. [Google Scholar] [CrossRef]
  31. Tang, L.; Li, J.; Bao, M.; Xiang, J.; Chen, Y.; Wang, Y. Genetic association between HER2 and ESR2 polymorphisms and ovarian cancer: A meta-analysis. Oncotargets Ther. 2018, 11, 1055–1066. [Google Scholar] [CrossRef]
  32. Mao, X.; Chen, Y.; Lu, X.; Jin, S.; Jiang, P.; Deng, Z.; Zhu, X.; Cai, Q.; Wu, C.; Kang, S. Tissue resident memory T cells are enriched and dysfunctional in effusion of patients with malignant tumor. J. Cancer 2023, 14, 1223–1231. [Google Scholar] [CrossRef]
  33. Heeren, T.; Ueno, T.; Wang, D.; Namihira, T.; Katsuki, S.; Akiyama, H. Novel dual Marx Generator for microplasma applications. IEEE Trans. Plasma Sci. 2005, 33, 1205–1209. [Google Scholar] [CrossRef]
  34. Merensky, L.M. Fast switching of drift step recovery diodes based on all epi-Si growth. In Proceedings of the 2009 IEEE International Conference on Microwaves, Communications, Antennas and Electronics Systems, Tel Aviv, Israel, 4–6 November 2009; pp. 1–4. [Google Scholar] [CrossRef]
  35. Kesar, A.S.; Sharabani, Y.; Merensky, L.M.; Shafir, I.; Sher, A. Drift-Step-Recovery Diode Characterization by a Bipolar Pulsed Power Circuit. IEEE Trans. Plasma Sci. 2012, 40, 3100–3104. [Google Scholar] [CrossRef]
  36. Geng, L.; Feng, J.; Sun, Q.; Liu, J.; Hua, W.; Li, J.; Ao, Z.; You, K.; Guo, Y.; Liao, F.; et al. Nanomechanical clues from morphologically normal cervical squamous cells could improve cervical cancer screening. Nanoscale 2015, 7, 15589. [Google Scholar] [CrossRef]
  37. Li, Z.; Wang, S.; He, Y.; Li, Q.; Gao, G.; Tong, G. Regulation of Apelin-13 on Bcl-2 and Caspase-3 and Its Effects on Adipocyte Apoptosis. Evid.-Based Complement. Altern. Med. 2021, 2021, 1687919. [Google Scholar] [CrossRef]
  38. Chen, S.; Chen, Y.; Yu, L.; Hu, X. Overexpression of SOCS4 inhibits proliferation and migration of cervical cancer cells by regulating JAK1/STAT3 signaling pathway. Eur. J. Gynaecol. Oncol. 2021, 42, 554–560. [Google Scholar] [CrossRef]
  39. Shamoon, D.; Dermol-Černe, J.; Rems, L.; Reberšek, M.; Kotnik, T.; Lasquellec, S.; Brosseau, C.; Miklavčič, D. Assessing the electro-deformation and electro-poration of biological cells using a three-dimensional finite element model. Appl. Phys. Lett. 2019, 114, 063701. [Google Scholar] [CrossRef]
  40. Guo, F.; Qian, K.; Deng, H.; Li, X. Multiphysics analysis of nsPEF induced electrodeformation in a dispersive cell model. Appl. Phys. Lett. 2021, 118, 083701. [Google Scholar] [CrossRef]
  41. Chen, J.; Abdelgawad, M.; Yu, L.; Shakiba, N.; Chien, W.-Y.; Lu, Z.; Geddie, W.R.; Jewett, M.A.S.; Sun, Y. Electrodeformation for single cell mechanical characterization. In Proceedings of the 2011 IEEE 24th International Conference on Micro Electro Mechanical Systems, Cancun, Mexico, 23–27 January 2011; pp. 1119–1122. [Google Scholar] [CrossRef]
  42. Xu, H.; Wang, H.; Zhao, W.; Fu, S.; Li, Y.; Ni, W.; Xin, Y.; Li, W.; Yang, C.; Bai, Y.; et al. SUMO1 modification of methyltransferase-like 3 promotes tumor progression via regulating Snail mRNA homeostasis in hepatocellular carcinoma. Theranostics 2020, 10, 5671–5686. [Google Scholar] [CrossRef]
  43. He, B.; Zhang, Y.; Zhou, Z.; Wang, B.; Liang, Y.; Lang, J.; Lin, H.; Bing, P.; Yu, L.; Sun, D.; et al. A Neural Network Framework for Predicting the Tissue-of-Origin of 15 Common Cancer Types Based on RNA-Seq Data. Front. Bioeng. Biotechnol. 2020, 8, 737. [Google Scholar] [CrossRef]
  44. Available online: http://mde-lab.aegean.gr/index.php/downloads (accessed on 10 March 2023).
  45. Khorasani, A. The Effect of Conductivity Changes on Temperature Rise During Irreversible Electroporation. Front. Biomed. Technol. 2020, 7, 178–185. [Google Scholar] [CrossRef]
  46. Davalos, R.V.; Mir, L.M.; Rubinsky, B. Tissue Ablation with Irreversible Electroporation. Ann. Biomed. Eng. 2005, 33, 223–231. [Google Scholar] [CrossRef]
  47. Huang, H.; Zhang, B.; Zhong, J.; Han, G.; Zhang, J.; Zhou, H.; Mao, T.; Liu, Y. The behavior between fluid and structure from coupling system of bile, bile duct, and polydioxanone biliary stent: A numerical method. Med. Eng. Phys. 2023, 113, 103966. [Google Scholar] [CrossRef]
Figure 1. Flowchart for the extraction of contours from the Pap smear images.
Figure 1. Flowchart for the extraction of contours from the Pap smear images.
Micromachines 14 02136 g001
Figure 2. Three-dimensional view of cervical cells in a multiphysics environment, with copper plates on the uppermost plate and the lower plate grounded: (a) Containing cancerous cervical cells and (b) containing normal cervical cells.
Figure 2. Three-dimensional view of cervical cells in a multiphysics environment, with copper plates on the uppermost plate and the lower plate grounded: (a) Containing cancerous cervical cells and (b) containing normal cervical cells.
Micromachines 14 02136 g002
Figure 3. Pulse generator circuit diagram using DSRD diode [34].
Figure 3. Pulse generator circuit diagram using DSRD diode [34].
Micromachines 14 02136 g003
Figure 4. (A) Simulation output plot of the DSRD pulse generator. (B) Pulse generator output pulse using DSRD diode.
Figure 4. (A) Simulation output plot of the DSRD pulse generator. (B) Pulse generator output pulse using DSRD diode.
Micromachines 14 02136 g004
Figure 5. (a) Temporal plot of the transmembrane potential of cervical cells at point P. (b) Temporal pore density (m 2 ) in log scale at point P in cervical cells. (c) Temporal plot of the evolution of the pore radius (nm) of cervical cells during RE. (d) Temporal plot of the strain energy of cervical cells during RE. (e) Solid displacement of cervical cells during electroporation.
Figure 5. (a) Temporal plot of the transmembrane potential of cervical cells at point P. (b) Temporal pore density (m 2 ) in log scale at point P in cervical cells. (c) Temporal plot of the evolution of the pore radius (nm) of cervical cells during RE. (d) Temporal plot of the strain energy of cervical cells during RE. (e) Solid displacement of cervical cells during electroporation.
Micromachines 14 02136 g005
Figure 6. (a) Temporal plot of the transmembrane potential of cervical cells at point P. (b) Temporal pore density (m 2 ) in log-scale at point P in cervical cells. (c) Temporal plot of the evolution of the pore radius (nm) of cervical cells during RE. (d) Temporal plot of the Maxwell stress tensor (N/m 2 ) of cervical cells at point P during RE. (e) Solid displacement of the cervical cells during electroporation. (f) Temporal plot of the strain energy of cervical cells during RE. (g) Temporal plot of the temperature increase during electroporation.
Figure 6. (a) Temporal plot of the transmembrane potential of cervical cells at point P. (b) Temporal pore density (m 2 ) in log-scale at point P in cervical cells. (c) Temporal plot of the evolution of the pore radius (nm) of cervical cells during RE. (d) Temporal plot of the Maxwell stress tensor (N/m 2 ) of cervical cells at point P during RE. (e) Solid displacement of the cervical cells during electroporation. (f) Temporal plot of the strain energy of cervical cells during RE. (g) Temporal plot of the temperature increase during electroporation.
Micromachines 14 02136 g006
Figure 7. (a) Plot of the TMV (V) at 20 ns for normal cervical cells. (b) Plot of the pore density in log scale at 20 ns for normal cervical cells. (c) Plot of the TMV (V) at 20 ns for cancerous cervical cells. (d) Plot of the pore density in log scale at 20 ns for cancerous cervical cells.
Figure 7. (a) Plot of the TMV (V) at 20 ns for normal cervical cells. (b) Plot of the pore density in log scale at 20 ns for normal cervical cells. (c) Plot of the TMV (V) at 20 ns for cancerous cervical cells. (d) Plot of the pore density in log scale at 20 ns for cancerous cervical cells.
Micromachines 14 02136 g007
Figure 8. (a) Comparison between experimental [41] and simulation results of cervical cells under different electric fields applied for a duration of 1 μ s. (b) Maximum displacement for the plasma membrane under a 1 kV/cm electric field applied for a duration of 1 μ s.
Figure 8. (a) Comparison between experimental [41] and simulation results of cervical cells under different electric fields applied for a duration of 1 μ s. (b) Maximum displacement for the plasma membrane under a 1 kV/cm electric field applied for a duration of 1 μ s.
Micromachines 14 02136 g008
Table 1. Dimensions of cervical cells.
Table 1. Dimensions of cervical cells.
Cell TypeCellNucleus
XYZXYZ
Intermediate52 ± 10 μ m52 ± 14 μ m15 ± 1.1 μ m8.4 ± 1.1 μ m8.4 ± 1.1 μ m5.8 ± 1.4 μ m
CIN141 ± 10 μ m41 ± 10 μ m14 ± 1.2 μ m7.3 ± 2.2 μ m7.3 ± 2.1 μ m5.4 ± 3.3 μ m
CIN225 ± 6 μ m25 ± 6 μ m17 ± 1.0 μ m8.1 ± 1.7 μ m8.1 ± 1.7 μ m6.9 ± 1.7 μ m
CIN318 ± 3.2 μ m18 ± 3.2 μ m16 ± 1.8 μ m9.3 ± 1.7 μ m8.6 ± 1.7 μ m8.6 ± 1.7 μ m
Table 2. Supplementary data.
Table 2. Supplementary data.
ϵ 13.9 pF m 1 High-frequency permittivity [11]
ϵ 0 8.85 pF m 1 Dielectric permittivity of vacuum [11]
ϵ r e x 72Relative permittivity of E x medium [25,26]
ϵ r C p 86Relative permittivity of C p medium [25,26]
ϵ r N p 145Relative permittivity of N p medium [25,26]
σ E x 1.2 S m 1 Conductivity of E x medium [25,26]
σ 0 9.5 nS m 1 Passive conductivity of P m [11]
σ C p 0.6 S m 1 Conductivity of C p [25,26]
σ N p 0.8 S m 1 Conductivity of N p [25,26]
q2.46Electroporation constant [11]
V e q 224 mVCharacteristic voltage of electroporation [11]
α 10 9   m 2 s 1 Pore-creation density [11]
N e q 3.3 × 10 6 m 2 Equilibrium pore density [11]
h5 nmPlasma membrane thickness [40]
w 0 3.2Energy barrier inside the pore [40]
T295 KTemperature
η 0.15Relative length of pore entrance area [11]
ρ 1050Uniform density for all regions [40]
Y e x 1000 PaYoung’s modulus for extracellular medium [40]
Y c 500 PaYoung’s modulus for normal cell cytoplasm [36]
Y m e m 500 PaYoung’s modulus for normal cell plasma membrane [36]
Y c 300 PaYoung’s modulus for cancer cell cytoplasm [36]
Y m e m 300 PaYoung’s modulus for cancer cell plasma membrane [36]
v e x 0.4Poisson ratio for extracellular medium [40]
v c 0.4Poisson ratio for cytoplasm [40]
v m e m 0.4Poisson ratio for plasma membrane [40]
S m e m 1500 PaBranch shear modulus of membrane [40]
t0.1 sBranch viscous relaxation time of membrane [40]
K E 0.41 W/mKThermal conductivity of extracellular fluid [45]
C E 3780 J/kg-KHeat capacity at a constant pressure of the extracellular fluid [45]
K C 0.6 W/mKThermal conductivity of the cytoplasm [45]
C C 4718 J/kg-KHeat capacity at a constant pressure of the cytoplasm [45]
K N 0.3 W/mKThermal conductivity of the nucleus [45]
C N 3000 J/kg-KHeat capacity at a constant pressure of the nucleus [45]
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

Kumar, M.; Kumar, S.; Chakrabartty, S.; Poulose, A.; Mostafa, H.; Goyal, B. Dispersive Modeling of Normal and Cancerous Cervical Cell Responses to Nanosecond Electric Fields in Reversible Electroporation Using a Drift-Step Rectifier Diode Generator. Micromachines 2023, 14, 2136. https://doi.org/10.3390/mi14122136

AMA Style

Kumar M, Kumar S, Chakrabartty S, Poulose A, Mostafa H, Goyal B. Dispersive Modeling of Normal and Cancerous Cervical Cell Responses to Nanosecond Electric Fields in Reversible Electroporation Using a Drift-Step Rectifier Diode Generator. Micromachines. 2023; 14(12):2136. https://doi.org/10.3390/mi14122136

Chicago/Turabian Style

Kumar, Mayank, Sachin Kumar, Shubhro Chakrabartty, Alwin Poulose, Hala Mostafa, and Bhawna Goyal. 2023. "Dispersive Modeling of Normal and Cancerous Cervical Cell Responses to Nanosecond Electric Fields in Reversible Electroporation Using a Drift-Step Rectifier Diode Generator" Micromachines 14, no. 12: 2136. https://doi.org/10.3390/mi14122136

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