Eur J Appl Physiol. 2007 Nov;101(4):495-502. Epub 2007 Aug 3.
Short-term effects of pulsed electromagnetic fields after physical exercise are dependent on autonomic tone before exposure.
Grote V, Lackner H, Kelz C, Trapp M, Aichinger F, Puff H, Moser M.
Institute of Noninvasive Diagnosis, JOANNEUM RESEARCH, Weiz, Austria.
Abstract
The therapeutic application of pulsed electromagnetic fields (PEMFs)
can accelerate healing after bone fractures and also alleviate pain
according to several studies. However, no objective criteria have been
available to ensure appropriate magnetic field strength or type of
electromagnetic field. Moreover, few studies so far have investigated
the physical principles responsible for the impact of electromagnetic
fields on the human body. Existing studies have shown that PEMFs
influence cell activity, the autonomic nervous system and the blood
flow. The aim of this study is to examine the instantaneous and
short-term effects of a PEMF therapy and to measure the impact of
different electromagnetic field strengths on a range of physiological
parameters, especially the autonomic nervous systems, determined by
heart rate variability (HRV) as well as their influence on subjects’
general feeling of well-being. The study comprised experimental,
double-blind laboratory tests during which 32 healthy male adults (age:
38.4+/-6.5 years) underwent four physical stress tests at standardised
times followed by exposure to pulsed magnetic fields of varying
intensity [HPM, High Performance magnetic field; Leotec; pulsed signal;
mean intensity increase: zero (placebo), 0.005, 0.03 and 0.09 T/s].
Exposure to electromagnetic fields after standardised physical effort
significantly affected the very low frequency power spectral components
of HRV (VLF; an indicator for sympathetically controlled blood flow
rhythms). Compared to placebo treatment, exposure to 0.005 T/s resulted
in accelerated recovery after physical strain. Subjects with lower
baseline VLF power recovered more quickly than subjects with higher VLF
when exposed to higher magnetic field strengths. The application of
electromagnetic fields had no effect on subjects’ general feeling of
well-being. Once the magnetic field exposure was stopped, the described
effects quickly subsided. PEMF exposure has a short-term
dosage-dependent impact on healthy subjects. Exposure to PEMF for 20 min
resulted in more rapid recovery of heart rate variability, especially
in the very low frequency range after physical strain. The study also
showed the moderating influence of the subjects’ constitutional VLF
power on their response to PEMF treatment. These findings have since
been replicated in a clinical study and should be taken into
consideration when PEMF treatment is chosen.
Bioelectromagnetics. 2007 Jan;28(1):64-8.
A pilot investigation of the effect of extremely low frequency pulsed electromagnetic fields on humans’ heart rate variability.
Baldi E, Baldi C, Lithgow BJ.
Diagnostic and Neurosignal Processing Research Group, Electrical
& Computer System Engineering, Monash University, Victoria,
Australia. Emilio.Baldi@eng.monash.edu.au
Abstract
The question whether pulsed electromagnetic field (PEMF) can affect
the heart rhythm is still controversial. This study investigates the
effects on the cardiocirculatory system of ELF-PEMFs. It is a follow-up
to an investigation made of the possible therapeutic effect ELF-PEMFs,
using a commercially available magneto therapeutic unit, had on soft
tissue injury repair in humans. Modulation of heart rate (HR) or heart
rate variability (HRV) can be detected from changes in periodicity of
the R-R interval and/or from changes in the numbers of heart-beat/min
(bpm), however, R-R interval analysis gives only a quantitative insight
into HRV. A qualitative understanding of HRV can be obtained considering
the power spectral density (PSD) of the R-R intervals Fourier
transform. In this study PSD is the investigative tool used, more
specifically the low frequency (LF) PSD and high frequency (HF) PSD
ratio (LF/HF) which is an indicator of sympatho-vagal balance. To obtain
the PSD value, variations of the R-R time intervals were evaluated from
a continuously recorded ECG. The results show a HR variation in all the
subjects when they are exposed to the same ELF-PEMF. This variation can
be detected by observing the change in the sympatho-vagal equilibrium,
which is an indicator of modulation of heart activity. Variation of the
LF/HF PSD ratio mainly occurs at transition times from exposure to
nonexposure, or vice versa. Also of interest are the results obtained
during the exposure of one subject to a range of different ELF-PEMFs.
This pilot study suggests that a full investigation into the effect of
ELF-PEMFs on the cardiovascular system is justified.
Magn Reson Med. 2003 Dec;50(6):1180-8.
Influence of magnetically-induced E-fields on cardiac electric activity during MRI: A modeling study.
Liu F, Xia L, Crozier S.
School of Information Technology and Electrical Engineering, University of Queensland, St. Lucia, Brisbane, Australia.
Abstract
In modern magnetic resonance imaging (MRI), patients are exposed to
strong, time-varying gradient magnetic fields that may be able to induce
electric fields (E-fields)/currents in tissues approaching the level of
physiological significance. In this work we present theoretical
investigations into induced E-fields in the thorax, and evaluate their
potential influence on cardiac electric activity under the assumption
that the sites of maximum E-field correspond to the myocardial
stimulation threshold (an abnormal circumstance). Whole-body cylindrical
and planar gradient coils were included in the model. The calculations
of the induced fields are based on an efficient, quasi-static,
finite-difference scheme and an anatomically realistic, whole-body
model. The potential for cardiac stimulation was evaluated using an
electrical model of the heart. Twelve-lead electrocardiogram (ECG)
signals were simulated and inspected for arrhythmias caused by the
applied fields for both healthy and diseased hearts. The simulations
show that the shape of the thorax and the conductive paths significantly
influence induced E-fields. In healthy patients, these fields are not
sufficient to elicit serious arrhythmias with the use of contemporary
gradient sets. However, raising the strength and number of repeated
switching episodes of gradients, as is certainly possible in local chest
gradient sets, could expose patients to increased risk. For patients
with cardiac disease, the risk factors are elevated. By the use of this
model, the sensitivity of cardiac pathologies, such as abnormal
conductive pathways, to the induced fields generated by an MRI sequence
can be investigated.
Auton Neurosci. 2003 Apr 30;105(1):53-61.
Can extremely low frequency alternating magnetic fields modulate heart rate or its variability in humans?
Kurokawa Y, Nitta H, Imai H, Kabuto M.
Environmental Health Science Region, National Institute for
Environmental Studies, 16-2 Onogawa, Ibaraki Tsukuba 305-0053, Japan. kurokawa@nies.go.jp
Abstract
This study is a reexamination of the possibility that exposure to
extremely low frequency alternating magnetic field (ELF-MF) may
influence heart rate (HR) or its variability (HRV) in humans. In a
wooden room (cube with 2.7-m sides) surrounded with wire, three series
of experiments were performed on 50 healthy volunteers, who were exposed
to MFs at frequencies ranging from 50 to 1000 Hz and with flux
densities ranging from 20 to 100 microT for periods ranging from 2 min
to 12 h. In each experiment, six indices of HR/HRV were calculated from
the RR intervals (RRIs): average RRI, standard deviation of RRIs, power
spectral components in three frequency ranges (pVLF, pLF and pHF), and
the ratio of pLF to pHF. Statistical analyses of results revealed no
significant effect of ELF-MFs in any of the experiments, and suggested
that the ELF-MF to which humans are exposed in their daily lives has no
acute influence on the activity of the cardiovascular autonomic nervous
system (ANS) that modulates the heart rate.
Neuropsychobiology. 1998 Nov;38(4):251-6.
No effects of pulsed high-frequency electromagnetic fields on heart rate variability during human sleep.
Mann K, Roschke J, Connemann B, Beta H.
Department of Psychiatry, University of Mainz, Germany.
The influence of pulsed high-frequency electromagnetic fields emitted
by digital mobile radio telephones on heart rate during sleep in
healthy humans was investigated. Beside mean RR interval and total
variability of RR intervals based on calculation of the standard
deviation, heart rate variability was assessed in the frequency domain
by spectral power analysis providing information about the balance
between the two branches of the autonomic nervous system. For most
parameters, significant differences between different sleep stages were
found. In particular, slow-wave sleep was characterized by a low ratio
of low- and high-frequency components, indicating a predominance of the
parasympathetic over the sympathetic tone. In contrast, during REM sleep
the autonomic balance was shifted in favor of the sympathetic activity.
For all heart rate parameters, no significant effects were detected
under exposure to the field compared to placebo condition. Thus, under
the given experimental conditions, autonomic control of heart rate was
not affected by weak-pulsed high-frequency electromagnetic fields.
Bioelectromagnetics. 1998;19(2):98-106.
Nocturnal exposure to intermittent 60 Hz magnetic fields alters human cardiac rhythm.
Heart rate variability (HRV) results from the action of neuronal and
cardiovascular reflexes, including those involved in the control of
temperature, blood pressure and respiration. Quantitative spectral
analyses of alterations in HRV using the digital Fourier transform
technique provide useful in vivo indicators of beat-to-beat variations
in sympathetic and parasympathetic nerve activity. Recently, decreases
in HRV have been shown to have clinical value in the prediction of
cardiovascular morbidity and mortality. While previous studies have
shown that exposure to power-frequency electric and magnetic fields
alters mean heart rate, the studies reported here are the first to
examine effects of exposure on HRV. This report describes three
double-blind studies involving a total of 77 human volunteers. In the
first two studies, nocturnal exposure to an intermittent, circularly
polarized magnetic field at 200 mG significantly reduced HRV in the
spectral band associated with temperature and blood pressure control
mechanisms (P = 0.035 and P = 0.02), and increased variability in the
spectral band associated with respiration (P = 0.06 and P = 0.008). In
the third study the field was presented continuously rather than
intermittently, and no significant effects on HRV were found. The
changes seen as a function of intermittent magnetic field exposure are
similar, but not identical, to those reported as predictive of
cardiovascular morbidity and mortality. Furthermore, the changes
resemble those reported during stage II sleep. Further research will be
required to determine whether exposure to magnetic fields alters stage
II sleep and to define further the anatomical structures where
field-related interactions between magnetic fields and human physiology
should be sought.
Med Biol Eng Comput. 1992 Mar;30(2):162-8.
Closed-chest cardiac stimulation with a pulsed magnetic field.
Hillenbrand Biomedical Engineering Center, School of Electrical Engineering, Purdue University, West Lafayette, IN 47907.
Abstract
Magnetic stimulators, used medically, generate intense rapidly
changing magnetic fields, capable of stimulating nerves. Advanced
magnetic resonance imaging systems employ stronger and more rapidly
changing gradient fields than those used previously. The risk of
provoking cardiac arrhythmias by these new devices is of concern. In the
paper, the threshold for cardiac stimulation by an externally-applied
magnetic field is determined for 11 anaesthetised dogs. Two coplanar
coils provide the pulsed magnetic field. An average energy of
approximately 12 kJ is required to achieve closed-chest magnetically
induced ectopic beats in the 17-26 kg dogs. The mean peak induced
electric field for threshold stimulation is 213 V m-1 for a 571
microseconds damped sine wave pulse. Accounting for waveform efficacy
and extrapolating to long-duration pulses, a threshold induced electric
field strength of approximately 30 V m-1 for the rectangular pulse is
predicted. It is now possible to establish the margin of safety for
devices that use pulsed magnetic fields and to design therapeutic
devices employing magnetic fields to stimulate the heart.
Bioelectromagnetics. 1992;13(4):303-11.
Preliminary report: modification of cardiac contraction rate by pulsed magnetic fields.
Ramon C, Powell MR.
Institute of Applied Physiology and Medicine, Seattle, WA 98122.
Abstract
Isolated rat hearts and excised canine cardiac tissues were subjected
to pulsed magnetic fields. The fields excited in coils by tandem
pairings of sinusoidal pulses were presented at various inter-pair
delays and repetition rates. The waveform of the magnetic field was a
single or multiple sinusoid followed after a variable delay by another
single or multiple sinusoid. Small but reliable increases in the beating
rate of rat heart were observed. Similar increases occurred in
contraction rates of canine tissues. Both preparations exhibited a
contraction-rate dependency on the repetition rate of the paired
magnetic pulses: 4.5-6 rep/s for canine tissue, and 20-25 and 40-55
reps/s for rat heart. Flux-density thresholds for both preparations
approximated 10 mT (100 gauss) rms.
Extracorporeal pulsed electromagnetic
field (PEMF) has shown the ability to regenerate tissue by promoting
cell proliferation. In the present study, we investigated for the first
time whether PEMF treatment could improve the myocardial
ischaemia/reperfusion (I/R) injury and uncovered its underlying
mechanisms.
In our study, we demonstrated for the
first time that extracorporeal PEMF has a novel effect on myocardial I/R
injury. The number and function of circulating endothelial progenitor
cells (EPCs) were increased in PEMF treating rats. The in vivo results
showed that per-treatment of PEMF could significantly improve the
cardiac function in I/R injury group. In addition, PEMF treatment also
reduced the apoptosis of myocardial cells by up-regulating the
expression of anti-apoptosis protein B-cell lymphoma 2 (Bcl-2) and
down-regulating the expression of pro-apoptosis protein (Bax). In vitro,
the results showed that PEMF treatment could significantly reduce the
apoptosis and reactive oxygen species (ROS) levels in primary neonatal
rat cardiac ventricular myocytes (NRCMs) induced by
hypoxia/reoxygenation (H/R). In particular, PEMF increased the
phosphorylation of protein kinase B (Akt) and endothelial nitric oxide
synthase (eNOS), which might be closely related to attenuated cell
apoptosis by increasing the releasing of nitric oxide (NO). Therefore,
our data indicated that PEMF could be a potential candidate for I/R
injury.Keywords: apoptosis, Bax, B-cell lymphoma 2 (Bcl-2), ischaemia/reperfusion (I/R) injury, pulsed electromagnetic field (PEMF)
INTRODUCTION
Hypertension, arrhythmia, myocardial
infarction (MI) and myocardial ischaemia/reperfusion (I/R) injury are
all the most common cardiac diseases, which are the major causes of
mortality in the world [1].
Among them, myocardial I/R injury is the most important cause of
cardiac damage. Its pathological process is closely related to
postoperative complications [2,3]
caused by coronary artery vascular formation, coronary
revascularization and heart transplantation. After myocardium suffered
severe ischaemia, restoration of the blood flow is a prerequisite for
myocardial salvage [2]. However, reperfusion may induce oxidative stress [4], inflammatory cell infiltration and calcium dysregulation [5]. All these players contribute to the heart damage such as contraction and arrhythmias [6], generally named myocardial I/R injury. Recently, more and more evolving therapies have been put into use for I/R injury.
Pulsed electromagnetic field (PEMF) is the most widely
tested and investigated technique in the various forms of
electromagnetic stimulations for wound healing [7], alleviating traumatic pain and neuronal regeneration [8,9].
The rats were randomly divided into PEMF-treated (5 mT, 25 Hz, 1 h
daily) and control groups. They hypothesized the possible mechanism that
PEMF would increase the myofibroblast population, contributing to wound
closure during diabetic wound healing. It is a non-invasive and
non-pharmacological intervention therapy. Recent studies indicated that
PEMF also stimulated angiogenesis in patients with diabetes [10], and could improve arrhythmia, hypertension and MI [1]. The MI rats were exposed to active PEMF for 4 cycles per day (8 min/cycle, 30±3 Hz, 6 mT) after MI induction. In vitro,
PEMF induced the degree of human umbilical venous endothelial cells
tubulization and increased soluble pro-angiogenic factor secretion [VEGF
and nitric oxide (NO)] [7].
However, the role of PEMF in ischaemia and reperfusion diseases remains
largely unknown. Our study aimed to investigate the effects of PEMF
preconditioning on myocardial I/R injury and to investigate the involved
mechanisms.
In our study, we verified the
cardioprotective effects of PEMF in myocardial I/R rats and the
anti-apoptotic effects of PEMF in neonatal rat cardiac ventricular
myocytes (NRCMs) subjected to hypoxia/reoxygenation (H/R). We
hypothesized that PEMF treatment could alleviate myocardial I/R injury
through elevating the protein expression of B-cell lymphoma 2 (Bcl-2),
phosphorylation of protein kinase B (Akt). Meanwhile, it could decrease
Bax. We emphatically made an effort to investigate the MI/R model and
tried to uncover the underlying mechanisms.
MATERIALS AND METHODS
Animals
Male, 12-week-old Sprague Dawley
(SD) rats (250–300 g) were purchased from Shanghai SLAC Laboratory
Animal. Animals were housed in an environmentally controlled breeding
room and given free access to food and water supplies. All animals were
handled according to the “Guide for the Care and Use of Laboratory
Animals” published by the US National Institutes of Health (NIH).
Experimental procedures were managed according to the Institutional
Aminal Care and Use Committee (IACUC), School of Pharmacy, Fudan
University.
The measurement of blood pressure in SHR rats
At the end of 1 week treatment with
PEMF, the rats were anesthetized with chloral hydrate (350 mg/kg, i.p.),
the right common carotid artery (CCA) was cannulated with polyethylene
tubing for recording of the left ventricle pressures (MFlab 200, AMP
20130830, Image analysis system of physiology and pathology of Fudan
University, Shanghai, China).
Myocardial I/R injury rat model and measurement of infarct size
All the rats were divided into
three groups: (1) Sham: The silk was put under the left anterior
descending (LAD) without ligation; (2) I/R: Hearts were subjected to
ischaemia for 45 min and then reperfusion for 4 h; (3) I/R + PEMF: PEMF
device was provided by Biomobie Regenerative Medicine Technology. The
I/R rats were pre-exposed to active PEMF for 2 cycles per day (8 min per
cycle), whereas other two groups were housed with inactive PEMF
generator. I/R was performed by temporary ligation of the LAD coronary
artery for 45 min through an incision in the fourth intercostal space
under anaesthesia [11].
Then, the ligature was removed after 45 min of ischaemia, and the
myocardium was reperfused for 4 h. Ischaemia and reperfusion were
confirmed and monitored by electrocardiogram (ECG) observation. The
suture was then tightened again, and rats were intravenously injected
with 2% Evans Blue (Sigma–Aldrich). After explantation of the hearts,
the left ventricles were isolated, divided into 1 mm slices, and
subsequently incubated in 2% 2,3,5-triphenyltetrazolium chloride (TTC;
Sigma–Aldrich) in 0.9% saline at 37°C for 25 min, to distinguish
infarcted tissue from viable myocardium. These slices were flushed with
saline and then fixed in 10% paraformaldehyde in PBS (pH 7.4) for 2 h.
Next, the slices were placed on a glass slice and photographed by
digital camera, the ImageJ software (NIH) was used in a blind fashion
for analysis. Infarct size was expressed as a ratio of the infarct area
and the area at risk [12].
Pulsed electromagnetic field treatment
PEMF were generated by a
commercially available healing device (length × width × height: 7 cm ×
5cm × 3cm) purchased from Biomoble Regenerative Medicine Technology. The
adapter input voltage parameter is approximately 100–240 V and output
parameter is 5 V. Fields were asymmetric and consisted of 4.5 ms pulses
at 30±3 Hz, with an adjustable magnetic field strength range (X-axis 0.22±0.05 mT, Y-axis 0.20±0.05 mT, Z-axis
0.06±0.02 mT). The I/R rats were housed in custom designed cages and
exposed to active PEMF for 2 cycles per time (8 min for 1 cycle),
whereas the I/R rats were housed in identical cages with inactive PEMF
generator. For in vitro study, culture dishes were directly exposed to PEMF for 1–2 cycles as indicated (8 min for 1 cycle, 30 Hz, X-axis 0.22 mT, Y-axis 0.20 mT, Z-axis 0.06 mT) [1]. The background magnetic field in the room area of exposure animals/samples and controls is 0 mT.
Detection of myocardium apoptosis
Terminal deoxynucleotidyl
transferase-mediated dUTP nick-end labelling (TUNEL) assay was applied
to analyse cardiomyocyte apoptosis. Heart samples were first fixed in
10% formalin and then paraffin embedded at day 14. Then, the hearts were
cut into 5 ?m sections. TUNEL staining was carried out as described
previously [12]. When apoptosis occurred, cells would look green.
Determination of myocardial enzymes in plasma
Blood samples were collected after haemodynamic measurement and centrifuged at 3000 g for
15 min to get the plasma. Creatine kinase (CK), lactate dehydrogenase
(LDH), creatine kinase isoenzyme-MB (CKMB) and ?-hydroxybutyrate
dehydrogenase (HBDH) were quantified by automatic biochemical analyzer
(Cobas 6000, Roche). All procedures were performed according to the
manufacturer’s protocols.
Myocardium cells morphology via TEM
At the end of the experiment,
sections from myocardial samples of left ventricular were immediately
fixed overnight in glutaraldehyde solution at 4°C and then incubated
while protected from light in 1% osmium tetroxide for 2 h. After washing
with distilled water for three times (5 min each), specimens were
incubated in 2% uranyl acetate for 2 h at room temperature and then
dehydrated in graded ethanol concentrations. Finally, sections were
embedded in molds with fresh resin. The changes in morphology and
ultrastructure of the myocardial tissues were observed and photographed
under a TEM [13].
Scal-1+/flk-1+ cells counting of endothelial progenitor cells
We applied antibodies to the stem
cell antigen-1 (Sca-1) and fetal liver kinase-1 (flk-1) to sign
endothelial progenitor cells (EPCs) as described before, and used the
isotype specific conjugated anti-IgG as a negative control. The amount
of Scal-1+/flk-1+ cells would be counted by flow cytometry technique [14].
Measurement of nitric oxide concentration and Western blotting
Plasma concentrations of NO were
measured with Griess assay kit (Beyotime Institute of Biotechnology)
according to the manufacturer’s protocol. The expressions of Bax, Bcl-2,
p-Akt, Akt, p-endothelial nitric oxide synthase (eNOS), eNOS and
glyceraldehyde-3-phosphate dehydrogenase (GAPDH) were assessed using
Western blot as described recently [15].
Proteins were measured with Pierce BCA Protein Assay Kit (Thermo).
Hippocampal protein lysates (50 mg/well) were separated using (SDS/PAGE)
under reducing conditions. Following electrophoresis, the separated
proteins were transferred to a PVDF membrane (Millipore). Subsequently,
non-specific proteins were blocked using blocking buffer (5% skim milk
or 5% BSA in T-TBS containing 0.05% Tween 20), followed by overnight
incubation with primary rabbit anti-rat antibodies specific for target
proteins as mentioned before (Cell Signaling Technology) at 4°C. Blots
were rinsed three times (5 min each) with T-TBS and incubated with
horseradish peroxidase (HRP)-conjugated secondary antibody (1:10000,
Proteintech) for 2 h at room temperature. The blots were visualized by
using enhanced chemiluminescence (ECL) method (Thermo). GAPDH was
applied to be the internal control protein. Intensity of the tested
protein bands was quantified by densitometry.
Cell culture
Primary neonatal rat cardiac ventricular myocytes (NRCMs) were collected as previously described [15].
Briefly, the ventricles of new born SD rats (1–3 days old) were minced
and digested with 0.125% trypsin. Isolated cardiomyocytes were cultured
in Dulbecco’s modified Eagle’s medium/F-12 (DMEM/F12, Life Technologies)
supplemented with 10% (v/v) FBS (Life Technologies), 100 units/ml
penicillin and 100 mg/ml streptomycin. The following experiments used
spontaneously beating cardiomyocytes 48–72 hours after plating. (37°C
with 5% CO2).
Cell treatment (hypoxia/reoxygenation)
NRCMs were prepared according to the methods recently described [15].
To establish the H/R model, the cells were cultured in DMEM/F-12
without glucose and serum. The cells were exposed to hypoxia (99% N2+5% CO2)
for 8 h, followed by reoxygenation for 16 h. The cells were pretreated
with PEMF for 30 min before the H/R procedure. The control group was
cultured in DMEM/F-12 with low glucose (1000 mg/l) and 2% serum under
normoxic air conditions for the corresponding times.
Cell viability assays
The viability of NRCMs cultured in
96-well plates was measured by using the Cell Counting Kit-8 (CCK-8)
(Dojindo Molecular Technologies) according to the manufacturer’s
instructions. The absorbance of CCK-8 was obtained with a microplate
reader at 450 nm.
Measurement of intracellular reactive oxygen species levels
Reactive oxygen species (ROS)
levels in NRVMs were determined by dihydroethidium (DHE, Sigma–Aldrich)
fluorescence using confocal microscopy (Zeiss, LSM 710). After different
treatments, cells were washed with D-PBS and incubated with DHE
(10 ?mol/l) at 37°C for 30 min in the dark. Then, residual DHE was
removed by PBS-washing. Fluorescent signals were observed (excitation,
488 nm; emission, 610 nm) under a laser confocal microscope (Zeiss).
Data analysis
All the data were presented as
means ± S.E.M. Differences were compared by one-way ANOVA analysis by
using SPSS software version 19.0 (SPSS) and P value <0.05 was taken as statistically significant.
RESULTS
PEMF could lower blood pressure under treatment of certain PEMF intensity in SHR rat model (double-blind)
To determine whether PEMF has any
effects on blood pressure of SHR rats, we treated SHR rats with
different PEMF intensity 1–4 cycles per day for 7 days and measured the
blood pressure changes via CCA. We observed that PEMF treatment could
significantly lower the blood pressure in the Bioboosti WIN235 and
WI215-stimulating groups than that in non-treated ones (Figures 1A and ?and1B).1B).
But Bioboosti WIN221 and WC65 treating groups did not have any effects
on the blood pressure in SHR rats, compared with the non-treated ones (Figures 1C and ?and1D).1D). Fields were asymmetric and consisted of 4.5 ms pulses at 30±3 Hz, with an adjustable magnetic field strength range (X-axis 0.22±0.05 mT, Y-axis 0.20±0.05 mT, Z-axis
0.06±0.02 mT). The I/R rats were housed in custom designed cages and
exposed to active PEMF for 2 cycles per time (8 min for 1 cycle),
whereas the I/R rats were housed in identical cages with inactive PEMF
generator.
Figure 1The effect of PEMF on SHR rats in vivo. PEMF could
lower the blood pressure in SHR rats. At day 7 treatment with different
intensity PEMF, blood pressure was recorded via CCA [1(A), 1(B), 1(C)
and 1(D)]. Data were represented as the mean ±…
According to this result, we chose Bioboosti WIN235 as our needed PEMF to carry out the following experiments.
PEMF treatment could observably improve the abundance of EPCs
Amplifying EPCs abundance and function
is an active focus of research on EPCs-mediated neovascularization after
I/R. Thus, the number of circulating EPCs was identified by Sca-1/flk-1
dual positive cells as described. We determined that PEMF treatment
could remarkably increase the number of Scal-1+/flk-1+ cells in peripheral blood at postoperative days 7 and 14 (Figure 2).
Figure 2The effect of PEMF on the number of Scal-1+/flk-1+ cells after treating EPSc for 7 and 14 days. PEMF treatment notably increased the number of Scal-1+/flk-1+ cells after treating EPSc for 7 and 14 days. Data were represented as the mean…
Preliminary assessment of PEMF showed great protective effect against myocardial infarction/reperfusion injury (MI/RI) rat model
To examine the effect of PEMF on
myocardial I/R, male SD rats were divided into three groups: Sham, I/R
and I/R+ PEMF (2 cycles per day, 8 min per cycle) per day until 28 days.
We observed that PEMF stimulation could significantly decrease four
plasma myocardial enzymes (LDH, CK, CKMB and HBDH) in I/R rats (Figure 3A).
Additionally, we found that pre-stimulating PEMF could improve the
cardiac morphology via TEM, compared with I/R+ PEMF group. TEM revealed
the rupture of muscular fibres, together with mitochondrial swelling,
and intracellular oedema in Group I/R. The shape of nucleus was
irregular, with evidence of mitochondrial overflow after cell death.
Compared with Group I/R+ PEMF, less muscular fibres were ruptured, with
mild swelling of mitochondria, mild intercellular oedema and less cell
death. In Group Sham, the ruptured muscular fibres, mitochondrial or
intracellular oedema and dead cells were not observed (Figure 3B).
To further confirm protective effect of PEMF, we measured the MI size
by applying TTC and Evans Blue staining in all three groups. The MI area
in I/R+ PEMF group could be reduced, compared with the model rats in
I/R group (Figure 3C).
Figure 3Protective effect of PEMF on I/R rats in vivo. Plasma myocardial enzymes (LDH, CK, HBDH and CKMB) content was quantified by automatic biochemical analyzer (A) (n=18 in each group). Changes on cardiac cell morphology via TEM (B) (n=6 in…
In vivo, PEMF dramatically reduced cell apoptosis induced by I/R injury
As H/R of cardiomyocytes contributed to
cell death, we also detected the effect on myocardial apoptosis by using
TUNEL kit, as shown in Figure 4(A).
We uncovered that PEMF pretreating could dramatically decrease
apoptosis of myocardial cells in I/R + PEMF group, compared with I/R
group. In addition, we also found that PEMF treatment could
significantly increase the expression of anti-apoptosis protein Bcl-2,
p-eNOS and p-Akt and down-regulated the expression of pro-apoptosis
protein Bax in the heart tissue, as shown in Figure 4(B).
Figure 4Apoptotic cardiomyocyte was identified by TUNEL analysis,
apoptotic cardiomyocyte appears green whereas TUNEL-negative appears
blue (A), photomicrographs were taken at ×200 magnification.
Apoptosis-related protein Bcl-2, Bax, p-Akt level of different…
The effect of PEMF on cell viability in neonatal rat cardiac ventricular myocytes
To further investigate whether PEMF has the same effect in vitro, we simulated the I/R injury model in vitro.
We applied NRCMs and hypoxia incubator to mimic myocardial I/R injury
via H/R as described in the section ‘Materials and Methods’. We found
that PEMF treatment (2 cycles) could remarkably improve cell viability,
compared with the H/R group (Figure 5). For in vitro study, culture dishes were directly exposed to PEMF for 1–2 cycles as indicated (8 min for 1 cycle, 30±3 Hz, X-axis 0.22±0.05 mT, Y-axis 0.20±0.05 mT, Z-axis 0.06±0.02 mT).
Figure 5NRCMs viability measured by CCK-8 assay at the end of the
treatment for 72 h. PEMF treatment enhanced the cell viability of
hypoxia NRCMs. Data were represented as the mean ± S.E.M.
Specific-density PEMF could decrease intracellular ROS levels of primary cardiomyocytes subjected to hypoxia/reperfusion
As shown in Figure 6(A),
NRCMs that were subjected to H/R increased significantly the ROS level,
whereas the ROS level had been decreased in PEMF group (2 cycles), in
contrast with the H/R group. Representative images of the ROS level were
displayed in Figure 6(B). At the same time, we identified the effect on NRCMs apoptosis after suffering H/R by using TUNEL kit. As shown in Figure 6(C),
cell apoptosis in the H/R group was aggravated, whereas PEMF treatment
could reduce the cell death. Representative images of TUNEL staining
were shown in Figure 6(D).
Figure 6PEMF protected Neonatal rat cardiac ventricular myocytes
(NRCMs) from hypoxia/reoxygenation (H/R)-induced apoptosis via
decreasing ROS levelat the end of the treatment for 72 h in vitro.
Effect of PEMF on NO releasing via Akt/eNOS pathway
Cultured NRCMs were treated with PEMF
stimulation for 1 to 2 cycles and the supernatant and cell lysate were
collected. When cells suffered H/R, intracellular levels of p-Akt,
p-eNOS and Bcl-2 were decreased, whereas PEMF treatment could increase
the phosphorylation of Akt, p-eNOS and Bcl-2 (Figures 7A–7C). The expression of Bax was increased when cells subjected to H/R whereas PEMF treatment reversed such increase (Figure 7C). Western blot analysis was shown in Figure 7(D) for p-Akt/Akt, Figure 7(E) for p-eNOS/eNOS, Figure 7(F) for Bcl-2 and Figure 7(G) for Bax.
Figure 7The related protein expression about the effect of PEMF on
apoptosis induced by hypoxia/reoxygenationat the end of the treatment
for 72 h in vitro. PEMF increased the phosphorylation of Akt, endothelial nitric oxide synthase (eNOS), and the expression…Go to:
DISCUSSION
Our present study provides the first
evidence that PEMF has novel functions as follows: (1) We treated SHR
rats with different PEMF intensity (8 min for 1 cycle, 30±3 Hz, X-axis 0.22±0.05 mT, Y-axis 0.20±0.05 mT, Z-axis
0.06±0.02 mT) 1–4 cycles per day for 7 days. PEMF can lower blood
pressure under treatment of certain PEMF intensity in SHR rat model
(double-blind). (2) PEMF has a profound effect on improving cardiac
function in I/R rat model. (3) PEMF plays a vital role in inhibiting
cardiac apoptosis via Bcl-2 up-regulation and Bax down-regulation. (4) In vitro,
PEMF treatment also has a good effect on reducing ROS levels by
Akt/eNOS pathway to release NO and improving cell apoptosis in NRCMs
subjected to hypoxia.
Many previous studies showed that extracorporeal
PEMF-treated(5 mT, 25 Hz, 1 h daily) could enhance osteanagenesis, skin
rapture healing and neuronal regeneration, suggesting its regenerative
potency [8,16,17].
And some researchers had found that PEMF therapy (8 min/cycle, 30±3 Hz,
6 mT) could improve the myocardial infarct by activating VEGF–Enos [18] system and promoting EPCs mobilized to the ischaemic myocardium [1,19].
Consistent with the previous work, our present study demonstrated that
PEMF therapy could significantly alleviate cardiac dysfunction in I/R
rat model.
Recent evidence suggest that circulating EPCs can be
mobilized endogenously in response to tissue ischaemia or exogenously by
cytokine stimulation and the recruitment of EPCs contributes to the
adult blood vessels formation [19,20,21].
We hypothesized that PEMF could recruit more EPCs to the vessels. To
confirm our hypothesis, we applied antibodies to the Sca-1 and flk-1 to
sign EPC. The results indicated that PEMF could remarkably increase the
number of EPCs in the PEMF group, compared with the I/R group.
Previous evidence indicated that when heart suffered I/R, cardiac apoptosis would be dramatically aggravated [22–24].
Myocardial apoptosis plays a significant role in the pathogenesis of
myocardial I/R injury. We assumed that PEMF might play its role in
improving cardiac function through inhibiting cell apoptosis. The Bcl-2
family is a group of important apoptosis-regulating proteins that is
expressed on the mitochondrial outer membrane, endoplasmic reticulum
membrane and nuclear membrane. Overexpression of Bcl-2 proteins blocks
the pro-apoptosis signal transduction pathway, thereby preventing
apoptosis caused by the caspase cascade [25].
The role Bax plays in autophagy is a debatable. Recently, new genetic
and biochemical evidence suggest that Bcl-2/Bcl-xL may affect apoptosis
through its inhibition of Bax [26].
Overexpression of Bax protein promotes the apoptosis signal pathway. In
the present study, we applied TUNEL staining to find that PEMF has a
perfect effect on cardiac cell apoptosis by regulating apoptosis-related
proteins Bcl-2 and Bax [25,26,27,28].
To verify our findings in the rat model, we mimicked I/R condition in vitro by hypoxia exposure in NRCMs. Results showed that not only in vivo, hypoxia could induce cell apoptosis in vitro.
And we also found that PEMF treatment could significantly alleviate
cell apoptosis induced by hypoxia. At the basal level, ROS play an
important role in mediating multiple cellular signalling cascades
including cell growth and stress adaptation. Conversely, excess ROS can
damage tissues by oxidizing important cellular components such as
proteins, lipids and DNA, as well as activating proteolytic enzymes such
as matrix metalloproteinases [29].
Previous studies showed that when cells were subjected to hypoxia, the
intracellular ROS level would be sharply increased, and the
overproduction of ROS would result in cell damage [19,30,31].
In the present study, PEMF treatment could prominently down-regulate
ROS levels. We also investigated how PEMF reduced the intracellular ROS
level.
NO appears to mediate distinct pathways in response to oxidative stress via AKt–eNOS pathway [32,33].
NO is identified as gaseous transmitters. In vascular tissue, NO is
synthesized from L-arginine by nitric oxide synthase (NOS) and it is
considered to be the endothelium-derived relaxing factor. Evidence show
that the NO generation in endothelium cells was damaged in hypertensive
patients [34]. NO could also prevent platelet activation and promote vascular smooth muscle cells proliferation [35]. NO generation from eNOS is considered to be endothelium-derived relaxing and ROS-related factor [36,37]. Some researchers found that bradykinin limited MI induced by I/R injury via Akt/eNOS signalling pathway in mouse heart [38].
And bradykinin inhibited oxidative stress-induced cardiomyocytes
senescence by acting through BK B2 receptor induced NO release [39].
Such evidence indicated that Akt phosphorylation could activate eNOS,
which lead to NO releasing, and resulted in ROS reducing. In the present
study, we found that PEMF decreased ROS via Akt/eNOS pathway.
In conclusion, this is the first study
suggesting that PEMF treatment could improve cardiac dysfunction through
inhibiting cell apoptosis. Furthermore, in vitro, we first
clarified PEMF still plays a profound effect on improving cell death and
removing excess ROS via regulating apoptosis-related proteins and
Akt/eNOS pathway. All these findings highlight that PEMF would be
applied as a potentially powerful therapy for I/R injury cure.
Acknowledgments
We thank all of the members of the Laboratory of Pharmacology of Chen Y., Ding Y.J. for their technical assistance.
Fenfen Ma designed and performed experiments on MI/RI rat model, histological stain and Western blot. Wenwen Li assisted the in vivo experiments, validated the effect in vitro experiments,
analysed data and wrote the manuscript. Xinghui Li interpreted data and
formatted manuscript. Rinkiko Suguro, Ruijuan Guan, Cuilan Hou, Huijuan
Wang and Aijie Zhang interpreted data and edited manuscript. Yichun Zhu
and YiZhun Zhu proposed the idea and supervised the project.
FUNDING
This work was supported by the key
laboratory program of the Education Commission of Shanghai Municipality
[grant number ZDSYS14005]
.
References
1. Hao
C.N., Huang J.J., Shi Y.Q., Cheng X.W., Li H.Y., Zhou L., Guo X.G., Li
R.L., Lu W., Zhu Y.Z., Duan J.L. Pulsed electromagnetic field improves
cardiac function in response to myocardial infarction. Am. J. Transl. Res. 2014;6:281–290. [PMC free article] [PubMed]
2. Eltzschig H.K., Eckle T. Ischemia and reperfusion–from mechanism to translation. Nat. Med. 2011;17:1391–1401. doi: 10.1038/nm.2507. [PMC free article] [PubMed] [Cross Ref]
3. Thygesen
K., Alpert J.S., Jaffe A.S., Simoons M.L., Chaitman B.R., White H.D.
Third universal definition of myocardial infarction. Nat. Rev. Cardiol. 2012;9:620–633. doi: 10.1038/nrcardio.2012.122.[PubMed] [Cross Ref]
4. Nah D.Y., Rhee M.Y. The inflammatory response and cardiac repair after myocardial infarction. Korean Circ. J. 2009;39:393–398. doi: 10.4070/kcj.2009.39.10.393. [PMC free article] [PubMed] [Cross Ref]
5. Yellon D.M., Hausenloy D.J. Myocardial reperfusion injury. N. Engl. J. Med. 2007;357:1121–1135. doi: 10.1056/NEJMra071667. [PubMed] [Cross Ref]
6. Herron
T.J., Milstein M.L., Anumonwo J., Priori S.G., Jalife J. Purkinje cell
calcium dysregulation is the cellular mechanism that underlies
catecholaminergic polymorphic ventricular tachycardia. Heart Rhythm. 2010;7:1122–1128. doi: 10.1016/j.hrthm.2010.06.010. [PMC free article] [PubMed] [Cross Ref]
7. Kim
S.S., Shin H.J., Eom D.W., Huh J.R., Woo Y., Kim H., Ryu S.H., Suh
P.G., Kim M.J., Kim J.Y., et al. Enhanced expression of neuronal nitric
oxide synthase and phospholipase C-gamma1 in regenerating murine
neuronal cells by pulsed electromagnetic field. Exp. Mol. Med. 2002;34:53–59. doi: 10.1038/emm.2002.8. [PubMed] [Cross Ref]
8. Tepper
O.M., Callaghan M.J., Chang E.I., Galiano R.D., Bhatt K.A., Baharestani
S., Gan J., Simon B., Hopper R.A., Levine J.P., Gurtner G.C.
Electromagnetic fields increase in vitro and in vivo angiogenesis through endothelial release of FGF-2. FASEB J. 2004;18:1231–1233. [PubMed]
9. Weintraub
M.I., Herrmann D.N., Smith A.G., Backonja M.M., Cole S.P. Pulsed
electromagnetic fields to reduce diabetic neuropathic pain and stimulate
neuronal repair: a randomized controlled trial. Arch. Phys. Med. Rehabil. 2009;90:1102–1109. doi: 10.1016/j.apmr.2009.01.019. [PubMed] [Cross Ref]
10. Graak
V., Chaudhary S., Bal B.S., Sandhu J.S. Evaluation of the efficacy of
pulsed electromagnetic field in the management of patients with diabetic
polyneuropathy. Int. J. Diab. Dev. Ctries. 2009;29:56–61. doi: 10.4103/0973-3930.53121. [PMC free article] [PubMed] [Cross Ref]
11. Kin
H., Zhao Z.Q., Sun H.Y., Wang N.P., Corvera J.S., Halkos M.E., Kerendi
F., Guyton R.A., Vinten-Johansen J. Postconditioning attenuates
myocardial ischemia-reperfusion injury by inhibiting events in the early
minutes of reperfusion. Cardiovasc. Res. 2004;62:74–85. doi: 10.1016/j.cardiores.2004.01.006.[PubMed] [Cross Ref]
12. Yao
L.L., Huang X.W., Wang Y.G., Cao Y.X., Zhang C.C., Zhu Y.C. Hydrogen
sulfide protects cardiomyocytes from hypoxia/reoxygenation-induced
apoptosis by preventing GSK-3beta-dependent opening of mPTP. Am. J. Physiol. Heart. Circ. Physiol. 2010;298:H1310–H1319. doi: 10.1152/ajpheart.00339.2009. [PubMed] [Cross Ref]
13. Zhikun
G., Liping M., Kang G., Yaofeng W. Structural relationship between
microlymphatic and microvascullar blood vessels in the rabbit
ventricular myocardium. Lymphology. 2013;46:193–201.[PubMed]
14. Tsai
S.H., Huang P.H., Chang W.C., Tsai H.Y., Lin C.P., Leu H.B., Wu T.C.,
Chen J.W., Lin S.J. Zoledronate inhibits ischemia-induced
neovascularization by impairing the mobilization and function of
endothelial progenitor cells. PLoS ONE. 2012;7:e41065. doi: 10.1371/journal.pone.0041065.[PMC free article] [PubMed] [Cross Ref]
15. Jin S., Pu S.X., Hou C.L., Ma F.F., Li N., Li X.H., Tan B., Tao B.B., Wang M.J., Zhu Y.C. Cardiac H2S generation is reduced in ageing diabetic mice. Oxid. Med. Cell. Longev. 2015;2015:758358.[PMC free article] [PubMed]
16. Cheing
G.L., Li X., Huang L., Kwan R.L., Cheung K.K. Pulsed electromagnetic
fields (PEMF) promote early wound healing and myofibroblast
proliferation in diabetic rats. Bioelectromagnetics. 2014;35:161–169. doi: 10.1002/bem.21832. [PubMed] [Cross Ref]
17. Weintraub
M.I., Herrmann D.N., Smith A.G., Backonja M.M., Cole S.P. Pulsed
electromagnetic fields to reduce diabetic neuropathic pain and stimulate
neuronal repair: a randomized controlled trial. Arch. Phys. Med. Rehabil. 2009;90:1102–1109. doi: 10.1016/j.apmr.2009.01.019. [PubMed] [Cross Ref]
18. Li
J., Zhang Y., Li C., Xie J., Liu Y., Zhu W., Zhang X., Jiang S., Liu
L., Ding Z. HSPA12B attenuates cardiac dysfunction and remodelling after
myocardial infarction through an eNOS-dependent mechanism. Cardiovasc. Res. 2013;99:674–684. doi: 10.1093/cvr/cvt139. [PubMed] [Cross Ref]
19. Goto
T., Fujioka M., Ishida M., Kuribayashi M., Ueshima K., Kubo T.
Noninvasive up-regulation of angiopoietin-2 and fibroblast growth
factor-2 in bone marrow by pulsed electromagnetic field therapy. J. Orthop. Sci. 2010;15:661–665. doi: 10.1007/s00776-010-1510-0. [PubMed] [Cross Ref]
20. Asahara
T., Masuda H., Takahashi T., Kalka C., Pastore C., Silver M., Kearne
M., Magner M., Isner J.M. Bone marrow origin of endothelial progenitor
cells responsible for postnatal vasculogenesis in physiological and
pathological neovascularization. Circ. Res. 1999;85:221–228. doi: 10.1161/01.RES.85.3.221. [PubMed] [Cross Ref]
21. Takahashi
T., Kalka C., Masuda H., Chen D., Silver M., Kearney M., Magner M.,
Isner J.M., Asahara T. Ischemia- and cytokine-induced mobilization of
bone marrow-derived endothelial progenitor cells for
neovascularization. Nat. Med. 1999;5:434–438. doi: 10.1038/8462. [PubMed] [Cross Ref]
22. Freude
B., Masters T.N., Robicsek F., Fokin A., Kostin S., Zimmermann R.,
Ullmann C., Lorenz-Meyer S., Schaper J. Apoptosis is initiated by
myocardial ischemia and executed during reperfusion. J. Mol. Cell Cardiol. 2000;32:197–208. doi: 10.1006/jmcc.1999.1066. [PubMed] [Cross Ref]
23. Martindale
J.J., Fernandez R., Thuerauf D., Whittaker R., Gude N., Sussman M.A.,
Glembotski C.C. Endoplasmic reticulum stress gene induction and
protection from ischemia/reperfusion injury in the hearts of transgenic
mice with a tamoxifen-regulated form of ATF6. Circ. Res. 2006;98:1186–1193. doi: 10.1161/01.RES.0000220643.65941.8d. [PubMed] [Cross Ref]
24. Yu
L., Lu M., Wang P., Chen X. Trichostatin A ameliorates myocardial
ischemia/reperfusion injury through inhibition of endoplasmic reticulum
stress-induced apoptosis. Arch. Med. Res. 2012;43:190–196. doi: 10.1016/j.arcmed.2012.04.007. [PubMed] [Cross Ref]
25. Maiuri
M.C., Criollo A., Tasdemir E., Vicencio J.M., Tajeddine N., Hickman
J.A., Geneste O., Kroemer G. BH3-only proteins and BH3 mimetics induce
autophagy by competitively disrupting the interaction between Beclin 1
and Bcl-2/Bcl-X(L) Autophagy. 2007;3:374–376. doi: 10.4161/auto.4237.[PubMed] [Cross Ref]
26. Lindqvist
L.M., Heinlein M., Huang D.C., Vaux D.L. Prosurvival Bcl-2 family
members affect autophagy only indirectly, by inhibiting Bax and Bak. Proc. Natl. Acad. Sci. U.S.A. 2014;111:8512–8517. doi: 10.1073/pnas.1406425111. [PMC free article] [PubMed] [Cross Ref]
27. Chandna
S., Suman S., Chandna M., Pandey A., Singh V., Kumar A., Dwarakanath
B.S., Seth R.K. Radioresistant Sf9 insect cells undergo an atypical form
of Bax-dependent apoptosis at very high doses of gamma-radiation. Int. J. Rad. Biol. 2013;89:1017–1027. doi: 10.3109/09553002.2013.825059. [PubMed][Cross Ref]
28. Xu
M., Zhou B., Wang G., Wang G., Weng X., Cai J., Li P., Chen H., Jiang
X., Zhang Y. miR-15a and miR-16 modulate drug resistance by targeting
bcl-2 in human colon cancer cells. Zhonghua Zhong Liu Za Zhi. 2014;36:897–902. [PubMed]
29. Zuo L., Best T.M., Roberts W.J., Diaz P.T., Wagner P.D. Characterization of reactive oxygen species in diaphragm. Acta Physiol. (Oxf.) 2015;213:700–710. doi: 10.1111/apha.12410. [PubMed] [Cross Ref]
30. Kalogeris
T., Bao Y., Korthuis R.J. Mitochondrial reactive oxygen species: a
double edged sword in ischemia/reperfusion vs preconditioning. Redox Biol. 2014;2:702–714. doi: 10.1016/j.redox.2014.05.006.[PMC free article] [PubMed] [Cross Ref]
31. Levraut
J., Iwase H., Shao Z.H., Vanden H.T., Schumacker P.T. Cell death during
ischemia: relationship to mitochondrial depolarization and ROS
generation. Am. J. Physiol. Heart Circ. Physiol. 2003;284:H549–H558. doi: 10.1152/ajpheart.00708.2002. [PubMed] [Cross Ref]
32. Dong
R., Chen W., Feng W., Xia C., Hu D., Zhang Y., Yang Y., Wang D.W., Xu
X., Tu L. Exogenous bradykinin inhibits tissue factor induction and deep
vein thrombosis via activating the eNOS/phosphoinositide 3-kinase/Akt
signaling pathway. Cell. Physiol. Biochem. 2015;37:1592–1606. doi: 10.1159/000438526. [PubMed] [Cross Ref]
33. Jin R.C., Loscalzo J. Vascular nitric oxide: formation and function. J. Blood Med. 2010;2010:147–162.[PMC free article] [PubMed]
34. Taddei
S., Virdis A., Mattei P., Ghiadoni L., Sudano I., Salvetti A. Defective
L-arginine-nitric oxide pathway in offspring of essential hypertensive
patients. Circulation. 1996;94:1298–1303. doi: 10.1161/01.CIR.94.6.1298. [PubMed] [Cross Ref]
35. Tang E.H., Vanhoutte P.M. Endothelial dysfunction: a strategic target in the treatment of hypertension? Pflugers Arch. 2010;459:995–1004. doi: 10.1007/s00424-010-0786-4. [PubMed] [Cross Ref]
36. Beltowski J., Jamroz-Wisniewska A. Hydrogen sulfide and endothelium-dependent vasorelaxation. Molecules. 2014;19:21183–21199. doi: 10.3390/molecules191221183. [PubMed] [Cross Ref]
37. Wu
D., Hu Q., Liu X., Pan L., Xiong Q., Zhu Y.Z. Hydrogen sulfide protects
against apoptosis under oxidative stress through SIRT1 pathway in H9c2
cardiomyocytes. Nitric Oxide. 2015;46:204–212. doi: 10.1016/j.niox.2014.11.006. [PubMed] [Cross Ref]
38. Li
Y.D., Ye B.Q., Zheng S.X., Wang J.T., Wang J.G., Chen M., Liu J.G., Pei
X.H., Wang L.J., Lin Z.X., et al. NF-kappaB transcription factor p50
critically regulates tissue factor in deep vein thrombosis. J. Biol. Chem. 2009;284:4473–4483. doi: 10.1074/jbc.M806010200. [PMC free article] [PubMed] [Cross Ref]
39. Dong
R., Xu X., Li G., Feng W., Zhao G., Zhao J., Wang D.W., Tu L.
Bradykinin inhibits oxidative stress-induced cardiomyocytes senescence
via regulating redox state. PLoS ONE. 2013;8:e77034. doi: 10.1371/journal.pone.0077034. [PMC free article] [PubMed] [Cross Ref]
Klin Med (Mosk). 2003;81(1):24-7.
Am J Transl Res. 2014; 6(3): 281–290.
Published online May 15, 2014
Clinico-functional efficacy of medicinal and photon stabilization in patients with angina pectoris.
Modification of erythrocytic membrane and the trend in
clinicofunctional indices were studied in 90 patients with angina of
effort (FC I-IV) in the course of treatment with a combination of
membranoprotective drugs (group 1), magneto-laser radiation (group 2)
and imitation of laser radiation (group 3). In patients of groups 1 and 2
the treatment resulted in stabilization of cell membrane accompanied
with a hypotensive effect and increased exercise tolerance due to more
effective cardiac performance.um chloride baths. Its efficacy was 63%.
The effect was due to inhibition of high sympathico-adrenal system.
Klin Med (Mosk). 2000;78(3):23-5.
Characteristics of microcirculation and vascular responsiveness in
elderly patients with hypertension and ischemic heart disease.
[Article in Russian]
Abramovich SG.
Microcirculation and vascular responsiveness were studied in 52
patients with arterial hypertension and ischemic heart disease versus 48
healthy elderly persons. The patients were found to have defects of the
end blood flow in all links of microcirculation, longer and more severe
vasoconstriction of conjunctival and skin vessels in response to
norepinephrine and cold stimulation tests.
Crit Rev Biomed Eng. 2000;28(1-2):339-47.
The use of millimeter wavelength electromagnetic waves in cardiology.
Lebedeva AYu.
2nd Department of urgent cardiology at State Clinical Hospital, Russian State Medical University, Moscow.
Abstract
This paper concerns the problems of the use of millimeter wavelength
electromagnetic waves for the treatment of cardiovascular disease. The
prospects for this use are considered.
Central hemodynamics, diastolic and pumping functions of the heart,
myocardial reactivity, microcirculation and biological age of
cardiovascular system were studied in 66 elderly patients suffering from
hypertension and ischemic heart disease. The patients received systemic
magnetotherapy which produced a geroprotective effect as shown by
improved microcirculation, myocardial reactivity, central hemodynamics
reducing biological age of cardiovascular system and inhibiting its
ageing.
The effect of exposure to magnetics and lasers on the clinical
status and the electrophysiological indices of the heart in patients
with cardiac arrhythmias.
Magnetolaser radiation has a considerable influence on
electrophysiological condition of the sinus node and sinoatrial zone.
There are cases when patients with sick sinus syndrome get rid of
arrhythmia. The treatment is safe and promising for further studies.
The effect of a low-frequency magnetic field on erythrocyte membrane
function and on the prostanoid level in the blood plasma of children
with parasystolic arrhythmia.
As shown by clinical and biochemical evidence on 23 parasystolic
children, the treatment with low-frequency magnetic field improves
humoral and cellular processes participating in cardiac rhythm
regulation. There is activation of Ca, Mg-ATPase in the red cells, a
reduction of plasma thromboxane levels. Red cell phospholipid
composition insignificantly change. Further courses of magnetotherapy
may lower the risk of recurrent arrhythmia.
Changes in intracellular regeneration and the indices of endocrine
function and cardiac microcirculation in exposure to decimeter waves.
[Article in Russian]
Korolev IuN, Geniatulina MS, Popov VI.
Abstract
An electron-microscopic study of rabbit heart with experimental
myocardial infarction revealed that extracardiac exposure to decimetric
waves (DW) activated intracellular regeneration in the myocardium. This
was associated with enhanced circulation and endocrine activity in the
heart. Most pronounced regeneration was registered in adrenal exposure,
the effect of the parietal exposure being somewhat less.
Lik Sprava. 1992 May;(5):40-3.
The effect of combined treatment with the use of magnetotherapy on
the systemic hemodynamics of patients with ischemic heart disease and
spinal osteochondrosis.
[Article in Russian]
Dudchenko MA, Vesel’skii ISh, Shtompel’ VIu.
The authors examined 66 patients with ischemic heart disease and
concomitant cervico-thoracic osteochondrosis and 22 patients without
osteochondrosis. Differences were revealed in values of the systemic
hemodynamics with prevalence of the hypokinetic type in patients with
combined pathology. Inclusion of magnetotherapy in the treatment complex
of patients with ischemic heart disease and osteochondrosis favours
clinical improvement, normalization of indices of central and regional
blood circulation.
The effect of decimeter waves on the metabolism of the myocardium
and its hormonal regulation in rabbits with experimental ischemia.
[Article in Russian]
Frenkel’ ID, Zubkova SM, Liubimova NN, Popov VI.
Abstract
Biochemical and morphometric methods were employed to study the
effect of decimetric waves (460 MHz, 10 and 120 mW/cm2) in cardiac and
thyroid exposure on oxygen metabolism, myocardial microcirculation and
contractility, thyroid and adrenal hormonal activity, kallikrein-kinin
system activity in rabbits with experimental myocardial ischemia.
Hypoxia discontinued in all the treatment regimens, but the exposure of
the heart (field density 10 mW/sm2) had the additional effect on lipid
peroxidation which reduced in the serum and normalized in the
myocardium, on myocardial contractility, kallikrein-kinin system and on
the adrenal and thyroid hormones.
OBJECTIVE: The aim of this study was to investigate if laser therapy
in combination with pulsed electromagnetic field therapy/repetitive
transcranial magnetic stimulation (rTMS) and the control of reactive
oxygen species (ROS) would lead to positive treatment results for
hyperacusis patients.
BACKGROUND DATA: Eight of the first ten patients treated for
tinnitus, who were also suffering from chronic hyperacusis, claimed
their hyperacusis improved. Based upon that, a prospective, unblinded,
uncontrolled clinical trial was planned and conducted. ROS and
hyperacusis pain thresholds were measured.
MATERIALS AND METHODS: Forty-eight patients were treated twice a week
with a combination of therapeutic laser, rTMS, and the control and
adjustment of ROS. A magnetic field of no more than 100 microT was
oriented behind the outer ear, in the area of the mastoid bone. ROS were
measured and controlled by administering different antioxidants. At
every treatment session, 177-504 J of laser light of two different
wavelengths was administered toward the inner ear via meatus acusticus.
RESULTS: The improvements were significantly better in the verum
group than in a placebo group, where 40% of the patients were expected
to have a positive treatment effect. The patients in the long-term
follow-up group received significantly greater improvements than the
patients in the short-term follow-up group.
CONCLUSION: The treatment is effective in treating chronic hyperacusis.
Vestn Otorinolaringol. 2002;(1):11-4.
Electrophysical effects in combined treatment of neurosensory hypoacusis.
Article in Russian]
Morenko VM, Enin IP.
The authors consider different methods of electrobiophysical impacts
on the body in the treatment of neurosensory hypoacusis: laser beam,
laser puncture, electrostimulation, magnetotherapy, magnetolasertherapy,
electrophoresis, etc. These methods find more and more intensive
application in modern medicine. Further success of physiotherapy for
neurosensory hypoacusis depends on adequate knowledge about mechanisms
of action of each physical method used and introduction of novel
techniques.
Vestn Otorinolaringol. 2001;(4):10-2.
Cerebral hemodynamics in patients with neurosensory hearing loss before and after magnetotherapy. a prospective clinical study.
Article in Russian]
Morenko VM, Enin IP.
Magnetotherapy effects on cerebral hemodynamics were studied using
rheoencephalography (REG). When the treatment results and changes in
cerebral hemodynamics were compared it was evident that normalization or
improvement of vascular status in vertebrobasilar and carotid
territories registered at REG results in better hearing. This confirms
the role of vascular factor in pathogenesis of neurosensory hypoacusis
of different etiology and effectiveness of magnetotherapy in such
patients.
Vestn Otorinolaringol. 1996 Nov-Dec;(6):23-6.
The treatment of hypoacusis in children by using a pulsed low-frequency electromagnetic field.
[Article in Russian]
Bogomil’skii MR, Sapozhnikov IaM, Zaslavskii AIu, Tarutin NP.
The authors provide specifications of the unit INFITA supplied with
ELEMAGS attachment of their own design; the technique of treating
hypoacusis in children with utilization of impulse low-frequency
electromagnetic field; the results of this treatment in 105 hypoacusis
children. The method was found highly effective and valuable for wide
practice.
Med Tekh. 1995 Mar-Apr;(2):40-1.
ELEMAGS. apparatus and clinical experience in its use in the treatment of children with hypoacusis and otalgia.
To enhance effectiveness of magnetotherapy in the treatment of otic
diseases the authors propose to use impulse low-frequency
electromagnetic field in combination with constant magnetic field.
ELEMAGS equipment based on the above principles is introduced to treat
cochlear neuritis and neurosensory hypoacusis in children.
The magnetic amplipulse therapy of vestibular dysfunctions of
vascular origin by using the Sedaton apparatus (experimental research).
[Article in Russian]
Mal’tsev AE.
The paper describes the results of combined utilization of magnetic
field (MF), sinusoidal modulated current (SMC) and galvanic current (GC)
generated by a specially devised unit “Sedaton”. This multimodality
physiotherapy was tested in chronic experiments on 25 cats with
experimental vascular and vestibular dysfunction. MF in combination with
SMC displayed greater efficacy than in monotherapy. Positive
physiological reactions were more pronounced.
Headache. 2010 Jul;50(7):1153-63. Epub 2010 Jun 10.
Transcranial magnetic stimulation for migraine: a safety review.
Dodick DW, Schembri CT, Helmuth M, Aurora SK.
Source
Mayo Clinic, Phoenix, AZ, USA.
Abstract
OBJECTIVE:
To review potential and theoretical safety concerns of transcranial
magnetic stimulation (TMS), as obtained from studies of single-pulse
(sTMS) and repetitive TMS (rTMS) and to discuss safety concerns
associated with sTMS in the context of its use as a migraine treatment.
METHODS:
The published literature was reviewed to identify adverse events
that have been reported during the use of TMS; to assess its potential
effects on brain tissue, the cardiovascular system, hormone levels,
cognition and psychomotor tests, and hearing; to identify the risk of
seizures associated with TMS; and to identify safety issues associated
with its use in patients with attached or implanted electronic
equipment or during pregnancy.
RESULTS:
Two decades of clinical experience with sTMS have shown it to be a
low risk technique with promise in the diagnosis, monitoring, and
treatment of neurological and psychiatric disease in adults. Tens of
thousands of subjects have undergone TMS for diagnostic, investigative,
and therapeutic intervention trial purposes with minimal adverse
events or side effects. No discernable evidence exists to suggest that
sTMS causes harm to humans. No changes in neurophysiological function
have been reported with sTMS use.
CONCLUSIONS:
The safety of sTMS in clinical practice, including as an acute
migraine headache treatment, is supported by biological, empirical, and
clinical trial evidence. Single-pulse TMS may offer a safe
nonpharmacologic, nonbehavioral therapeutic approach to the currently
prescribed drugs for patients who suffer from migraine.
Appl Psychophysiol Biofeedback. 2007 Dec;32(3-4):191-207. Epub 2007 Nov 2.
Headache treatment with pulsing electromagnetic fields: a literature review.
Vincent W, Andrasik F, Sherman R.
Department of Psychology, Georgia State University, PO Box 5010, Atlanta, GA 30302-5010, USA.
Abstract
Pulsing electromagnetic field (PEMF) therapy may be a viable form of
complementary and alternative medicine. Clinical applications include
the treatment of fractures, wounds, and heart disease. More recent
applications involve treatment of recurrent headache disorders. This
paper reviews available studies investigating PEMF for headache
management. Possible mechanisms for effects (neurochemical,
electrophysical, and cardiovascular) are discussed. The available data
suggest that PEMF treatment for headache merits further study.
Suggestions for future research are provided.
EEG EMG Z Elektroenzephalogr Elektromyogr Verwandte Geb. 1985 Dec;16(4):227-30.
Cerebral use of a pulsating field in neuropsychiatry patients with long-term headache.
[Article in German]
Grunner O.
A pulsed magnetic field (f = 260 Hz; t = 3 ms; induction B = 1.9 mT;
gradient = 0.5 mT/cm) was applied at 40 patients with headaches of
various etiology. The change of cephalea intensity was evaluated
according the patients statements. These statements were further
compared with the changes of the EEG. By means of frequency analysis of
the EEG significant changes in percentages of delta and alpha 1
activities (7.5-9.5/s) were stated after the application of the real
treatment regarding the sham treatment. Any treatment lasted one half
hour. The retreat of subjective difficulties as well as the amelioration
of EEG were stated accordingly at headaches, which were bounded with
cerebral arteriosclerosis, with states after cerebral concussion, with
depressive neurosis, or with tension headache. Pulsed magnetic field
could be applied only there, where the visual evaluation stated EEG as
physiological.
Headache: The Journal of Head and Face Pain
Volume 39 Issue 8 Page 567 – September 1999
doi:10.1046/j.1526-4610.1999.3908567.x
Treatment of Migraine With Pulsing Electromagnetic Fields: A Double-Blind, Placebo-Controlled Study
Richard A. Sherman, PhD; Nancy M. Acosta, BS; Linda Robson, BA
The effect of exposure to pulsing
electromagnetic fields on migraine activity was evaluated by having
42 subjects (34 women and 8 men), who met the International Headache
Society’s criteria for migraine, participate in a double-blind,
placebo-controlled study. Each subject kept a 1-month, pretreatment,
baseline log of headache activity prior to being randomized to having
either actual or placebo pulsing electromagnetic fields applied to
their inner thighs for 1 hour per day, 5 days per week, for 2 weeks.
After exposure, all subjects kept the log for at least 1 follow-up
month. During the first month of follow-up, 73% of those receiving
actual exposure reported decreased headaches (45% good decrease,
14% excellent decrease) compared to half of those receiving the
placebo (15% worse, 20% good, 0% excellent). Ten of the 22 subjects
who had actual exposure received 2 additional weeks of actual
exposure after their initial 1-month follow-up. All showed decreased
headache activity (50% good, 38% excellent). Thirteen subjects from
the actual exposure group elected not to receive additional exposure.
Twelve of them showed decreased headache activity by the second
month (29% good, 43% excellent). Eight of the subjects in the placebo
group elected to receive 2 weeks of actual exposure after the
initial 1-month follow-up with 75% showing decreased headache
activity (38% good, 38% excellent).
In conclusion, exposure of the inner thighs to pulsing
electromagnetic fields for at least 3 weeks is an effective,
short-term intervention for migraine, but not tension headaches.
Headache: The Journal of Head and Face Pain
Volume 38 Issue 3 Page 208 – March 1998
doi:10.1046/j.1526-4610.1998.3803208.x
Initial Exploration of Pulsing Electromagnetic Fields for Treatment of Migraine
Richard A. Sherman, PhD; Linda Robson, BA; Linda A. Marden, MD
Two studies were conducted during which
23 patients with chronic migraine were exposed to pulsing
electromagnetic fields over the inner thigh. In an open study, 11
subjects kept a 2-week headache log before and after 2 to 3 weeks of
exposure to pulsing electromagnetic fields for 1 hour per day, 5 days
per week. The number of headaches per week decreased from 4.03
during the baseline period to 0.43 during the initial 2-week
follow-up period and to 0.14 during the extended follow-up which
averaged 8.1 months. In a double-blind study, 9 subjects kept a
3-week log of headache activity and were randomly assigned to receive
2 weeks of real or placebo pulsing elactromagnetic field exposures
as described above. They were subsequently switched to 2 weeks of the
other mode, after which they kept a final 3-week log. Three
additional subjects in the blind study inadvertently received
half-power pulsing electromagnetic field exposures. The 6 subjects
exposed to the actual device first showed a change in headache
activity from 3.32 per week to 0.58 per week. The 3 subjects exposed
to only half the dose showed no change in headache activity. Large
controlled studies should be performed to determine whether this
intervention is actually effective
Curr Rev Pain. 1999;3(5):342-347.
Sphenopalatine Ganglion Analgesia.
Day M.
Texas Tech University Health Sciences Center, Department of
Anesthesiology, 3601 4th Street, Room 1C282, Lubbock, TX 79430, USA.
The sphenopalatine ganglion and its involvement in the
pathogenesis of pain has been the subject of debate for the last 90
years. The ganglion is a complex neural center composed of sensory,
motor, and autonomic nerves, which makes it difficult to determine
its pathophysiology. Current indications for blockade of the
sphenopalatine ganglion include sphenopalatine and trigeminal
neuralgia, migraine and cluster headaches, and atypical facial pain.
Methods of blockade use local anesthetics, steroids, phenol, and
conventional radiofrequency and electromagnetic field- pulsed
radiofrequency lesioning. The techniques for blockade range from
superficial to highly invasive. Efficacy studies, though few and
small, show promise in patients who have failed pharmacologic or
surgical therapies.
Anesth Pain Control Dent. 1992 Spring;1(2):85-9.
The management of craniofacial pain in a pain relief unit.
Hillman L, Burns MT, Chander A, Tai YM.
Russells Hall Hospital, Dudley, United Kingdom.
This paper reports the results of 34 craniofacial pain sufferers who
were treated at the Dudley Pain Relief Unit over a 1-year period. Most
of the patients were referred by their general medical practitioners.
They were adults representing all age groups, with a female-male ratio
of 4:1. The average history of pain was 5.5 years. Neuralgic pain (as
distinct from temporomandibular joint dysfunction syndrome, migrainous
disorders, and pain of iatrogenic origin) was most frequently seen. Oral
drug therapy, local injection of corticosteroids and analgesics,
peripheral neurolysis, magnetotherapy, hypnotherapy, and acupuncture
were the lines of management available. By the end of this study period,
pain had been relieved or eliminated in 30 of the patients (88%).
J Physiol Pharmacol. 2012 Sep;63(4):397-401.
Pulsating electromagnetic field stimulation of urothelial cells
induces apoptosis and diminishes necrosis: new insight to magnetic
therapy in urology.
Juszczak K, Kaszuba-Zwoinska J, Thor PJ.
Source
Department of Pathophysiology, Jagiellonian University, Medical College, Cracow, Poland. kajus13@poczta.onet.pl.
Abstract
The evidence of electromagnetic therapy (EMT) efficacy in stress
and/or urge urinary incontinence, as well as in detrusor overactivity is
generally lacking in the literature. The potential EMT action of
neuromuscular tissue depolarization has been described. Because there is
no data on the influence of pulsating electromagnetic fields (PEMF) on
the urothelium, we evaluated the effect of PEMF stimulation on rat
urothelial cultured cells (RUCC). In our study 15 Wistar rats were used
for RUCC preparation. RUCC were exposed to PEMF (50 Hz, 45±5 mT) three
times for 4 hours each with 24-hour intervals. The unexposed RUCC was in
the same incubator, but in a distance of 35 cm from the PEMF generator.
Annexin V-APC (AnV+) labelled was used to determine the percentage of
apoptotic cells and propidium iodide (PI+), as standard flow cytometric
viability probe to distinguish necrotic cells from viable ones. The
results are presented in percentage values. The flow cytometric analysis
was carried out on a FACS calibur flow cytometer using Cell-Quest
software. In PEMF-unstimulated RUCC, the percentage of AnV+, PI+, and
AnV+PI+ positive cells were 1.24±0.34%, 11.03±1.55%, and 12.43±1.96%,
respectively. The percentages of AnV+, PI+, and AnV+PI+ positive cells
obtained after PEMF stimulation were 1.45±0.16% (p=0.027), 7.03±1.76%
(p<0.001), and 9.48±3.40% (p=0.003), respectively. The PEMF
stimulation of RUCC induces apoptosis (increase of AnV+ cells) and
inhibits necrosis (decrease of PI+ cells) of urothelial cells. This
leads us to the conclusion that a low-frequency pulsating
electromagnetic field stimulation induces apoptosis and diminishes
necrosis of rat urothelial cells in culture.
Vopr Kurortol Fizioter Lech Fiz Kult. 2005 Jan-Feb;(1):26-8.
Efficacy of general magnetotherapy in conservative therapy of uterine myoma in women of reproductive age.
[Article in Russian]
Kulishova TV, Tabashnikova NA, Akker LV.
Sixty women of the reproductive age with uterine myoma were divided
into two groups. Thirty patients of the study group received combined
therapy plus general magnetotherapy (GMT). Patients of the control group
received only combined treatment. Ultrasound investigation registered a
reduction in the size of myoma nodes by 16.7% in the study group, while
in the controls myoma size did not change (p < 0.05). 1-year
follow-up data for the study group demonstrated no cases of the myoma
growth while 16.6% of the controls showed growth of myoma nodes, in 6.6%
of them supravaginal myoma amputation was made for rapidly growing
myoma.
Urologiia. 2004 Mar-Apr;(2):20-2.
Combined therapy of interstitial cystitis using the “Aeltis-Synchro-02-Iarilo” device.
[Article in Russian]
Kalinina SN, Molchanov AV, Rutskaia NS.
Multiple modality therapy of interstitial cystitis (IC)–the disease
characterized by nicturia, pelvic pains, imperative pollakiuria–is
considered. As IC nature is not well known, its treatment remains
empiric. Among the underlying causes, most probable are autoimmune,
allergic, infectious, neurological, vascular. Therefore, the treatment
should be multi-modality. Most usable now is combined chemotherapy.
Perspective is also IC treatment with medicines in combination with
physiotherapy (electromagnetolaser AELTIS-SYNCHRO-02-YARILO”).
Endovesical electrophoresis can be also applied.
The effect of a low-frequency magnetic field on the
clinico-immunological indices of patients with chronic inflammatory
diseases of the organs of the female genital system.
Low-frequency magnetic field generated by the vaginal inductor used
in 120 females with chronic genital inflammation promoted a decrease in
leukocytosis, elevation of total population of T-lymphocytes, inhibition
of high proliferative activity in PHA test. However, marked
immunocorrection was not reached.
Eur J Surg Suppl. 1994;(574):83-6.
Electrochemical therapy of pelvic pain: effects of pulsed electromagnetic fields (PEMF) on tissue trauma.
Jorgensen WA, Frome BM, Wallach C.
International Pain Research Institute, Los Angeles, California.
Unusually effective and long-lasting relief of pelvic pain of
gynaecological origin has been obtained consistently by short exposures
of affected areas to the application of a magnetic induction device
producing short, sharp, magnetic-field pulses of a minimal amplitude to
initiate the electrochemical phenomenon of electroporation within a 25
cm2 focal area. Treatments are short, fasting-acting, economical and in
many instances have obviated surgery. This report describes typical
cases such as dysmenorrhoea, endometriosis, ruptured ovarian cyst, acute
lower urinary tract infection, post-operative haematoma, and persistent
dyspareunia in which pulsed magnetic field treatment has not, in most
cases, been supplemented by analgesic medication. Of 17 female patients
presenting with a total of 20 episodes of pelvic pain, of which 11
episodes were acute, seven chronic and two acute as well as chronic, 16
patients representing 18 episodes (90%) experienced marked, even
dramatic relief, while two patients representing two episodes reported
less than complete pain relief.
Urol Nefrol (Mosk). 1996 Sep-Oct;(5):10-4.
Magnetic-laser therapy in inflammatory and postraumatic lesions of the urinary system.
[Article in Russian]
Loran OB, Kaprin AD, Gazimagomedov GA.
The authors discuss disputable problem of renal and ureteral tissue
after trauma or inflammation. These cause irreversible morphological
changes in the tissue. Poor results of the standard therapy urged the
authors to try magnetic-laser therapy in urological clinic. The
technique has been developed on experimental animal models. The
resultant morphological characteristics of ureteral wall and parenchyma
support the validity of magnetic-laser therapy in urological practice.
A permanent magnetic field in the combined treatment of acute endometritis after an artificial abortion.
[Article in Russian]
Strugatskii VM, Strizhakov AN, Kovalenko MV, Istratov VG, Iakubovich DV.
117 patients with acute endometritis after induced abortion were
examined using markers of wound process phases and treated according to
the original method. This consists in combination of constant magnetic
field with other modalities. Application of the constant magnetic field
produced a significant clinical response and reduced the hospital stay
through positive effect on healing of the endometrial wound.
Zh Nevropatol Psikhiatr Im S S Korsakova. 1989;89(4):41-4.
Current methods of pathogenetic therapy of infectious-allergic polyradiculoneuritis.
[Article in Russian]
Neretin VIa, Ki’riakov VA, Sapfirova VA, Agafonov BV.
This is a survey of the experience in using corticosteroids,
plasmapheresis, immunodepressants, hyperbaric oxygenation, laser and
magnetotherapy in treating the infectious-allergic Guillain-Barre
polyradiculoneuritis. The indications and counter-indications to
individual techniques are presented as related to the character and
course of the disease. The principles of interrelation of these
techniques with other drug and physical therapies are discussed. The
authors infer that combination of plasmapheresis with corticosteroids is
the best for acute polyradiculoneuritis and prolonged use of
maintenance doses of corticosteroids and immunodepressants, physical
methods and gymnastics are recommended for chronic polyradiculoneuritis.
1
Siegfried Weller Institute for Trauma Research,
Eberhard-Karls-Universität Tübingen, Schnarrenbergstr. 95, D-72076,
Tübingen, Germany. sabrina.ehnert@med.uni-tuebingen.de.
2
Sachtleben GmbH, Hamburg, Spectrum UKE, Martinistraße 64, D-20251, Hamburg, Germany.
3
Siegfried Weller Institute for Trauma Research,
Eberhard-Karls-Universität Tübingen, Schnarrenbergstr. 95, D-72076,
Tübingen, Germany.
4
Wuhan Union Hospital, Tongji Medical College, Huazhong University of
Science and Technology, Jiefang Dadao 1277#, 430022, Wuhan, China.
Abstract
Recently, we identified a specific extremely low-frequency pulsed
electromagnetic field (ELF-PEMF) that supports human osteoblast (hOBs)
function in an ERK1/2-dependent manner, suggesting reactive oxygen
species (ROS) being key regulators in this process. Thus, this study
aimed at investigating how ELF-PEMF exposure can modulate hOBs function
via ROS. Our results show that single exposure to ELF-PEMF induced ROS
production in hOBs, without reducing intracellular glutathione.
Repetitive exposure (>3) to ELF-PEMF however reduced ROS-levels,
suggesting alterations in the cells antioxidative stress response. The
main ROS induced by ELF-PEMF were •O2– and H2O2,
therefore expression/activity of antioxidative enzymes related to these
ROS were further investigated. ELF-PEMF exposure induced expression of
GPX3, SOD2, CAT and GSR on mRNA, protein and enzyme activity level.
Scavenging •O2– and H2O2 diminished the ELF-PEMF effect on hOBs function (AP activity and mineralization). Challenging the hOBs with low amounts of H2O2 on
the other hand improved hOBs function. In summary, our data show that
ELF-PEMF treatment favors differentiation of hOBs by producing non-toxic
amounts of ROS, which induces antioxidative defense mechanisms in these
cells. Thus, ELF-PEMF treatment might represent an interesting adjunct
to conventional therapy supporting bone formation under oxidative stress
conditions, e.g. during fracture healing.
Med Pr. 2014;65(3):343-9.
Effect of extremely low frequency magnetic field on glutathione in rat muscles.
[Article in Polish]
Ciejka E, Jakubowska E, Zelechowska P, Huk-Kolega H, Kowalczyk A, Goraca A.
Abstract
BACKGROUND:
Free radicals (FR) are atoms,
molecules or their fragments. Their excess leads to the development of
oxidizing stress, the cause of many neoplastic, neurodegenerative and
inflammatory diseases, and aging of the organism. Industrial pollution,
tobacco smoke, ionizing radiation, ultrasound and magnetic field are the
major FR exogenous sources. The low frequency magnetic field is still
more commonly applied in the physical therapy. The aim of the presented
study was to evaluate the effect of extremely low frequency magnetic
field used in the magnetotherapy on the level of total glutathione,
oxidized and reduced, and the redox state of the skeletal muscle cells,
depending on the duration of exposure to magnetic field.
MATERIAL AND METHODS:
The male rats, weight of 280-300 g,
were randomly devided into 3 experimental groups: controls (group I) and
treatment groups exposed to extremely low frequency magnetic field
(ELF-MF) (group II exposed to 40 Hz, 7 mT for 0.5 h/day for 14 days and
group III exposed to 40 Hz, 7 mT for 1 h/day for 14 days). Control rats
were kept in a separate room not exposed to extremely low frequency
magnetic field. Immediately after the last exposure, part of muscles was
taken under pentobarbital anesthesia. Total glutathione, oxidized and
reduced, and the redox state in the muscle tissue of animals were
determined after exposure to magnetic fields.
RESULTS:
Exposure to low magnetic field: 40 Hz,
7 mT for 30 min/day and 60 min/day for 2 weeks significantly increased
the total glutathione levels in the skeletal muscle compared to the
control group (p < 0.001).
CONCLUSIONS:
Exposure to magnetic fields used in
the magnetic therapy plays an important role in the development of
adaptive mechanisms responsible for maintaining the oxidation-reduction
balance in the body and depends on exposure duration.
Long-standing research on electric
and electromagnetic field interactions with biological cells and their
subcellular structures has mainly focused on the low- and high-frequency
regimes. Biological effects at intermediate frequencies between 100 and
300 kHz have been recently discovered and applied to cancer cells as a
therapeutic modality called Tumor Treating Fields (TTFields). TTFields
are clinically applied to disrupt cell division, primarily for the
treatment of glioblastoma multiforme (GBM). In this review, we provide
an assessment of possible physical interactions between 100 kHz range
alternating electric fields and biological cells in general and their
nano-scale subcellular structures in particular. This is intended to
mechanistically elucidate the observed strong disruptive effects in
cancer cells. Computational models of isolated cells subject to TTFields
predict that for intermediate frequencies the intracellular electric
field strength significantly increases and that peak dielectrophoretic
forces develop in dividing cells. These findings are in agreement with
in vitro observations of TTFields’ disruptive effects on cellular
function. We conclude that the most likely candidates to provide a
quantitative explanation of these effects are ionic condensation waves
around microtubules as well as dielectrophoretic effects on the dipole
moments of microtubules. A less likely possibility is the involvement of
actin filaments or ion channels.Keywords: electric fields, biological cells, cancer cells, microtubules, ions, TTFields
1. Introduction
The effects of external electric fields
on biological cells have been extensively studied both in the direct
current (DC) and alternating current (AC) cases [1].
In order to elucidate possible impact of electric fields on cells,
various experimental assays as well as analytical and computational
models have been developed in the past. Experimentally obtained findings
were further translated into biomedical applications. While DC or
low-frequency AC fields are used to induce stimulation of excitable
cells through membrane depolarization or to promote wound healing,
high-frequency AC fields are associated with tissue heating and membrane
rupture, thus finding its application in diathermy or ablation
techniques.
Intermediate-frequency AC electric fields in the kHz to MHz
range were commonly assumed to lead to no significant biological
effects [1]. However, in a major breakthrough paper, Kirson et al. [2]
reported the discovery that low-intensity (1–3 V/cm), intermediate
frequency (100–300 kHz) electric fields have a profoundly inhibitory
effect on the growth rate of various mammalian tumor cell lines [2,3,4].
This discovery has been translated into a clinical application termed
Tumor Treating Fields (TTFields). Based on the results of a Phase III
clinical trial [5],
TTFields have been approved by the United States Food and Drug
Administration (FDA) in 2011 for the treatment of recurrent glioblastoma
multiforme (GBM) and their efficacy in treating other solid tumor types
is currently being investigated clinically [6]. In late 2015, TTFields were also approved for newly diagnosed GBM patients in combination therapy with temozolomide [7,8] due to significantly increased survival times.
It should be noted that electromagnetic (EM) fields may
affect the regulation of cellular growth and differentiation, including
the growth of tumors [9,10].
Both static magnetic and electric fields have altered the mitotic index
and cell cycle progression of a number of cell types in various species
[10].
EM low-frequency fields in the range of 50–75 Hz cause perturbations in
the mitotic activity of plant and animal cells and a significant
inhibitory effect on mitotic activity occurs early during exposure [11,12,13].
While the field amplitudes used are consistent with those of interest
to this report, the frequencies are orders of magnitude lower.
The reduction in the cell number due to an application of
TTFields was studied by in vitro experiments with various cancer cell
lines. A significant prolongation of mitosis was predicted, where
treated cells remain stationary at metaphase for several hours, which
was accompanied by abnormal mitotic figures as well as membrane rupture
and blebbing leading to apoptosis [2,3].
Furthermore, these experiments showed that the inhibitory effect
increases with an increasing electric field intensity, resulting in a
complete proliferation arrest of rat glioma cells after 24 h exposure to
a field intensity of 2.25 V/cm. Additionally, the effects of TTFields
have been shown to be frequency-dependent, with a cancer cell
line-specific peak frequency of the maximal inhibitory effect, e.g., 200
kHz for glioma cells [3]. Following these experimental results, two specific mechanisms of action of TTFields have been proposed [2,3,4] which we describe below.
Firstly, the applied field is expected to interfere with
proper microtubule (MT) formation preventing a functioning mitotic
spindle, due to the force of interactions with the large intrinsic
dipole moments of the tubulin dimers [14,15,16]
that make up MTs. It has been hypothesized that the tubulin dimers
might align parallel to the direction of the applied electric field,
rather than along the MT axis. Secondly, the cellular morphology during
cytokinesis gives rise to a non-uniform intracellular electric field,
with a high density at the cleavage furrow between the dividing cells.
This non-uniform field leads to the development of dielectrophoretic
(DEP) forces [17]
acting on polarizable macromolecules such as MTs, organelles and all
charged structures present in the cell, such as ions, proteins or DNA.
Thus, TTFields are considered to be suitable as a novel
anti-mitotic cancer treatment modality. In fact, it has been suggested
by numerous researchers that endogenous electric fields may play a key
role during mitosis. Similar to Cooper [18], Pohl et al. [19]
proposed that the onset of mitosis is associated with a ferroelectric
phase transition, which establishes an axis of oscillation for the
cellular polarization wave. The mitotic spindle apparatus would
delineate the polarization field with MTs lined up along the electric
field lines. The poles are expected to experience the highest field
intensities while the equatorial plane is likely to provide a nodal
manifold for the fields. Consequently, the chromosome condensation
during this transformation was predicted to be induced by the static
dielectric polarization of the chromatin complex as a result of the
cellular ferroelectric phase transition. These conclusions have been
supported by experimental evidence for peak EM activity during mitosis [20,21]
and by physical modeling of the electrostatic forces generated by MTs
which generate mechanical force required for chromosome segregation
during mitosis and influence chromosomal motion [15,16,22]. A detailed review of this aspect can be found elsewhere [1].
Put together, there is reasonable evidence that especially
during mitosis, electric field effects are relevant for the functioning
of a dividing cell, especially in the creation of the mitotic spindle.
However, to date a rigorous quantitative analysis of the magnitude of
these effects within cells exposed to TTFields has not been performed.
Furthermore, an analysis of how TTFields might interact with subcellular
structures has also never been reported. In a quantitative model, which
attempts to explain these effects, an energetic constraint, both from
below and above, must be kept in mind. Firstly, for an effect to be of
significance at a molecular level, its interaction energy must exceed
thermal energy, i.e., kT per degree of freedom (i.e., 4 × 10?21 J).
Otherwise, thermal fluctuations will disrupt the action of electric
fields. Secondly, it must not produce so much thermal energy as to
seriously increase the temperature of the cell. In terms of practical
comparisons, a cell generates approximately 3 × 10?12 W of power (3 × 10?12 J/s),
much of which is used to maintain a constant physiological temperature.
This is found from a simple estimate of energy production by the human
body which is 100 W divided by the number of cells in the body which is
approximately 3 × 1013 [23].
In terms of subcellular forces at work, a minimal amount of useful
force at a nanometer scale is 1 pN. Motor proteins generate forces on
the order of several pN. A force of 1 pN applied to a tip of a
microtubule may be used to bend it by as much as 1 µm [24]. Below, we review electric conduction effects for subcellular structures of interest.
The paper is structured as follows. In Section 2,
we review what is known about the shape and intensity of the electric
field within cells exposed to externally applied electric fields,
focusing on cells during mitosis. As a preparation for following
sections, Section 3 offers a general introduction to subcellular electrical conduction and electrostatics. Section 4 and Section 5 are
devoted to a comprehensive review of the literature concerned with the
effects of electric fields on biopolymers, and with the identification
of additional mechanisms by which TTFields might interact with cells. Section 4 covers electric field interactions with the cell membrane and the cytosol, whereas the focus of Section 5 penetrates
deeper into the cell, shedding light on the electric field effects on
subcellular structures of interest, i.e., microtubules (MTs), actin
filaments (AFs), ionic charges and DNA. Finally, in Section 6,
we present a discussion about the significance of our findings and
about future directions of research that should be undertaken in this
area. We hope this paper will set a solid theoretical foundation for
future studies into the biophysics of TTFields.
2. Induced Electric Fields within Biological Cells in Mitosis
The topic of induced electric fields in
and around biological cells subject to DC or AC fields has been
investigated for decades. The preliminary and most popular studies on
the analytical description of steady-state trans-membrane potential
induced on spherical cells go back to the work of H.P. Schwan and
colleagues [25].
Arguments were presented to account for the influence of the membrane
conductance, surface admittance and spatial charge effects [25], as well as for the geometric and material properties of the cell and the surrounding medium [26].
The impact of external electric fields on a living cell significantly
depends on the cell’s shape. Concerning analytical solutions for
non-spherical cell shapes, many authors proposed appropriate adaption of
the governing equations going back to the work presented in Reference [27].
Later models aimed to study electric polarization effects on oblate and
prolate homogeneous and single-shell spheroids have been developed [28]. They were later extended to arbitrarily oriented cells of the general ellipsoidal shape [29].
Importantly, the induced electric field inside a spherical cell is
uniform, whereas increasing non-uniformities are predicted for
deviations of the regular shape.
Another important aspect is the frequency-dependency of
the induced trans-membrane voltage and thus also the intracellular field
strength, as predicted by the above-mentioned studies and additional
research reported elsewhere [30,31,32,33,34].
For low frequencies, the intracellular space is shielded to a large
degree from extracellular electric fields. For example, the electric
field strength inside a typical spherical cell is approximately five
orders of magnitude lower than that outside the cell [35,36].
However, as the frequency of the field increases, the high membrane
field gain diminishes, allowing for higher field intensities to
penetrate into the cell.
Recently, Wenger et al. [37]
developed a computational model to study the application of TTFields to
isolated cells during mitosis, specifically during metaphase and at
different stages of cytokinesis. Comsol Multiphysics (www.comsol.com) was used to solve for the scalar electric potential V for
frequency ranges between 60 Hz and 10 GHz. With voltages of opposite
signs set as boundary conditions, a uniform field of 1 V/cm was induced
in the model domain. Following 3D confocal microscopy findings [38,39],
the metaphase cell was represented by a sphere with a 10 µm radius and
three different stages of cytokinesis were modeled with increased
distance between the elliptical mother and daughter cell (see Figure 1,
left panel). Three model domains, the extracellular space, the
cytoplasm and the membrane, were assigned typical dielectric properties,
electrical conductivity and relative permittivity [37].
Figure 1
(Left) Schematic diagrams of the cell geometries
for metaphase and three stages of cytokinesis. Black lines indicate the
electric field contours. (Right) The maximum intracellular electric field strength in V/cm plotted as a function of field frequency …
For a spherical cell during metaphase, the modeling
results predict that for frequencies lower than 10 kHz only small
changes of the field are detected and the intracellular field strength, Ei, almost equals zero. A first significant increase of Ei is observed at approximately 200 kHz, and Ei increases rapidly as the frequencies increase above this value. This can be seen in the inset of the right panel in Figure 1,
which shows a zoomed view of the blue M-phase cell. This transition
region depends on the dielectric properties of the cell and its
membrane. Nonetheless, above 1 MHz electric current is shunted across
the membrane and the impedance is dominated by the cytoplasm. Thus, for
an increasing frequency, the electric field inside the cell is augmented
and at 1 GHz the cellular structure becomes ‘‘electrically invisible’’
as previously reported [33]. The directions of the electrical field near the cell membrane resemble already predicted results [40].
The model further showed that within the dividing cell the
intracellular electric field distribution is non-uniform with highest
field intensities at the cleavage plane (black lines in the left panel
of Figure 1).
These maximum intensities are much higher than the applied field and
appear for frequencies in the range 100–500 kHz depending on the stage
of cytokinesis, i.e., how far the cell division has already progressed.
The corresponding curves are plotted in the right panel of Figure 1, where the highest maximum intracellular field strength of ~22 V/cm is observed for the cell in late cytokinesis.
Due to the inhomogeneity in the electric field
distribution, significant dielectrophoretic (DEP) forces are expected to
develop within the cell and these DEP forces are thought to be
important factors in the mechanism of action of TTFields [41].
This DEP force causes the motion of polarizable particles as a result
of the interaction of a non-uniform electric field with their induced
dipole moment F=p??E [42]. The DEP force is proportional to the volume of the particle v, its effective polarizability ? and the square of the gradient of the electric field according to: ?F DEP?=1/2?v?Re[?(E˜??)E˜?] using complex phasor notation [42,43].
Thus, the magnitude of the DEP force component is proportional to the
magnitude of the gradient of the squared electric field, |F|????|E|2?? in (V2/m3).
The DEP force component showed well-defined peak frequencies at 500,
200, and 100 kHz, respectively, for the three stages considered, from
the earliest to the latest stage [37]. This coincides with the peaks of the maximum electric field inside the cell, which are presented in the right panel of Figure 1.
Apart form testing different field intesities, the
computational study tested another aspect of TTFields. Namely, it has
been shown that the optimal frequency for the inhibitory effect of
TTFields is inversely related to cell size [2,44] and that cell volume is increased in almost all cell lines treated with TTFields [45].
The simulation results predicted that the above-mentioned peak
frequencies decrease and converge as a function of an increasing cell
radius. The corresponding maximum values of the DEP force component also
decrease with an increasing cell size with equal decay rates for all
cytokinesis stages [37].
In summary, these results obtained by computational
modeling confirmed several predicted outcomes of the application of
TTFields to biological cells. During metaphase a uniform non-zero Ei is
induced. Depending on cell properties, the frequency window of the
predicted transition range might be shifted. During cytokinesis, a
non-uniform Eiis induced with a substantially
increased strength at the cleavage furrow. Frequency-, cell size-, and
field-intensity dependences were confirmed.
Experimental validation of the predicted
induced field strength values would be of great interest. Electric field
strengths have typically only been able to be measured inside membranes
with voltage dye and patch-clamp techniques. A promising technique by
Tyner et al. [46]
reported the generation of a nanovoltmeter that can report local
electric fields in the cell and its use would be ideal to calibrate the
strength and local distribution of electric fields in the presence of
externally applied AC electric fields.
3. Subcellular Electrical Conduction and Electrostatics
3.1. Protein Conduction
Biological polymers are made up of
various proteins, such as actin and tubulin, or nucleic acids as is the
case of DNA and RNA. These structures have uncompensated electrical
charges when immersed in water but ionic solutions such as the cytoplasm
provide a bath of counter-ions that at least partially neutralize the
net electric charge. This, however, results in dipolar and higher-moment
electric field distributions complicating the situation greatly.
Biological water is also believed to create structures with ordered
dipole moments and complex dynamics at multiple scales [47],
which adds to the complexity of subcellular electric field effects.
Additionally, free ions endow the cell with conducting properties along
well-defined polymeric pathways as well as in a diffusive way. Membranes
support strong electric fields (on the order of 105 V/cm),
which, due to counter-ion attraction to charged surfaces in solution,
result in Debye screening. This causes an exponential decay of these
electric fields on a nm scale [48] but not their complete disappearance when measured in the cell interior (hence a field strength of 105 V/cm decreases to approximately 0.01 V/cm over 100 nm).
The idea that proteins in organisms may have semiconducting properties dates back many decades [49,50] but protein conductivity has been found to be strongly dependent on the hydration state of proteins [51]. Electrical properties of cells and their components were promoted by Szent-Györgyi [52,53],
but significant experimental challenges of measuring electric fields
and currents at a sub-cellular level impeded progress in this field. The
development of more precise experimental tools in the area of
nanotechnology holds great promise for rapid progress in the near future
[54,55]. Owing to the fact that there have been many previous reviews of electromagnetic effects in biology [1,56,57,58],
here we mainly focus on the electrical properties of MTs, actin AFs,
ion channels, cytoplasmic ions and DNA with special interest into
dynamical electrical properties involving AC fields in the range of 100
kHz. A crucial role of water in the transmission of electrical pulses
due to the structure imparted by hydrophilic surfaces [59]
is also worth noting. Charge carriers related to protein
semi-conduction have largely been electrons, protons as well as ions
surrounding proteins in physiological solution. AFs and MTs have been
implicated in facilitating numerous electrical processes involving ionic
and electronic conduction [60,61] and have been theorized to support dipolar and/or ionic kink-like soliton waves traveling at speeds in the 2–100 m/s range [14,62].
Due to strong coupling between electrical and mechanical degrees of
freedom in proteins, mechano-electric vibrations of MTs have been
modeled both analytically and computationally [63,64]. Electric fields generated by MTs have been modeled extensively and reviewed recently [65,66,67], although experimental measurement of these fields remains extremely difficult, especially in a live cellular environment.
3.2. Electrostatic Interactions Involving Charges and Dipoles of Tubulin
The net charge on a tubulin dimer depends on pH and changes from +5 at pH 4.5 to 0 at pH 5 and drops to ?30 at pH 8 [68].
However, in the cytoplasm, a vast majority of electrostatic charges are
screened over the distances greater than the Debye length (which varies
between 0.6 and 1.5 nm depending on the ionic strength). Therefore,
calculating the force due to an electric field of a static electric
field with a strength of 1 V/cm acting on a 10 µm-long microtubule, we
find from F = qE, with q = 10?13 C for unscreened charges, that results in F =
10 pN assuming the field is largely undiminished when penetrating a
cell, which is in general a major oversimplification. This latter issue
will be addressed at the end of this review. Even if the force is
essentially unchanged, the Debye screening of electrostatic charges
means that less than 5% of the charge remains exposed to the field
resulting in a net force of at most 0.5 pN, most likely insufficient to
exert a major influence on the cytoskeleton. If the field oscillates
rapidly, the net force would cancel out over the period of these
oscillations, i.e., on a time scale of microseconds or less.
The next aspect of MT electrostatics is the effect on the
dipole moments of tubulin dimers and of entire MTs. The dipole moment of
tubulin (excluding the very flexible and dynamic C-termini which we
discuss separately below) has been estimated to be between 566 debye for
the ?-monomer and 1714 for the ?-monomer [69].
However, this is also strongly tubulin-isotype dependent, so these
numbers vary a lot between various tubulin isotypes from 500 to 4000
debye [70]. Note that 1 debye is a unit of electric polarization and is equal to 3.33 × 10?30 Cm. Therefore, taking the dipole moment of a free tubulin dimer as p = 3000 debye as a representative number, we find the interaction energy U with an electric field of E = 1 V/cm, and obtain U= ?pE, and hence U = 10?24 J. This is clearly too small (4000 times smaller than thermal energy kT)
to affect the dynamics of an individual tubulin dimer. However, a
single MT contains 1625 dimers per 1 µm of its length, so it could
eventually accumulate enough net dipole strength to be significantly
affected by the field. Unfortunately, this is very unlikely because of
the almost perfect radial symmetry of tubulin dipole arrangements in an
MT, which has been predicted by a computer simulation [70].
The individual dipole moments of constituent dimers will almost
perfectly cancel out in the radial arrangement of an MT cylinder. There
is a small non-cancellation effect along the MT axis but this amounts to
less than 10% of the next dipole moment, hence it is doubtful that an
entire MT can be aligned in electric fields with intensities lower than
10 V/cm. Unless one uses time-dependent fields (e.g., those used in
Reference [71]), much stronger fields are needed for static effects. To put it another way, the torque ? between a dipole moment of an MT, p, and an external electric field, E, is proportional to their vector product: ? = pxE.
For the force to have a meaningful effect on a microtubule, it should
exceed 1 pN for lever arm on the order of 1 µm giving a torque of 10?18 Nm.
With fields on the order of 1 V/cm, and a dipole moment of 3000 debye
per tubulin dimer, even if these dipoles were perfectly aligned, it
would result in a 1 µm MT only experiencing a torque of 10?21 Nm, which is approximately 1000 times too low to be of relevance.
Various special situations involving electrostatic effects on MTs were calculated earlier [68]. Note that a force between a charge and an electric field is given by F = qE(x) where E(x)
is screened exponentially over the Debye length, which is approximately
1 nm. Hence, a test charge of +5e a distance of 5 nm from the MT
surface for a 10-µm MT, experiences a force of 12 pN in water and 1 pN
in ionic solution. A tubulin dimer with a dipole of 3000 debye in the
vicinity of a microtubule experiences an electrostatic energy of 3 meV.
MT-MT interactions due to their net charges with Debye screening
accounted for lead to a net force of 9 pN when separated by 40 nm
resulting in net repulsion between them. However, at longer distances
attractive forces prevail and the corresponding dipole-dipole attraction
at 90 nm is only 0.08 pN. The authors of the references [15,16,22] estimated the maximum electrostatic force in the mitotic plate, which was given as F = 6n2 pN per MT where n is the number of elementary charges on each protofilament. Since F is
estimated to be 1–74 pN for a typical MT, the estimate is 0.4–3.5
uncompensated elementary charges per protofilament. The range of values
of the forces involved is certainly within the realm of possible force
requirements for chromosome segregation (about 700 pN per chromosome).
3.3. MT Conductivity
The building block of an MT is a tubulin
dimer, containing approximately 900 amino acid residues with a combined
mass of 110 kDa (1 Da is the atomic unit of mass, 1 Da = 1.7 × 10?27 kg).
Each tubulin dimer in an MT has a length of 8 nm, along the MT cylinder
axis, a width of about 6.5 nm and the radial dimension of 4.6 nm. The
inner core of the cylinder, known as the lumen, is approximately 15 nm
in diameter. MTs have been predicted to exhibit intrinsic electronic
conductivity as well as ionic conductivity along their length [72].
MTs have a highly electro-negatively charged outer surface as well as
C-terminal tails (TTs), resulting in a cloud of counter-ions surrounding
them. Experiment and theory demonstrate that ionic waves are amplified
along MTs [72,73].
Since MTs form a cylinder with a hollow inner volume (lumen), MTs have
also been theorized to have special conducting properties involving the
lumen [55]
but there has been no direct experimental determination of the electric
properties of the MT lumen. Many diverse experiments were performed to
date in order to measure the various conductivities of MTs, with a range
of results largely dependent on the experimental method, and this has
been reviewed elsewhere [74].
Interestingly, Sahu et al. [75,76]
measured conductivity along the periphery of MTs, where the DC
intrinsic conductivities of MTs, from a 200 nm gap, were found to be
approximately 10?1 to 102 S/m. Unexpectedly, MTs
at certain specific AC frequencies (in several frequency ranges) were
found to be approximately 1000 times more conductive, exhibiting
astonishing values for the MT conductivities in the range of 103to 105 S/m [55,76].
Some resonance peaks for solubilized tubulin dimers were reported as:
37, 46, 91, 137, 176, 281, and 430 MHz; 9, 19, 78, 160, and 224 GHz; and
28, 88, 127, and 340 THz. However, for MTs, the corresponding resonance
peaks were given as: 120, 240, and 320 kHz; 12, 20, 22, 30, 101, 113,
185, and 204 MHz; and 3, 7, 13, and 18 GHz. Therefore, for MTs there is
some overlap with the 100 kHz range indicating a possible independent
confirmation of the sensitivity of MT AC conductivity to this electric
field frequency range. These authors showed experimental evidence that
the high conductivity of the MT at specific AC frequencies only occurred
when the water channel inside the lumen of the MT remained intact [55].
Electro-orientation experiments involving MTs have shown
an increased ionic conductivity (0.150 S/m) compared to the buffer
solution free of tubulin by as much as 15-fold [77]. MTs exposed to low frequency AC fields (f <
10 kHz) exhibit a flow motion due to ionic convection. However, for
frequencies above 10 kHz this convection effect is absent. Electric
fields with intensities above 500 V/cm and frequencies in the range of
10 kHz–5 MHz, are able to orient MTs in solution. As a point of
interest, this frequency range overlaps with the range used by Kirson et
al. [2]. However, the intensities of the electric fields used are substantially higher. For instance, a 900 V/cm field with f = 1 MHz was able to align MTs within several seconds [77]. Impedance spectroscopy enabled the measurements of the dielectric constant of tubulin as ? = 8.41 [78]. Uppalapati et al. [79]
exposed taxol-stabilized MT’s in solution to an AC field, which
exhibited electro-osmotic and electro-thermal flow, in addition to MT
dielectrophoresis effects. Interestingly above f = 5 MHz, electro-hydrodynamic flows were virtually eliminated, and the conductivity of MTs was estimated at 0.25 S/m.
Priel et al. demonstrated MTs’ ability to amplify ionic
charge conductivity, with current transmission increasing by 69% along
MTs [60],
which was explained by the highly negative surface charge density of
MTs that creates a counter-ionic cloud subjected to amplification along
the MT axis [60]. From Priel et al.’s conductance data, the approximate ionic conductivity of MTs is found to be an astonishing 367 S/m [74].
Below, in the second part of this review, we quantitatively assess AC
electric fields on these ionic conductivity experiments, which are
expected to be sensitive to the electric field frequencies in the 100
kHz to 1 MHz range.
The multiple mechanisms of MT conductance
provide ample possibility to explain the varied reports on MT
conductivity in the literature. Ionic conductivity along the outer edge
of the MT, intrinsic conductivity through the MT itself, and possible
proton jump conduction and conductivity through the inner MT lumen have
all been suggested. It is conceivable that TTFields may affect ionic
conductivities along MTs as is argued below.
4. Collective Effects in the Membrane and Cytoplasm
4.1. Membrane Depolymerization Effects
The electric field across the membrane is on the order of 105 V/cm
(0.1 V over 8–10 nm), which is 4–5 orders of magnitude greater than
TTFields’ amplitude. Therefore, a direct effect of TTFields on cancer
cells’ membrane potential is expected to be very minor.
4.2. Ion Channel Conduction Effects
Liu et al. [80] reported activation of a Na+ pumping
mode with an oscillating electric field with a strength of 20 V/cm,
which is comparable to the fields of interest in this review, but at a
much higher frequency (1.0 MHz) than those of interest. Moreover,
neither K+ efflux nor Na+ influx was stimulated by
the applied field in the frequency range from 1 Hz to 10 MHz. These
results indicate that only those transport modes that require ATP
splitting under the physiological condition were affected by the applied
electric fields, although the field-stimulated K+ influx and Na+ efflux
did not depend on the cellular ATP concentration in the range 5 to 800
µM. Computer simulation of a four-state enzyme electro-conformationally
coupled to an alternating electric field [81,82] reproduced the main features of the above results.
Channel densities strongly vary among different neuronal
phenotypes reflecting different stabilities of resting potentials and
signal reliabilities. In model cell types such as in mammalian medial
enthorinal cortex cells, modeled and experimental results match best for
an average of 5 × 105 fast conductance Na+ and delayed rectifier K+ channels per neuron [83]. In unmyelinated squid axons counts can reach up to 108 channels per cell. In model channels such as the bacterial KcsA channel one K+ ion crosses the channel per 10–20 ns under physiological conductances of roughly 80–100 pS [84],
which is consistent with the frequencies of external electric
stimulation mentioned above. This allows for a maximum conduction rate
of about 108 ions/s. Estimating the distances between the
center of the channel pore and the membrane surface to scale along 5 nm
and assuming the simplest watery-hole and continuum electro-diffusion
model of channels, this would provide an average speed of 0.5 m/s per
ion. Ion transition occurs through a sequence of stable multi-ion
configurations through the filter region of the channels, which allows
rapid and ion-selective conduction [85].
The motion of ions within the filter was intensively studied applying
classical molecular dynamics (MD) methods (for a summary see Reference [86]) and density functional studies (e.g., [87]). MD methods used in these simulations solve Newton’s equations of motion for the trajectory of ions.
Time scales for the processes in ion channels can be estimated by the time for translocations (ttr) between two filter sites separated by ~0.3 nm, i.e., 5 × 10?10 s [88] and 5 × 10?11 s [87]. Transition rates (from potential mean force maps and the Kramer transition rate model [89]
are consistent with these numbers. Changes between a non-conductive and
a conductive state in the KcsA occur at a rate of 7.1 × 103 s?1, giving a life-time of the non-conducting state of 0.14 ms (~10?4 s) [89]. As the duration of the rather (stable) non-conducting state scales in the range 10?3–10?4 s and the within filter translocation time is on the order of 10?11 s, we can expect about 107 filter state changes during a non-conducting state and about 1010 switches
per second (10 GHz). Consequently, these time scales are incompatible
with those resulting from the effects of 100 kHz electric fields (10
?s).
4.3. Electric Field Effects on Cytoplasmic Ions
The cytoplasm provides a medium in which
fundamental biophysical processes, e.g., cellular respiration, take
place. Most biological cells maintain a neutral pH (7.25–7.35) and their
dry matter is composed of at least 50% of protein). The remaining dry
material is composed of nucleic acids, trace ions, lipids, and
carbohydrates. Most of the trace ions are positively charged. A few
metallic ions are found which are required for incorporation into
metallo-proteins, e.g., Fe2+, typically at nanomolar concentrations. In Table 1, we summarize the composition of the cytoplasm regarding the most abundant and important components.
Based on the above, we can estimate the net force on the total charge in the cytoplasm as F = qE, q = 4 × 1011 e and E =
1 V/cm, so the total force is approximately 6 µN, which is sufficient
to cause major perturbations in the cell interior. As discussed above,
this is strongly depended on the ability of the electric field to
penetrate into the cell’s interior, which is easier in the case of
non-spherical cells. The net outcome of these ionic oscillations away
and towards attractively interacting protein surfaces inside the
cytoplasm can be a concomitant series of oscillations of the structures
affected by the ionic clouds as schematically shown below.
The viscosity of cytoplasm is approximately ? = 0.002 Pa·s [90], hence we can estimate the friction coefficient for an ion in solution as ? = 6??r where r is the ionic radius (hydration shell radius) and find ? = 2 × 10?12 Pa·s·m. In an oscillating electric field of amplitude 1 V/cm and a frequency f = 200 kHz, an ion’s position will follow periodic motion given by: x(t) = 0.1·A·sin(2?ft),
i.e., will execute harmonic motion out of phase with the field, with
the same frequency and an amplitude A approximately 10% of the radius.
However, these ions are simultaneously subjected to the Brownian motion
due to their collisions with the molecules of the solvent.
To estimate the effect of an oscillating external electric
field on the diffusion of a single biomolecular particle (protein, DNA,
simple ion, etc.), the Langevin equation can be written down and
solved. In the Ito interpretation [91], the position Xt of such a particle is given as a function of time by [92]:
dXt= F(Xt)?dt+2kBT???????dWt
(1)
where ? is the friction coefficient of the particle, T=310 ? is the temperature and kB is the Boltzmann constant. The first term on the RHS of Equation (1) accounts for the influence of deterministic forces F(Xt). Assuming there is no interaction other than the coupling with an external electric field E(Xt), we can write F(Xt)=qE(Xt) where q is
the net charge of the particle. At intermediate frequencies, i.e.,
around 100–200 kHz, the wavelength is around 1000 m, which is obviously
much larger than the size of a typical cell. Thus, assuming no important
changes due to the dissipation of the field, E can be considered almost constant in a cellular environment: F(Xt)=qE(t).
The second term on the RHS represents the random motion, which is due
to the many kicks with the surrounding water molecules. Hence, dWt is usually given by [91]:
dWt~dt1/2 ?(t)
(2)
where ?(t) is
a random number, which follows a normal distribution with a mean equal
to 0 and a variance equal to 1. Since the Brownian motion is
proportional to dt1/2, an estimate of dt is needed to evaluate the influence of the external electric field over the thermal noise. The time step dt can
be estimated by the time interval between two series of collisions with
water molecules, each series being the sum of enough collisions so that
the outcome is approximately Gaussian. In other words, one can assume dt=dx/vH2O, where dx is the typical separation between two water molecules, i.e., dx=mH2O/?H2O??????????3 where mH2O is the mass of one water molecule and ?H2O is the mass density of water. Here, vH2O is the velocity of water molecules given by vH2O=3kBT/mH2O???????????. The use of the above parameters leads to a typical time step of dt~5.0×10?13 s.
The two terms in the RHS of Equation (1) above can be
compared to estimate the effect of an electric field over the thermal
noise. In the case of a spherical particle, we can assume ?=6??r, where the hydrodynamic radius is r=1.8 ? and the viscosity of the cytoplasm is ?=0.002 Pa·s [90]. By taking q=1 e (a single ion) and E=E0cos2?ft with E0=1 V/cm, it turns out that the amplitude of the coupling term associated with the electric field is qE0/?=2.36 × 10?6 m/s. On the other hand, the noise coefficient is 2kBT/???????? (dt)?1/2=50.2 m/s when the estimate obtained above is used: dt~5.0×10?13 s, which is much larger than the deterministic term. Even in the case of less frequent Brownian collisions, e.g., dt~10?6 s,
the noise coefficient is 0.035 m/s which is still much larger than the
coupling with the electric field, meaning that an electric field of
amplitude 1 V/cmhas an exceedingly small probability to influence the diffusion of a single Brownian particle even if the net charge q is 100–1000 times larger as in the case of a protein.
Alternatively, it can be shown that an oscillating electric field at intermediate frequencies with an amplitude of 1 V/cm has
no direct sizable effect on the diffusion of biomolecules by
considering an ensemble of molecules instead of a single Brownian
particle. Assuming a constant electric field E, the distribution of particles as a function of time is given by [93]:
P(x,t)= 12Dt????exp?????(x –x0?qEt?)22Dt????
(3)
Here, D=kBT/? is
the diffusion coefficient for one particle. From the above equation, a
typical time when the particles start to be drifted away because of the
electric field is t=2(kBT)?/(qE)2. For a single ion (q=1 e, r=1.8 ?), t=226.1 s, whereas for a typical globular protein (q~100 e, r~1.0 nm), t=0.13 s, which is much larger than the period of an electric field oscillating at hundreds of kHz.
For the sake of simplicity, we have not
discussed here how an electric field could induce conformational changes
in biomolecular structures, which would affect their charge
distributions and dipolar spectra, which, in turn, could modify their
diffusion by inducing new interactions with the surrounding molecules.
An estimate of such indirect effects would require careful
investigations of the studied system based on realistic MD simulations.
In this case, the external electric field can be either computationally
modeled by initializing the system with added kinetic energy in the
directions of the normal modes or by adding an extra coupling term to
the force field [94].
5. AC Electric Field Effects on Subcellular Structures
5.1. Electric Field Effects on MTs
Several experimental efforts were made aimed at measuring the electric field around MTs. Vassilev et al. [71]
observed alignment of MTs in parallel arrays due to the application of
electric fields with intensities of 0.025 V/cm and of pulsed shape. In
cell division, coherent polarization waves have been implicated as
playing the key role in chromosome alignment and their subsequent
separation [18,19]. Electric fields in the range of 3 V/cm were applied by Stracke et al. [75]
to suspended MTs, which moved at pH 6.8 from the negative electrode to
the positive one indicating a negative net charge, and an
electrophoretic mobility of about 2.6 × 10?4 cm2·V?1·s?1. The work of Uppalapati et al. [79]
covers the range of frequencies overlapping with TTFields, although the
amplitudes are much larger due to the voltage bias of 40 V across a
20-µm gap giving an electric field of 2 × 104 V/cm as opposed to 1 V/cm). Below 500 kHz, MTs flow toward the centerline of electrodes. The electro-osmotic force causes
the movement of the fluid in a vortex-like manner. This represents the
Coulomb force experienced by the ionic fluid due to the applied voltage.
The fluid flow velocity ? is proportional to the tangential component of the electric field Et, surface charge density ?, the solution’s viscosity ? and the inverse Debye length ? such that: ? = Et ?/??.
At lower frequencies, flow velocity is larger. On the other hand, due
to strong heating effects of the AC field, the electro-thermal force
causes motion of MTs along the length of the electrodes. Above 500 kHz
MTs flow toward the gap between the electrodes due to dielectrophoresis.
The DEP force experienced by MTs in a non-uniform electric field is
given by:
?FDEP?=14??m[?2?m(?p??m)+?m(?p??m)?2?2m+?2m]?|E|2
(4)
where the symbols with subscript “m” refer to the medium and “p”
to the particle. Hence, this process is largely driven by the
difference between the conductivities and permittivities of the MTs and
the medium, (?p ? ?m) and (?p ? ?m),
respectively. We predict that lowering the pH of the solution to the
isoelectric point of MTs around pH 5 should substantially reduce this
effect and additionally lowering the frequency will reduce it further
due to the dependence of the first term on the square of the frequency.
At ~5 MHz, the electro-osmotic and electro-thermal flow balance each
other out with the flow of MTs being solely due to dielectrophoresis. It
is important to compare the dielectrophoretic force to Brownian motion
in order to determine whether or not electric fields are sufficiently
strong to overcome random motion, i.e., to find out if the dielectric
potential exceeds the thermal energy, i.e.,
?r3?m[(?p??m)(?p+2?m)]E2>kT
(5)
where ?m is the dielectric constant of the medium and ?p is the dielectric constant of the particle. E is the electric field strength and r the
radius of the particle. Taking as an example a tubulin dimer in
solution and the corresponding values of the dielectric constants, one
finds that E must exceed 0.25 V/cm for the
field to be effective in orienting polarizable tubulin dimers.
Similarly, for a 10-µm long MT we replace the factor ?r3 with ?r2L, where r is the radius of a MT (12.5 nm) and L its length, to obtain a condition that E >
0.01 V/cm. Clearly, the electric field values of 1 V/cm (even if they
are screened by a large factor inside the cell) are sufficient to exert
electrophoretic effects on tubulin and MTs. The longer the MT, the more
pronounced the dielectrophoretic effect is predicted to occur.
Recently, Isozaki et al. [95]
used MTs labeled with dsDNA to manipulate the amount of net charge and
observe the mobility of these hybrid structures compared to control
where MTs where only labeled fluorescently with two different tags. It
was found for control MTs that the electrophoretic mobility is
approximately: 2 × 10?8 m2·V?1·s?1which is consistent with Stracke et al. [75].
For field strengths of approximately 1 V/cm, one can estimate the
average velocity of MT translocations as 2 µm/s. They also stated ?D = 0.74 nm as the Debye length, ? = 8.90 × 10?4 kg·m?1·s?1 and ? = 6.93 × 10?10C·V?1·m?1 as
the viscosity and dielectric constant of the buffer, respectively.
Importantly, they estimated the effective charges of the TAMRA- and
AlexaFluor 488-tagged tubulin dimer as 10 e? and 9.7 e?,
which obviously is only a fraction (approximately 20%–30%) of the
vacuum values but much larger than earlier experimental estimates.
Electrophoresis experiments were also performed by van den Heuvel et al.
[96], with electric field strengths of 40 V/cm, yielding MT electrophoretic mobility in the range of 2.6 × 10?8 m2·V?1·s?1, in line with previous reports. They found the effective charge of a tubulin dimer to be approximately 23 e?.
5.2. Tubulin’s C-Termini Dynamics and AC Electric Fields
Computer simulations demonstrate
that ionic waves can trigger C-termini to change from upright to
downward conformations initiating propagation of a travelling wave [97]. This wave is predicted to travel as a “kink” solitary wave with a phase velocity of vph = 2 nm/ps [97].
A typical time scale for C-termini motion is 100 ps, which is too fast
for the 100 kHz frequency range of TTFields. However, C-termini being
very flexible and highly charged (with approximately 40% of the
tubulin’s charge located there) are likely to dynamically respond to
electric fields as local changes of pH are correlated with positive and
negative electric field’s polarities, respectively. This effect can
cause MT instability as well as interference with motor protein
transport as discussed below. A stable dimer conformation is predicted
to have C-termini cross-linked between the monomers as shown in Figure 2.
Figure 2
A cross-linked conformation of C-termini stabilizes a straight
orientation of a tubulin dimer. A disruption of this conformation can
cause MT instability.
5.3. Ionic Waves along MTs and AC Electric Fields
Manning [98]
postulated that polyelectrolytes may have condensed ions in their
surroundings if a sufficiently high linear charge density is present on
the polymer’s surface [99]. The Bjerrum length, ?B,
is defined as the distance at which thermal fluctuations are equally
strong as the electrostatic interactions between charges in solution
whose dielectric constant is ? at a given temperature T in Kelvin. Here, ?0denotes the permittivity of the vacuum and kB is the Boltzmann constant. Counter-ion condensation occurs when the average distance between charges, b, is such that ?B/b = S>
1. In this case, the cylindrical volume of space depleted of ions
outside the counter-ion cloud surrounding the polymer functions as an
electrical shield. The “cable-like” electro-conducting behavior of such a
structure is supported by the polymer itself and the “adsorbed”
counter-ions, which are “bound” to the polymer in the form of an ionic
cloud (IC). Tuszynski et al. [68]
calculated an electrostatic potential around tubulin and extended this
to an MT, which demonstrated non-uniformity of the potential along the
MT radius with periodically repeating peaks and troughs along the MT
axis. Consequently, MTs have been viewed as “conducting cables” composed
of 13 parallel currents of ionic flux (corresponding to 13
protofilaments of MTs) and attracting an IC of positive counter-ions
close to its surface and along tubulin C-terminal tails (TT), while
negative ions of the cytosol are repelled away from the MT surface. The
thickness of the negative ion depleted area corresponds to the Bjerrum
length. An estimate of the respective condensate thickness ? of the
counter-ion sheath for the tubulin dimer (?TD) and C-termini (?TT) is ?TD = 2.5 nm and ?TT = 1.1 nm, as analyzed in [61]. Using a Poisson–Boltzmann approach, the capacitance of an elementary ring of an MT consisting of 13 dimers is found as [100]:
C0=2??0?lln(1+lBRIC)
(6)
where l stands for the length of a polymer unit and RIC = ?TD + ?TT for the outer radius of an IC. For a tubulin dimer: CTD = 1.4 × 10?16 F and for an extended TT: CTT = 0.26 × 10?16 F. Hence:
C0=C0+2×C0=1.92×10?16 F
(7)
Estimating the electrical resistance for a complete tubulin ring gives R0 = 6.2 × 107 ? [60,100].
Including the conductance of both nanopores through an MT surface
accounts for the leakage of IC cations into the lumen area and gives a
conductance G0, of a ring as G0 = ?1 + ?2 = (2.93 + 7.8) nS = 10.7 nS and the corresponding resistivity as R = 1/G0= 93 M?.
A simple equivalent periodic electric circuit simulating
one protofilament of an MT consists of a long ladder network composed of
elementary circuit units as shown in Figure 3 [61].
Figure 3
An effective circuit diagram for the n-th unit with characteristic
elements for Kirchhoff’s laws applied to a microtubule as an ionic cable
[61].
The longitudinal ionic current encounters a series of Ohmic resistors R0 for each ionic conduction unit (an MT ring). The nonlinear capacity with the charge Qn for the n-th site of the ladder is in parallel with the total conductance G0 of the two TTs of a dimer. Then using Kirchhoff’s law:
in?in+1=?Qn?t+G0?n,
(8)
?n?1??n=R0in,
(9)
we find the equations for the voltage propagation:
?Qn?t=C0??n?t?C0?0??n?C0?0?(t?t0)??n?t?2b0C0?n??n?t
(10)
Introducing an auxiliary function u(x, t) unifying the voltage and its accompanying IC current as:
un=Z1/2in=Z?1/2?n
(11)
with the characteristic impedance defined as:
Z=1?C0,
(12)
leads in the continuum limit to the electric signal propagation equation:
?2?u?x?l23?3u?x3?ZC0l?u?t+ZC0?0?l(t?t0)?u?t+2Z3/2b0C0lu?u?t?1l(ZG0+Z?1R0?ZC0?0?)u=0
(13)
The characteristic charging (discharging) time of an elementary unit capacitor C0through the resistance R0 is given by T0 = R0C0 with an estimate for T0 = 1.2 × 10?8 s and the characteristic propagation velocity of the ionic wave: v=l/T0 as v0 = 0.67 m/s. A standard travelling-wave with speed v, for the normalized function u(x, t), can
be used as a solution of the propagation equation, which is a soliton
that preserves its width but its amplitude decays over the length of
about 400 units corresponding to 3.2 µm, which is of the order of the MT
length. Interestingly, a characteristic time for this excitation can
readily be estimated as 1.2 × 10?5 s whose inverse, the frequency, f, is very close to the TTField value, i.e., 90 kHz. The maximum frequency allowed in this model is 68 MHz.
To summarize, ionic conduction along and away from charged
protein filaments such as MTs involves cable equations resulting from
equivalent RLC circuits surrounding each protein unit in the network.
Conduction along the filaments experiences resistance due to viscosity
in the ionic fluid. Capacitance is caused by charge separation forming a
double layer between the MT surface and ions with a distance separating
them comparable to the Bjerrum length. Inductance is caused by helical
nature of the MT surface and consequently, solenoidal flows of the ionic
fluid along and around the MT. The key numerical estimates of the RLC
circuit components are as follows [60]. For a single dimer: C = 6.6 × 10?16 F, R1 = 6 × 106 ? (along the MT), R2 = 1.2 × 106 ? (perpendicular to the MT) and L = 2 × 10?12 H.
These numbers can be used to estimate characteristic time scales for
the oscillations (LC) and exponential decay (RC) taking place in this
equivalent circuit. We obtain for decay times (? = RC) the following values: (a) ?1 = 10?8 s along the MT length and (b) ?2 = 10?9 s away from the MT surface. However, due a low value of inductance L, the corresponding time for electromagnetic oscillations is found using ?0 = (LC)1/2 as ?0 = 0.2 × 10?12 s
= 0.2 ps. Clearly, the oscillation times are too short for potential
effects with 100 kHz-range fields (the time of TTFields oscillations is
on the order of 5–10 µs). The decay times are much closer so we will
focus on these parameters. Repeating these calculations for a
microtubule of length l, we note that R1 scales with length of a microtubule, while R2 is length independent. The corresponding capacitance in both cases scales with length, therefore ?1 scales with length squared (l2) while ?2 scales
with length. To obtain actual values, we need to multiply the values
for a single ring by the number of rings in an MT. We use the values
found for a single ring, i.e., ?1 = 10?8 s and ?2 = 2 × 10?9 s
and scale them accordingly to estimate the length of MTs that could
experience resonant effects in terms of ionic currents along and away
from their surface. This way we find the scaling factor that leads to
the characteristic times on the order of 10 µs. Therefore, for
longitudinal effects, on the order of 50 rings, MTs only 400 nm long
would respond to 100 kHz stimulation. On the other hand, for ionic flows
pulsating radially around an MT, a 20-µm long MT would be required.
These results are very sensitive regarding the choice of parameter
values, especially the resistivity where diverse estimates can be found
in the literature. In general, there is strong overlap between the time
scales of ionic wave propagation and electric field stimulation. It is
conceivable that both effects play a role depending on the orientation
of the field vis a vis the geometry of mitotic spindles and the
MTs forming them. It appears that short MTs would be more sensitive to
the longitudinal wave generation by TTFields while long MTs should lead
to perpendicular wave generation.
Current densities should also be briefly discussed in relation to previously reported endogenous current densities, j, in cells, which range from 0.2 to 60 µA/ cm2 [101]. This translates into 0.002 < j < 0.6 A/m2. Since j = ?E where E =
1 V/cm and ? of the cytoplasm has a large range of values reported
between 0.1 and 100, we see that even taking the lower limit of 0.1
would result in ionic currents along MTs that would overwhelm the
intrinsic ion flows in a dividing cell. It is possible that these
externally stimulated currents cause a major disruption of the process
of mitosis and associated intra-cellular effects.
It is also worth mentioning that recently metabolic
oscillations in cells with a period of approximately 10 to 12 s, were
measured in vivo [102]
which is many orders of magnitude slower than any AC electric field
effects discussed here. Hence, it is safe to assume that there is a very
unlikely possibility of electric field effects in the 100 kHz range to
interfere with cellular metabolism.
Finally, it is interesting to address the
issue of the power dissipated due to a current flowing along an MT.
Again, we take as an example a 10 µm-long MT, and we estimate the
average power drain as:
?P?=(1/2)V20[R/(R2+X2c)],
(14)
where Xc= 1/?C is the capacitive resistance. Substituting the relevant numbers we obtain the power dissipated to be in the 10?11 W
range which is comparable to the power generated by the cell in
metabolic processes (100 W of power generation in the body/3 × 1013 cells
in the body). Consequently, additional heat generated by these
processes may be disruptive to living cells although there is no
experimentally detected thermal effect of TTFields.
5.4. Resonance Effects on MTs
Cosic et al. [103,104]
reported EM resonances in biological molecules (proteins, DNA and RNA)
in THz, GHz, MHz and kHz ranges. They proposed the so-called resonant
recognition model (RRM) based on the distribution of energy of
delocalized proteins in a biological system and charge transfer under
resonance with a velocity of 7.87 × 105m/s and covering distances of 3.8 Å between amino acids, giving a characteristic frequency between 1013 and 1015 Hz.
Then they state a variety of charge transfer velocities yielding
different resonant frequencies. Of particular interest to this review is
the velocity v = 0.0005 m/s which produces EMF in the range of
108–325 kHz for TERT, TERT mRNA and Telomere. This velocity corresponds
the propagation of solitons on ?-helices. For tubulin and MTs, three
specific ranges of resonant frequencies have been predicted by the RRM
approach: 97–101 THz, 340–350 THz and 445–470 THz, none of which
overlaps with TTField frequencies.
H-bond strength in MTs has been recently computationally estimated [105]
as ranging from 11.9 k/mol for the weakest bond to 42.2 kJ/mol for the
strongest one and a total of 462 kJ/mol for the ?-tubulin/?-tubulin
interactions and 472 kJ/mol for the ?-tubulin/?-tubulin interactions,
which based on the Planck relationship between frequency and energy
translates into a range of frequency values between 0.3 × 1014 Hz and 1.3 × 1015Hz.
Again, these frequencies are much too high to be affected by TTFields.
Therefore, we do not expect TTFields to be capable of disrupting the MT
structure.
Furthermore, Pizzi et al. [106]
measured microwave resonance effects in MTs and found a resonant
frequency at 1.510 GHz. This may not correspond to bond-breaking between
tubulin dimers but simply to some specific electro-mechanical
oscillations. Finally, Preto et al. [92]
re-evaluated the Froehlich mechanism for long-range interactions and
concluded that classical electromagnetic dipole-dipole interactions at
high enough frequencies can lead to attraction between oscillating
dipoles over distances comparable to the size of the cell. However, even
including a coherently coupled layer of water molecules around a
protein, this would require frequencies in the THz range or higher.
Consequently, almost all of the resonant frequencies listed above fall
well outside the range of potential overlap with the 100 kHz frequencies
of TTFields.
5.5. Ionic Wave Conductivity along Actin Filaments and AC Fields
AFs are approximately 7 nm in diameter,
with a periodic helical structure repeating every 37 nm. Actin filaments
are arranged from actin monomers resulting in an alternating
distribution of electric dipole moments along the length of each
filament [107]. They are characterized by a high electrostatic charge density [108,109] resulting in ionic current conductivity involving the counter-ions surrounding them [109], which is very similar to the effects observed for MTs [60]. The observed wave patterns in electrically-stimulated AFs [30] were very similar to the solitary waveforms recorded for electrically-stimulated non-linear transmission lines [110]. In these experiments [30,42],
an input voltage pulse was applied with an amplitude of 200 mV for a
duration of 800 ms. Electrical signals were measured at the opposite end
of the AF demonstrating that AFs support axial non-linear ionic
currents. Since AFs produce a spatially-dependent electric field
arranged in peaks and troughs [111]
with an average pitch ~35–40 nm, they can be modeled as an electrical
circuit with the following non-linear components: (a) a non-linear
capacitor associated with the spatial charge distribution between the
ions located in the outer and inner areas of the polymer; (b) an
inductor; and (c) a resistor, similar to the model described above
developed for MTs. A helical distribution of ions winding around the
filament at an approximate distance of one Bjerrum length to the
filament corresponds to a solenoid in which an ionic current flows due
to the voltage gradient between the two ends. For an AF with n monomers, its effective resistance, inductance, and capacitance are, respectively:
Reff=(?ni=11R2,1)?1+?ni=1R1,i,
(15)
Leff=?ni=1Li,
(16)
Ceff=?ni=1C0,i,
(17)
where R1,i = 6.11 × 106 ?, and R2,i = 0.9 × 106 ?, such that R1,i = 7R2,i [112]. Hence, for a 1-µm length of an AF we find that Reff = 1.2 × 109 ?, Leff = 340 × 10?12 H and Ceff = 0.02 × 10?12 F.
The electrical model of an AF is an application of Kirchhoff’s laws to
one section of the effective electrical circuit that is coupled to
neighboring monomers. In the continuum limit [112] the following equation describes the spatio-temporal behavior of the electric potential propagating along the actin filament:
LC0?2V?t2=a2(?xxV)+ R2C0??t(a2(?xxV))? R1C0?V?t+R1C02bV?V?t.
(18)
Solitary ionic waves have been described as the solutions of the above nonlinear partial differential equation [112] with an estimated velocity of propagation between 1 and 100 m/s [72]. This model has been recently updated with a more plausible estimation of model parameters [100]. Like MTs [96], AFs can be manipulated by external electric fields [113].
In a similar manner to our analysis of the time scales for MTs as ionic
conduction cables with RLC components, we estimate similar time scales
for actin and AFs. We readily find for a single actin monomer, that the
time scale for LC oscillations is very fast, namely ?0 = (LC)1/2 and ?0 = 6 × 10?14 s. Secondly, the decay time for longitudinal ionic waves is ?1 = R1C = 6 × 10?10 s while the corresponding time for radial waves is ?2 = R2C = 0.9 × 10?10 s.
All of the above time scales are not compatible with interactions
involving electric fields in the 100 kHz range. However, the situation
changes drastically for AFs where there is a similar scaling with the
length of the filament as described above for MTs. Taking as an example a
1-µm AF, we find ?0 = 10?11 s, which is still too short but ?1 = R1C = 2.4 × 10?5 s
which is in the correct range of time for interactions with AC electric
fields in the 100 kHz range. It should be noted that AFs have been
found sensitive to AC fields under experimental conditions [114].
5.6. Electric Field Effects on DNA
Anderson and Record [115]
described ionic distribution around DNA in great detail. During
interphase, DNA contents present in the nucleus are expected to be
protected from external fields due to being enclosed in the nearly
spherical nuclear membrane [78].
In addition to the screening effects of being shielded both by the cell
membrane and the nuclear wall, the irregular geometry of the DNA
strands and their short persistence length indicate that while highly
charged, DNA is unlikely to participate in ionic conduction effects
shown either for AFs or MTs, both of which have very large persistence
lengths.
However, at the beginning of mitosis, the
nuclear membrane breaks down, thus potentially not shielding the DNA
any longer which would allow for the action of electric fields on
chromosomes.
5.7. Electric Field Effects on Motor Proteins
Kinesin participates in mass transport along MTs and propagates at a maximum speed of 10?6 m/s.
This value depends on the concentration of ATP and the ionic
concentrations in the medium. In the case of MTs, kinesin transports
various crucial cargo and for actin filaments, dynein does the same at
similar speeds. Hence each step of a motor protein takes place over the
period of a few ms, which is much longer than the period of AC field
oscillations. However, kinesin binds to MTs through C-termini, which are
very sensitive to electric field fluctuations and hence it is possible
that kinesin transport would be very strongly disrupted by these rapid
oscillations of C-termini. This aspect merits careful experimental
verification.
Another potential member of the cytoskeleton that has been found affected by TTFields [2]
is the protein called septin, which are GTP-binding like tubulin but
form oligomeric hetero-complexes including rings and filaments. There is
no information at the present time that could shed light on the
mechanism of TTField effects with septin-based structures.
6. Discussion
The cytoskeleton and especially, MTs,
may participate in numerous interactions with electromagnetic forces due
to the complex charge distribution in and around these protein
filaments surrounded by poly-ionic solutions. First of all, there are
large net charges on tubulin, which are largely but not completely
screened by counter-ions. Secondly, some of the charges are localized on
C-termini, which are very flexible leading to oscillating charge
configurations. Then, there are ions surrounding the protein that can be
partially condensed and susceptible to collective oscillations.
Moreover, there are large dipole moments on tubulin and microtubules
whose geometric organization importantly affects their response to
external fields. Finally, there can be induced dipole moments especially
in the presence of electric field gradients. Disentangling the relative
importance of the various effects under different conditions is not
trivial and requires careful examination.
Depending on the orientation of the electric fields with
the cell axis and in particular with the MT axis (however, they fan out
from centrosomes in mitotic cells, so there will be at different angles
to any field), there could in general be three types of ionic waves
generated:
Longitudinal waves propagating along the MT surface. In this case
each protofilament of a microtubule acts like a cable with its inherent
resistance r, so the resistance of an entire microtubule would be R = r/13 since all these cables are in parallel to each other.
Helical waves propagating around and along each microtubule, there
could be three or five such waves propagating simultaneously mimicking
the three-start or five-start geometry of a microtubule. The effective
resistance of such cables would be the individual resistance divided by
the number of cables in parallel.
Radial waves propagating perpendicularly to the microtubule surface.
If a field is oriented at an angle to the MT axis, it is
expected that all these wave types may be generated simultaneously. Once
AC fields generate oscillating ionic flows, these can in turn:
Interfere with ion flows in the cleavage area of dividing cells.
Interfere with motor protein motion and MAP-MT interactions.
May to a lesser degree affect ion channel dynamics.
May in general affect the net charge of the cytoplasm.
Finally, Kirson et al. [2]
mention intracellular charged and polar entities such as cytoplasmic
organelles as being potentially most directly affected by TTFields. This
is not specifically addressed in this paper due to size and scope
limitations as well as the scarcity of data in this regard. It has been
argued [2]
that inhomogeneity in field intensity may exert a uni-directional
electric force on all intracellular charged and polar entities, pulling
them toward the furrow (regardless of field polarity). It was determined
that cytoplasmic organelles are electrically polarized by the field
within dividing cells. As a consequence, the TTField-generated forces
acting on these organelles may reach values up to 60 pN resulting in
their movement toward the cleavage furrow. These organelles can move at
velocities up to 30 ?m/s and, as a result, they could pile up at the
cleavage furrow within a few minutes, interfering with cytokinesis,
which may lead to cell destruction. This aspect needs detailed
experimental investigation.
Some measurable heating effects in the
cytoplasm might also be expected. These fields are not expected to
affect permanent dipoles of proteins such as tubulin and actin. Although
TTField effects have not been specifically assessed for AFs, an earlier
paper [114]
investigated exposure of cells to AC electric fields in a low frequency
range of 1–120 Hz and found significant induced alterations in the AF
structure, which were both frequency- and amplitude dependent. An
application of 1–10 Hz AC fields caused AF reorganization from
continuous, aligned cable structures to discontinuous globular patches.
Cells exposed to 20–120 Hz electric fields were not visibly affected.
The extent of AF reorganization increased nonlinearly with the electric
field strength. The characteristic time for AF reorganization in cells
exposed to a 1 Hz, 20 V/cm electric field was approximately 5 min.
Importantly, applied AC electric fields were shown to initiate signal
transduction cascades, which in turn cause reorganization of
cytoskeletal structures. Therefore, in addition to direct effects of
TTFields, there may be indirect, down-stream interactions.
7. Conclusions
Based on the extensive analysis of the
various possible effects AC electric fields can have on living cells, we
conclude the following. Electric field gradients, especially in
dividing cells, cause substantial DEP forces on tubulin dimers and MTs.
The longer the MT, the more pronounced the effect. Additionally, another
likely scenario is that ionic current flows along and perpendicular to
MT surfaces (as well as actin filaments, but less likely) take place,
which can be generated by AC field oscillations in the 100–300 kHz
range. The specific frequency selection depends critically on the length
of each filament.
Identification of the strength, cause,
and function of intracellular electric fields has only recently been
experimentally accessible, although speculations in this area have
existed for over a decade. These insights may also assist in devising
and optimizing ways and means of affecting cells, especially cancer
cells, by the application of external electric fields. With the advent
of nanoprobe technology, which has shown promise in measuring these
fields at a subcellular level, it is very timely to explore the various
physical properties of the cytoplasmic environment including the
cytoskeleton and the ionic contents of the cytoplasm. This research
promises to contribute to our understanding of the cytoplasm in live
cells and the role of microtubules and mitochondria in creating dynamic
and structural order in healthy functioning cells. It will also be of
help to identify biophysical differences in cancer cells that lead to
increased metastatic behavior. Such an understanding may lead to
optimized therapies and the identification of specific targets to halt
metastatic transformation, as well as insights into the mechanism of
action of current electromagnetic cancer therapies that are FDA approved
and are in development.
Acknowledgments
Cornelia Wenger was supported by Novocure. Douglas E.
Friesen was supported by Novocure. Douglas E. Friesen also gratefully
acknowledges support from Alberta Innovates Health Solutions and the
Alberta Cancer Foundation. The funding for J.A.T.’s research comes from
the Natural Sciences and Engineering Research Council of Canada.
Abbreviations
The following abbreviations are used in this manuscript:
DC
direct current
AC
alternating current
TTFields
Tumor Treating Fields
GBM
glioblastoma multiforme
EM
electromagnetic
MT
microtubule
DEP
dielectrophoretic
AF
actin filament
TT
C-terminal tail
MAP
microtubule associated protein
Author Contributions
Jack A. Tuszynski produced the first draft of the
manuscript. Cornelia Wenger performed the computational studies and
contributed to editing the paper. Douglas E. Friesen helped conceive the
ideas presented in the paper and contributed to editing the paper.
Jordane Preto contributed the analysis of ion motion in electric fields.
Conflicts of Interest
Novocure had no role in the design of the study; in the
collection, analyses, or interpretation of data; in the writing of the
manuscript, and in the decision to publish the results.
References
1. Cifra M., Fields J.Z., Farhadi A. Electromagnetic cellular interactions. Prog. Biophys. Mol. Biol. 2011;105:223–246. doi: 10.1016/j.pbiomolbio.2010.07.003. [PubMed][Cross Ref]
2. Kirson
E.D., Gurvich Z., Schneiderman R., Dekel E., Itzhaki A., Wasserman Y.,
Schatzberger R., Palti Y. Disruption of cancer cell replication by
alternating electric fields. Cancer Res. 2004;64:3288–3295. doi: 10.1158/0008-5472.CAN-04-0083.[PubMed] [Cross Ref]
3. Kirson
E.D., Dbalý V., Tovarys F., Vymazal J., Soustiel J.F., Itzhaki A.,
Mordechovich D., Steinberg-Shapira S., Gurvich Z., Schneiderman R., et
al. Alternating electric fields arrest cell proliferation in animal
tumor models and human brain tumors. Proc. Natl. Acad. Sci. USA. 2007;104:10152–10157. doi: 10.1073/pnas.0702916104.[PMC free article] [PubMed] [Cross Ref]
4. Davies A.M., Weinberg U., Palti Y. Tumor treating fields: A new frontier in cancer therapy. Ann. N. Y. Acad. Sci. 2013;1291:86–95. doi: 10.1111/nyas.12112. [PubMed][Cross Ref]
5. Stupp
R., Wong E.T., Kanner A.A., Steinberg D., Engelhard H., Heidecke V.,
Kirson E.D., Taillibert S., Liebermann F., Dbalý V., et al. NovoTTF-100A
versus physician’s choice chemotherapy in recurrent glioblastoma: A
randomised phase III trial of a novel treatment modality. Eur. J. Cancer. 2012;48:2192–2202. doi: 10.1016/j.ejca.2012.04.011.[PubMed] [Cross Ref]
6. Kirson
E.D., Giladi M., Gurvich Z., Itzhaki A., Mordechovich D., Schneiderman
R.S., Wasserman Y., Ryffel B., Goldsher D., Palti Y. Alternating
electric fields (TTFields) inhibit metastatic spread of solid tumors to
the lungs. Clin. Exp. Metastasis. 2009;26:633–640. doi: 10.1007/s10585-009-9262-y. [PMC free article] [PubMed][Cross Ref]
7. Stupp
R., Taillibert S., Kanner A.A., Kesari S., Steinberg D.M., Toms S.A.,
Taylor L.P., Lieberman F., Silvani A., Fink K.L., et al. Maintenance
therapy with tumor-treating fields plus temozolomide vs. temozolomide
alone for glioblastoma: A randomized clinical trial. JAMA. 2015;314:2535–2543. doi: 10.1001/jama.2015.16669. [PubMed][Cross Ref]
8. Kirson
E.D., Schneiderman R.S., Dbalý V., Tovaryš F., Vymazal J., Itzhaki A.,
Mordechovich D., Gurvich Z., Shmueli E., Goldsher D., et al.
Chemotherapeutic treatment efficacy and sensitivity are increased by
adjuvant alternating electric fields (TTFields) BMC Med. Phys. 2009;9:1–13. doi: 10.1186/1756-6649-9-1.[PMC free article] [PubMed] [Cross Ref]
9. Berg
H., Günther B., Hilger I., Radeva M., Traitcheva N., Wollweber L.
Bioelectromagnetic field effects on cancer cells and mice tumors. Electromagn. Biol. Med. 2010;29:132–143. doi: 10.3109/15368371003776725. [PubMed] [Cross Ref]
10. Funk R.H.W., Monsees T., Ozkucur N. Electromagnetic effects—From cell biology to medicine. Prog. Histochem. Cytochem. 2009;43:177–264. doi: 10.1016/j.proghi.2008.07.001. [PubMed] [Cross Ref]
11. Dyshlovoi
V.D., Panchuk A.S., Kachura V.S. Effect of electromagnetic field of
industrial frequency on the growth pattern and mitotic activity of
cultured human fibroblastoid cells. Cytol. Genet. 1981;15:9–12. [PubMed]
12. Robertson D., Miller M.W., Cox C., Davis H.T. Inhibition and recovery of growth processes in roots of Pisum sativum L. exposed to 60-Hz electric fields. Bioelectromagnetics. 1981;2:329–340. doi: 10.1002/bem.2250020405. [PubMed][Cross Ref]
13. Jaffe L.F., Nuccitelli R. Electrical controls of development. Annu. Rev. Biophys. Bioeng. 1977;6:446–476. doi: 10.1146/annurev.bb.06.060177.002305. [PubMed][Cross Ref]
14. Tuszy?ski
J.A., Hameroff S., Satari? M.V., Trpisová B., Nip M.L.A. Ferroelectric
behavior in microtubule dipole lattices: Implications for information
processing, signaling and assembly/disassembly. J. Theor. Biol. 1995;174:371–380. doi: 10.1006/jtbi.1995.0105. [Cross Ref]
15. Gagliardi L.J. Electrostatic force in prometaphase, metaphase, and anaphase-A chromosome motions. Phys. Rev. E Stat. Nonlinear Soft Matter Phys. 2002;66:011901. doi: 10.1103/PhysRevE.66.011901. [PubMed] [Cross Ref]
16. Gagliardi L.J. Microscale electrostatics in mitosis. J. Electrostat. 2002;54:219–232. doi: 10.1016/S0304-3886(01)00155-3. [Cross Ref]
17. Pohl H.A. Dielectrophoresis. Cambridge University Press; Cambridge, UK: 1978.
18. Cooper M. Coherent polarization waves in cell division and cancer. Collect. Phenom. 1981;3:273–288.
19. Pohl
H.A., Braden T., Robinson S., Piclardi J., Pohl D.G. Life cycle
alterations of the micro-dielectrophoretic effects of cell. J. Biol. Phys. 1981;9:133–154. doi: 10.1007/BF01988247. [Cross Ref]
20. Pohl H.A. Oscillating fields about growing cells. Int. J. Quantum Chem. 1980;18:411–431. doi: 10.1002/qua.560180740. [Cross Ref]
21. Jelínek
F., Pokorný J., Saroch J., Trkal V., Hasek J., Palán B. Microelectronic
sensors for measurement of electromagnetic fields of living cells and
experimental results. Bioelectrochem. Bioenerg. 1999;48:261–266. doi: 10.1016/S0302-4598(99)00017-3.[PubMed] [Cross Ref]
22. Gagliardi L.J. Electrostatic force generation in chromosome motions during mitosis. J. Electrostat. 2005;63:309–327. doi: 10.1016/j.elstat.2004.09.007. [Cross Ref]
23. Tuszynski J.A., Dixon J.M. Biomedical applications of introductory physics. Eur. J. Phys. 2002;23:591. doi: 10.1088/0143-0807/23/5/601. [Cross Ref]
24. Howard J. Mechanics of Motor Proteins and the Cytoskeleton. Sinauer Associates; Sunderland, MA, USA: 2001.
25. Grosse C., Schwan H.P. Cellular membrane potentials induced by alternating fields. Biophys. J. 1992;63:1632–1642. doi: 10.1016/S0006-3495(92)81740-X.[PMC free article] [PubMed] [Cross Ref]
26. Kotnik T., Bobanovi? F., Miklav?i? D. Sensitivity of transmembrane voltage induced by applied fields—A theoretical analysis. Bioelectrochem. Bioenergy. 1997;43:285–291. doi: 10.1016/S0302-4598(97)00023-8. [Cross Ref]
27. Bernhardt
J., Pauly H. On the generation of potential differences across the
membranes of ellipsoidal cells in an alternating electrical field. Biophysik. 1973;10:89–98. doi: 10.1007/BF01189915. [PubMed] [Cross Ref]
28. Gimsa
J., Wachner D. A polarization model overcoming the geometric
restrictions of the laplace solution for spheroidal cells: Obtaining new
equations for field-induced forces and transmembrane potential. Biophys. J. 1999;77:1316–1326. doi: 10.1016/S0006-3495(99)76981-X. [PMC free article] [PubMed] [Cross Ref]
29. Gimsa
J., Wachner D. Analytical description of the transmembrane voltage
induced on arbitrarily oriented ellipsoidal and cylindrical cells. Biophys. J. 2001;81:1888–1896. doi: 10.1016/S0006-3495(01)75840-7. [PMC free article] [PubMed] [Cross Ref]
30. Kotnik T., Miklav?i? D. Second-order model of membrane electric field induced by alternating external electric fields. IEEE Trans. Biomed. Eng. 2000;47:1074–1081. doi: 10.1109/10.855935. [PubMed] [Cross Ref]
31. Kotnik
T., Miklav?i? D. Theoretical evaluation of voltage inducement on
internal membranes of biological cells exposed to electric fields. Biophys. J. 2006;90:480–491. doi: 10.1529/biophysj.105.070771. [PMC free article] [PubMed] [Cross Ref]
32. Gowrishankar T.R., Weaver J.C. An approach to electrical modeling of single and multiple cells. Proc. Natl. Acad. Sci. USA. 2003;100:3203–3208. doi: 10.1073/pnas.0636434100. [PMC free article] [PubMed] [Cross Ref]
33. Stewart
D.A., Gowrishankar T.R., Smith K.C., Weaver J.C. Cylindrical cell
membranes in uniform applied electric fields: Validation of a transport
lattice method. IEEE Trans. Biomed. Eng. 2005;52:1643–1653. doi: 10.1109/TBME.2005.856030.[PubMed] [Cross Ref]
34. Pavlin
M., Miklav?i? D. The effective conductivity and the induced
transmembrane potential in dense cell system exposed to DC and AC
electric fields. IEEE Trans. Plasma Sci. 2009;37:99–106. doi: 10.1109/TPS.2008.2005292. [Cross Ref]
35. Hobbie R.K., Roth B.J. Intermediate Physics for Medicine and Biology. 4th ed. Springer; New York, NY, USA: 2007.
36. King R.W.P., Wu T.T. Electric field induced in cells in the human body when this is exposed to low-frequency electric fields. Phys. Rev. E Stat. Nonlinear Soft Matter Phys. 1998;58:2363–2369. doi: 10.1103/PhysRevE.58.2363. [Cross Ref]
37. Wenger
C., Giladi M., Bomzon Z., Salvador R., Basser P.J., Miranda P.C.
Modeling Tumor Treating Fields (TTFields) application in single cells
during metaphase and telophase; In Proceedings of the 2015 37th Annual
International Conference of the IEEE Engineering in Medicine and Biology
Society (EMBC); Milan, Italy. 25–29 August 2015; pp. 6892–6895. [PubMed]
38. Boucrot E., Kirchhausen T. Mammalian cells change volume during mitosis. PLoS ONE. 2008;3:e1477 doi: 10.1371/journal.pone.0001477. [PMC free article] [PubMed][Cross Ref]
39. Habela C.W., Sontheimer H. Cytoplasmic volume condensation is an integral part of mitosis. Cell Cycle. 2007;6:1613–1620. doi: 10.4161/cc.6.13.4357. [PMC free article][PubMed] [Cross Ref]
40. Vajrala
V., Claycomb J.R., Sanabria H., Miller J.H. Effects of oscillatory
electric fields on internal membranes: An analytical model. Biophys. J. 2008;94:2043–2052. doi: 10.1529/biophysj.107.114611. [PMC free article] [PubMed] [Cross Ref]
41. Giladi
M., Porat Y., Blatt A., Wasserman Y., Kirson E.D., Dekel E., Palti Y.
Microbial growth inhibition by alternating electric fields. Antimicrob. Agents Chemother. 2008;52:3517–3522. doi: 10.1128/AAC.00673-08. [PMC free article][PubMed] [Cross Ref]
42. Sun
T., Morgan H., Green N. Analytical solutions of AC electrokinetics in
interdigitated electrode arrays: Electric field, dielectrophoretic and
traveling-wave dielectrophoretic forces. Phys. Rev. E Stat. Nonlinear Soft Matter Phys. 2007;76:046610. doi: 10.1103/PhysRevE.76.046610. [PubMed] [Cross Ref]
43. Jones T.B. Basic theory of dielectrophoresis and electrorotation. IEEE Eng. Med. Biol. Mag. 2003;22:33–42. doi: 10.1109/MEMB.2003.1304999. [PubMed] [Cross Ref]
44. Markx G.H. The use of electric fields in tissue engineering: A review. Organogenesis. 2008;4:11–17. doi: 10.4161/org.5799. [PMC free article] [PubMed][Cross Ref]
45. Giladi
M., Schneiderman R.S., Porat Y., Munster M., Itzhaki A., Mordechovich
D., Cahal S., Kirson E.D., Weinberg U., Palti Y. Mitotic disruption and
reduced clonogenicity of pancreatic cancer cells in vitro and in vivo by
tumor treating fields. Pancreatology. 2014;14:54–63. doi: 10.1016/j.pan.2013.11.009. [PubMed] [Cross Ref]
46. Tyner K.M., Kopelman R., Philbert M.A. “Nanosized voltmeter” enables cellular-wide electric field mapping. Biophys. J. 2007;93:1163–1174. doi: 10.1529/biophysj.106.092452. [PMC free article] [PubMed] [Cross Ref]
47. Qvist J., Persson E., Mattea C., Halle B. Time scales of water dynamics at biological interfaces: Peptides, proteins and cells. Faraday Discuss. 2009;141:131–144. doi: 10.1039/B806194G. [PubMed] [Cross Ref]
48. Tuszynski J.A. Molecular and Cellular Biophysics. Chapman & Hall/CRC; Boca Raton, FL, USA: 2008.
49. Szent-Györgyi A. The study of energy-levels in biochemistry. Nature. 1941;148:157–159. doi: 10.1038/148157a0. [Cross Ref]
50. Szent-Györgyi A. Bioenergetics. Academic Press; New York, NY, USA: 1957.
51. Gascoyne P.R.C., Pethig R., Szent-Györgyi A. Water structure-dependent charge transport in proteins. Proc. Natl. Acad. Sci. USA. 1981;78:261–265. doi: 10.1073/pnas.78.1.261. [PMC free article] [PubMed] [Cross Ref]
52. Szent-Györgyi A. Biolectronics and cancer. J. Bioenerg. 1973;4:533–562. doi: 10.1007/BF01516207. [PubMed] [Cross Ref]
53. Szent-Györgyi A. Electronic biology and its relation to cancer. Life Sci. 1974;15:863–875. doi: 10.1016/0024-3205(74)90003-4. [PubMed] [Cross Ref]
54. Craddock
T.J., Tuszy?ski J.A., Priel A., Freedman H. Microtubule ionic
conduction and its implications for higher cognitive functions. J. Integr. Neurosci. 2010;9:103–122. doi: 10.1142/S0219635210002421. [PubMed] [Cross Ref]
55. Sahu
S., Ghosh S., Ghosh B., Aswani K., Hirata K., Fujita D., Bandyopadhyay
A. Atomic water channel controlling remarkable properties of a single
brain microtubule: Correlating single protein to its supramolecular
assembly. Biosens. Bioelectron. 2013;47:141–148. doi: 10.1016/j.bios.2013.02.050. [PubMed] [Cross Ref]
56. Levin M. Bioelectromagnetics in morphogenesis. Bioelectromagnetics. 2003;24:295–315. doi: 10.1002/bem.10104. [PubMed] [Cross Ref]
57. McCaig C.D., Rajnicek A.M., Song B., Zhao M. Controlling cell behavior electrically: Current views and future potential. Physiol. Rev. 2005;85:943–978. doi: 10.1152/physrev.00020.2004. [PubMed] [Cross Ref]
58. Scholkmann F., Fels D., Cifra M. Non-chemical and non-contact cell-to-cell communication: A short review. Am. J. Transl. Res. 2013;5:586–593. [PMC free article][PubMed]
59. Zheng
J.M., Chin W.C., Khijniak E., Khijniak E.J., Pollack G.H. Surfaces and
interfacial water: Evidence that hydrophilic surfaces have long-range
impact. Adv. Colloid Interface Sci. 2006;127:19–27. doi: 10.1016/j.cis.2006.07.002. [PubMed][Cross Ref]
60. Priel A., Ramos A.J., Tuszynski J.A., Cantiello H.F. A biopolymer transistor: Electrical amplification by microtubules. Biophys. J. 2006;90:4639–4643. doi: 10.1529/biophysj.105.078915. [PMC free article] [PubMed] [Cross Ref]
61. Sekuli? D.L., Satari? B.M., Tuszy?ski J.A., Satari? M.V. Nonlinear ionic pulses along microtubules. Eur. Phys. J. E Soft Matter. 2011;34:49. doi: 10.1140/epje/i2011-11049-0. [PubMed] [Cross Ref]
62. Chou
K.C., Zhang C.T., Maggiora G.M. Solitary wave dynamics as a mechanism
for explaining the internal motion during microtubule growth. Biopolymers. 1994;34:143–153. doi: 10.1002/bip.360340114. [PubMed] [Cross Ref]
63. Ku?era O., Havelka D. Mechano-electrical vibrations of microtubules—Link to subcellular morphology. Biosystems. 2012;109:346–355. doi: 10.1016/j.biosystems.2012.04.009. [PubMed] [Cross Ref]
64. Havelka
D., Ku?era O., Deriu M.A., Cifra M. Electro-acoustic behavior of the
mitotic spindle: A semi-classical coarse-grained model. PLoS ONE. 2014;9:e86501 doi: 10.1371/journal.pone.0086501. [PMC free article] [PubMed] [Cross Ref]
65. Preto J., Pettini M., Tuszy?ski J.A. Possible role of electrodynamic interactions in long-distance biomolecular recognition. Phys. Rev. E Stat. Nonlinear Soft Matter Phys. 2015;91:052710. doi: 10.1103/PhysRevE.91.052710. [PubMed] [Cross Ref]
66. Cifra
M., Havelka D., Deriu M.A. Electric field generated by longitudinal
axial microtubule vibration modes with high spatial resolution
microtubule model. J. Phys. Conf. Ser. 2011;329:012013. doi: 10.1088/1742-6596/329/1/012013. [Cross Ref]
67. Cifra M., Pokorný J., Havelka D., Ku?era O. Electric field generated by axial longitudinal vibration modes of microtubule. Biosystems. 2010;100:122–131. doi: 10.1016/j.biosystems.2010.02.007. [PubMed] [Cross Ref]
68. Tuszy?ski
J.A., Brown J.A., Crawford E., Carpenter E.J., Nip M.L., Dixon J.M.,
Satari? M.V. Molecular dynamics simulations of tubulin structure and
calculations of electrostatic properties of microtubules. Math. Comput. Model. 2005;41:1055–1070. doi: 10.1016/j.mcm.2005.05.002. [Cross Ref]
69. Carpenter
E.J., Huzil J.T., Ludueña R.F., Tuszy?ski J.A. Homology modeling of
tubulin: Influence predictions for microtubule’s biophysical
properties. Eur. Biophys. J. 2006;36:35–43. doi: 10.1007/s00249-006-0088-0. [PubMed] [Cross Ref]
70. Tuszynski
J.A., Carpenter E.J., Huzil J.T., Malinski W., Luchko T., Luduena R.F.
The evolution of the structure of tubulin and its potential consequences
for the role and function of microtubules in cells and embryos. Int. J. Dev. Biol. 2006;50:341–358. doi: 10.1387/ijdb.052063jt. [PubMed] [Cross Ref]
71. Vassilev
P.M., Dronzine T., Vassileva M.P., Georgiev G.A. Parallel arrays of
microtubules formed in electric and magnetic fields. Biosci. Rep. 1982;2:1025–1029. doi: 10.1007/BF01122171. [PubMed] [Cross Ref]
72. Brown
J.A., Dixon J.M., Cantiello H.F., Priel A., Tuszy?ski J.A. Electronic
and ionic conductivities of microtubules and actin filaments, their
consequences for cell signaling and applications to bioelectronics. In:
Lyshevski S.E., editor. Nano and Molecular Electronics Handbook. CRC Press; Boca Raton, FL, USA: 2007.
73. Priel A., Tuszy?ski J. A nonlinear cable-like model of amplified ionic wave propagation along microtubules. Eur. Lett. 2008;83:68004. doi: 10.1209/0295-5075/83/68004. [Cross Ref]
74. Friesen
D.E., Craddock T.J.A., Kalra A.P., Tuszynski J.A. Biological wires,
communication systems, and implications for disease. Biosystems. 2015;127:14–27. doi: 10.1016/j.biosystems.2014.10.006. [PubMed] [Cross Ref]
75. Stracke
R., Böhm K.J., Wollweber L., Tuszynski J.A., Unger E. Analysis of the
migration behaviour of single microtubules in electric fields. Biochem. Biophys. Res. Commun. 2002;293:602–609. doi: 10.1016/S0006-291X(02)00251-6. [PubMed][Cross Ref]
76. Sahu
S., Ghosh S., Hirata K., Fujita D., Bandyopadhyay A. Multi-level
memory-switching properties of a single brain microtubule. Appl. Phys. Lett. 2013;102:123701. doi: 10.1063/1.4793995. [Cross Ref]
77. Minoura I., Muto E. Dielectric measurement of individual microtubules using the electroorientation method. Biophys. J. 2006;90:3739–3748. doi: 10.1529/biophysj.105.071324. [PMC free article] [PubMed] [Cross Ref]
78. Brown J.A., Tuszynski J.A. A review of the ferroelectric model of microtubules. Ferroelectrics. 1999;220:141–155. doi: 10.1080/00150199908216213. [Cross Ref]
79. Uppalapati M., Huang Y.-M., Jackson T.N., Hancock W.O. Microtubule alignment and manipulation using AC electrokinetics. Small. 2008;4:1371–1381. doi: 10.1002/smll.200701088. [PubMed] [Cross Ref]
80. Liu D.S., Astumian R.D., Tsong T.Y. Activation of Na+ and K+ pumping modes of (Na, K)-ATPase by an oscillating electric field. J. Biol. Chem. 1990;265:7260–7267.[PubMed]
81. Tsong T.Y., Astumian R.D. 863—Absorption and conversion of electric field energy by membrane bound ATPases. Bioelectrochem. Bioenergy. 1986;15:457–476. doi: 10.1016/0302-4598(86)85034-6. [Cross Ref]
82. Tsong
T.Y. Electrical modulation of membrane proteins: Enforced
conformational oscillations and biological energy and signal
transductions. Annu. Rev. Biophys. Biophys. Chem. 1990;19:83–106. doi: 10.1146/annurev.bb.19.060190.000503.[PubMed] [Cross Ref]
83. White J.A., Rubinstein J.T., Kay A.R. Channel noise in neurons. Trends Neurosci. 2000;23:131–137. doi: 10.1016/S0166-2236(99)01521-0. [PubMed] [Cross Ref]
84. Roux B., Schulten K. Computational studies of membrane channels. Structure. 2004;12:1343–1351. doi: 10.1016/j.str.2004.06.013. [PubMed] [Cross Ref]
85. Bernèche S., Roux B. Energetics of ion conduction through the K+ channel. Nature. 2001;414:73–77. doi: 10.1038/35102067. [PubMed] [Cross Ref]
86. Kuyucak S., Andersen O.S., Chung S.-H. Models of permeation in ion channels. Rep. Prog. Phys. 2001;64:1427–1472. doi: 10.1088/0034-4885/64/11/202. [Cross Ref]
87. Guidoni L., Carloni P. Potassium permeation through the KcsA channel: A density functional study. Biochim. Biophys. Acta. 2002;1563:1–6. doi: 10.1016/S0005-2736(02)00349-8. [PubMed] [Cross Ref]
88. Shrivastava I.H., Tieleman D.P., Biggin P.C., Sansom M.S.P. K+ versus Na+ ions in a K channels selectivity filter: A simulation study. Biophys. J. 2002;83:633–645. doi: 10.1016/S0006-3495(02)75197-7. [PMC free article] [PubMed] [Cross Ref]
89. Bernèche S., Roux B. A gate in the selectivity filter of potassium channels. Structure. 2005;13:591–600. doi: 10.1016/j.str.2004.12.019. [PubMed] [Cross Ref]
90. Mastro A.M., Babich M.A., Taylor W.D., Keith A.D. Diffusion of a small molecule in the cytoplasm of mammalian cells. Proc. Natl. Acad. Sci. USA. 1984;81:3414–3418. doi: 10.1073/pnas.81.11.3414. [PMC free article] [PubMed] [Cross Ref]
91. Gardiner C.W. Handbook of Stochastic Methods. Springer; Berlin, Germany: 1985.
92. Preto
J., Floriani E., Nardecchia I., Ferrier P., Pettini M. Experimental
assessment of the contribution of electrodynamic interactions to
long-distance recruitment of biomolecular partners: Theoretical basis. Phys. Rev. E Stat. Nonlinear Soft Matter Phys. 2012;85:041904. doi: 10.1103/PhysRevE.85.041904. [PubMed] [Cross Ref]
93. Brics M., Kaupuzs J., Mahnke R. How to solve Fokker-Planck equation treating mixed eigenvalue spectrum? Condens. Matter Phys. 2013;16:1–13. doi: 10.5488/CMP.16.13002. [Cross Ref]
94. Alexandrov B.S., Gelev V., Bishop A.R., Usheva A., Rasmussen K. DNA breathing dynamics in the presence of a terahertz field. Phys. Lett. A. 2010;374:1214–1217. doi: 10.1016/j.physleta.2009.12.077. [PMC free article] [PubMed] [Cross Ref]
95. Isozaki
N., Ando S., Nakahara T., Shintaku H., Kotera H., Meyhöfer E., Yokokawa
R. Control of microtubule trajectory within an electric field by
altering surface charge density. Sci. Rep. 2015;5:7669. doi: 10.1038/srep07669. [PMC free article] [PubMed][Cross Ref]
96. Van den Heuvel M.G.L., de Graaff M.P., Lemay S.G., Dekker C. Electrophoresis of individual microtubules in microchannels. Proc. Natl. Acad. Sci. USA. 2007;104:7770–7775. doi: 10.1073/pnas.0608316104. [PMC free article] [PubMed] [Cross Ref]
97. Priel
A., Tuszy?ski J.A., Woolf N.J. Transitions in microtubule C-termini
conformations as a possible dendritic signaling phenomenon. Eur. Biophys. J. 2005;35:40–52. doi: 10.1007/s00249-005-0003-0. [PubMed] [Cross Ref]
98. Manning
G.S. The molecular theory of polyelectrolyte solutions with
applications to the electrostatic properties of polynucleotides. Q. Rev. Biophys. 1978;11:179–246. doi: 10.1017/S0033583500002031. [PubMed] [Cross Ref]
99. Le Bret M., Zimm B. Distribution of counterions around a cylindrical polyelectrolyte and Manning’s condensation theory. Biopolymers. 1984;23:287–312. doi: 10.1002/bip.360230209. [Cross Ref]
100. Satari? M.V., Ili? D.I., Ralevi? N., Tuszynski J.A. A nonlinear model of ionic wave propagation along microtubules. Eur. Biophys. J. 2009;38:637–647. doi: 10.1007/s00249-009-0421-5. [PubMed] [Cross Ref]
101. Ussing H.H., Thorn N.A. Transport Mechanisms in Epithelia. Academic Press; New York, NY, USA: 1973.
102. Porat-Shilom
N., Chen Y., Tora M., Shitara A., Masedunskas A., Weigert R. In vivo
tissue-wide synchronization of mitochondrial metabolic oscillations. Cell Rep. 2014;9:514–524. doi: 10.1016/j.celrep.2014.09.022. [PMC free article] [PubMed][Cross Ref]
103. Cosic I., Lazar K., Cosic D. Prediction of Tubulin resonant frequencies using the Resonant Recognition Model (RRM) IEEE Trans Nanobiosci. 2014;14:491–496. doi: 10.1109/TNB.2014.2365851. [PubMed] [Cross Ref]
104. Cosic I., Cosic D., Lazar K. Is it possible to predict electromagnetic resonances in proteins, DNA and RNA? EPJ Nonlinear Biomed. Phys. 2015;3:5. doi: 10.1140/epjnbp/s40366-015-0020-6. [Cross Ref]
105. Ayoub
A.T., Craddock T.J., Klobukowski M., Tuszy?ski J. Analysis of the
strength of interfacial hydrogen bonds between tubulin dimers using
quantum theory of atoms in molecules. Biophys. J. 2014;107:740–750. doi: 10.1016/j.bpj.2014.05.047.[PMC free article] [PubMed] [Cross Ref]
106. Pizzi
R., Strini G., Fiorentini S., Pappalardo V., Pregnolato M. Evidences of
new biophysical propeties of microtubules. In: Kwon S.J., editor. Artificial Networks. Nova Science Publishers, Inc.; New York, NY, USA: 2010.
107. Kobayashi S., Asai H., Oosawa F. Electric birefringence of actin. Biochim. Biophys. Acta Spec. Sect. Biophys. Subj. 1964;88:528–540. doi: 10.1016/0926-6577(64)90096-8. [PubMed] [Cross Ref]
108. Cantiello H.F., Patenaude C., Zaner K. Osmotically induced electrical signals from actin filaments. Biophys. J. 1991;59:1284–1289. doi: 10.1016/S0006-3495(91)82343-8.[PMC free article] [PubMed] [Cross Ref]
109. Lin
E.C., Cantiello H.F. A novel method to study the electrodynamic
behavior of actin filaments. Evidence for cable-like properties of
actin. Biophys. J. 1993;65:1371–1378. doi: 10.1016/S0006-3495(93)81188-3. [PMC free article] [PubMed] [Cross Ref]
110. Lonngren K.E. Observations of solitons on nonlinear dispersive transmission lines. In: Lonngren K.E., Scott A., editors. Solitons in Action. Academic Press; New York, NY, USA: 1978. pp. 127–152.
111. Oosawa F. Polyelectrolytes. Marcel Dekker, Inc.; New York, NY, USA: 1971.
112. Tuszy?ski J.A., Portet S., Dixon J.M., Luxford C., Cantiello H.F. Ionic wave propagation along actin filaments. Biophys. J. 2004;86:1890–1903. doi: 10.1016/S0006-3495(04)74255-1. [PMC free article] [PubMed] [Cross Ref]
113. Arsenault
M.E., Zhao H., Purohit P.K., Goldman Y.E., Bau H.H. Confinement and
manipulation of actin filaments by electric fields. Biophys. J. 2007;93:L42–L44. doi: 10.1529/biophysj.107.114538. [PMC free article] [PubMed] [Cross Ref]
114. Cho M.R., Thatte H.S., Lee R.C., Golan D.E. Reorganization of microfilament structure induced by AC electric fields. FASEB J. 1996;10:1552–1558. [PubMed]
115. Anderson
C.F., Record M.T.J. Ion distributions around DNA and other cylindrical
polyions: Theoretical descriptions and physical implications. Annu. Rev. Biophys. Bioeng. Chem. 1990;19:4232–4265. doi: 10.1146/annurev.bb.19.060190.002231.[PubMed] [Cross Ref]
Vestn Oftalmol. 2000 Jul-Aug;116(4):3-5.
Differentiated approaches to the treatment of nonstabilized primary
open-angle glaucoma with normalized intraocular pressure considering
its pathogenic features.
[Article in Russian]
Egorov VV, Sorokin EL, Smoliakova GP.
Clinical efficiency of dedystrophic treatments for nonstabilized
primary open-angle glaucoma (POAG) in the presence of normalized
intraocular pressure is compared in 168 patients (246 eyes). In one
group of patients ischemic angiopathy and hyperreactivity of optic
vessel adrenoreceptors associated with hypokinetic central hemodynamics
and constitutional metabolic status of the organism was corrected by
cinnarisin and riboxin. Patients with predominating congestive
angiopathy symptoms, hyper- or eukinetic circulation and “slow”
acetylation were treated by pantothenic acid, endotelon, and hyperbaric
oxygenation. In both groups epithalamine, eiconol, and magnetic laser
therapy were used, if indicated. This helped stabilize the process in
91% patients with initial POAG, in 87.5% with well-developed condition
vs. 66.1% and 38.2% patients treated by traditional therapy (period of
observation 3 years).
Vestn Oftalmol. 1996 Jan-Mar;112(1):6-8.
Possibilities of magnetotherapy in stabilization of visual function in patients with glaucoma.
[Article in Russian]
Bisvas Shutanto Kumar, Listopadova NA.
Courses of magnetotherapy (MT) using ATOS device with 33 mT magnetic
field induction were administered to 31 patients (43 eyes) with primary
open-angle glaucoma with compensated intraocular pressure. The
operation mode was intermittent, with 1.0 to 1.5 Hz field rotation
frequency by 6 radii. The procedure is administered to a patient in a
sitting posture with magnetic inductor held before the eye. The
duration of a session is 10 min, a course consists of 10 sessions.
Untreated eyes (n = 15) of the same patients were examined for control.
The patients were examined before and 4 to 5 months after MT course.
Vision acuity improved by 0.16 diopters, on an average, in 29 eyes
(96.7%) out of 30 with vision acuity below 1.0 before treatment.
Visocontrastometry was carried out using Visokontrastometer-DT device
with spatial frequency range from 0.4 to 19 cycle/degree (12
frequencies) and 125 x 125 monitor. The orientation of lattices was
horizontal and vertical. The contrasts ranged from 0.03 to 0.9 (12
levels). MT brought about an improvement of spatial contrast
sensitivity by at least 7 values of 12 levels in 22 (84.6%) out of 26
eyes and was unchanged in 4 eyes. Visual field was examined using
Humphry automated analyzer. A 120-point threshold test was used. After a
course of MT, visual field deficit decreased by at least 10% in 31
(72%) out of 43 eyes, increased in 3, and was unchanged in 9 eyes; on
an average, visual field deficit decreased by 22.4% vs. the initial
value. After 4 to 5 months the changes in the vision acuity and visual
field deficit were negligible. In controls these parameters did not
appreciably change over the entire follow-up period.
Oftalmol Zh. 1990;(3):154-7.
The effect of a pulsed electromagnetic field on the hemodynamics of eyes with glaucoma.
[Article in Russian]
Tsisel’skii IuV, Kashintseva LT, Skrinnik AV.
The influence of pulse electromagnetic field (PEMF) on hemodynamics
of the eye in open-angle glaucoma has been studied by means of a method
and a device proposed at the Filatov Institute. The PEMF
characteristics are: impulse frequency–50 Hz, exposition–0,02 sec.,
impulse shape–square, rate of impulse rise–4.10(4) c rate of magnetic
induction rise–2.10(4) mT/c, amplitude value of magnetic induction at
the impulse height–9.0–8.5 mT, duration of the procedure–7 min., a
course–10 sessions. Observations over 150 patients (283 eyes) with
latent, initial and advanced glaucoma have shown a positive influence
of PEMF on hemodynamics of a glaucomatous eye: a rise of rheographic
coefficient and relative volume pulse in 87,99 and 81,63%,
respectively. The degree of the rise and restoration frequency of
rheographic values of the glaucomatous eye under the influence of PEMF
to the age norm was more expressed at initial stages of the
glaucomatous process (latent and initial glaucoma).
Oftalmol Zh. 1990;(3):154-7.
The effect of a pulsed electromagnetic field on the hemodynamics of eyes with glaucoma.
[Article in Russian]
Tsisel’skii IuV, Kashintseva LT, Skrinnik AV.
The influence of pulse electromagnetic field (PEMF) on hemodynamics
of the eye in open-angle glaucoma has been studied by means of a method
and a device proposed at the Filatov Institute. The PEMF characteristics
are: impulse frequency–50 Hz, exposition–0,02 sec., impulse
shape–square, rate of impulse rise–4.10(4) c rate of magnetic induction
rise–2.10(4) mT/c, amplitude value of magnetic induction at the impulse
height–9.0–8.5 mT, duration of the procedure–7 min., a course–10
sessions. Observations over 150 patients (283 eyes) with latent, initial
and advanced glaucoma have shown a positive influence of PEMF on
hemodynamics of a glaucomatous eye: a rise of rheographic coefficient
and relative volume pulse in 87,99 and 81,63%, respectively. The degree
of the rise and restoration frequency of rheographic values of the
glaucomatous eye under the influence of PEMF to the age norm was more
expressed at initial stages of the glaucomatous process (latent and
initial glaucoma).
Oftalmol Zh. 1990;(2):89-92.
The effect of a pulsed electromagnetic field on ocular hydrodynamics in open-angle glaucoma.
[Article in Russian]
Tsisel’skii IuV.
The influence of pulse electromagnetic field on the hydrodynamics of
the eye in open-angle glaucoma has been studied using the method and the
device suggested at the Filatov Institute. The characteristics of the
action were: impulse frequency–50 Hz, duration–0.02 sec., pulse
form–rectangular, rate of pulse rise–4/10(-4) sec., rate of magnetic
induction rise–2/10(-4) mT/sec., amplitude value of magnetic induction
at the pulse level–8.0-8.5 mT, duration of the procedure–7 min. Ten
session in a total. Observations over 150 patients (283 eyes) with
latent, initial and advanced glaucoma have shown that the usage of pulse
electromagnetic field exerts influence on the hydrodynamics of the eye
in open-angle glaucoma; stimulates the rise of aqueous outflow and
production, the reduction of the Becker’s coefficient. At the latent
stage of the disease, normalization of outflow was recorded in 25% of
cases, at the initial and advanced stages–in 17.8% and 16.0% of cases,
respectively. The investigations carried out allow to recommend the
mentioned method for a complex treatment of open-angle glaucoma.
Vestn Oftalmol. 1994 Apr-Jun;110(2):5-7.
The effect of noninvasive electrostimulation of the optic nerve and
retina on visual functions in patients with primary open-angle glaucoma.
[Article in Russian]
Kumar BSh, Nesterov AP.
Electrostimulation courses with OEC-2 Ophthalmologic
Electrostimulator were administered to 30 patients (36 eyes) with
primary open-angle glaucoma and normal intraocular pressure. An active
electrode was placed on the upper lid, an indifferent one on the
forearm. Electric pulses (150-900 mcA) were grouped in several sessions,
30 sec each, divided by 30-45 sec intervals. Total duration of a
procedure was 16 min, the course consisting of 10 procedures. Control
group included 24 eyes of the same patients. The patients were examined
before, immediately, and 4-5 months after the treatment. Noticeable
changes in vision acuity and visual field were detected. Visual field
was examined using Humphrey Field Analyzer and 120-point threshold
related test. The treatment resulted in reduction of visual field
deficit by 10% or more in 28 (78%) of 36 eyes, in its increase in 2
eyes, and in no changes in 2 cases. Visual field deficit decreased by
25% on an average as against the initial value. Four to five months
after the treatment the changes in this parameter were negligible.
Vision acuity increased after the treatment in 31 of 36 eyes by 0.17
diopters on an average; 4 to 5 months later no changes occurred. In
control eyes no changes were detected either in vision acuity or visual
field during and after the treatment.
Low-frequency pulsed electromagnetic field therapy in fibromyalgia: a randomized, double-blind, sham-controlled clinical study.
Sutbeyaz ST, Sezer N, Koseoglu F, Kibar S.
Fourth Physical Medicine and Rehabilitation Clinic, Ankara Physical
Medicine and Rehabilitation Education and Research Hospital, Ankara,
Turkey. ssutbeyaz@yahoo.com
Abstract
OBJECTIVE: To evaluate the clinical effectiveness of low-frequency
pulsed electromagnetic field (PEMF) therapy for women with fibromyalgia
(FM).
METHODS: Fifty-six women with FM, aged 18 to 60 years, were randomly
assigned to either PEMF or sham therapy. Both the PEMF group (n=28) and
the sham group (n=28) participated in therapy, 30 minutes per session,
twice a day for 3 weeks. Treatment outcomes were assessed by the
fibromyalgia Impact questionnaire (FIQ), visual analog scale (VAS),
patient global assessment of response to therapy, Beck Depression
Inventory (BDI), and Short-Form 36 health survey (SF-36), after
treatment (at 4 wk) and follow-up (at 12 wk).
RESULTS: The PEMF group showed significant improvements in FIQ, VAS
pain, BDI score, and SF-36 scale in all domains at the end of therapy.
These improvements in FIQ, VAS pain, and SF-36 pain score during
follow-up. The sham group also showed improvement were maintained on all
outcome measures except total FIQ scores after treatment. At 12 weeks
follow-up, only improvements in the BDI and SF-36 scores were present in
the sham group.
CONCLUSION: Low-frequency PEMF therapy might improve function, pain, fatigue, and global status in FM patients.
Aesthetic Plast Surg. 2008 Jul;32(4):660-6. Epub 2008 May 28.
Effects of pulsed electromagnetic fields on postoperative pain: a
double-blind randomized pilot study in breast augmentation patients.
Hedén P, Pilla AA.
Department of Plastic Surgery, Akademikliniken, Storängsvägen 10, 115 42, Stockholm, Sweden. per.heden@ak.se
Abstract
BACKGROUND: Postoperative pain may be experienced after breast
augmentation surgery despite advances in surgical techniques which
minimize trauma. The use of pharmacologic analgesics and narcotics may
have undesirable side effects that can add to patient morbidity. This
study reports the use of a portable and disposable noninvasive pulsed
electromagnetic field (PEMF) device in a double-blind, randomized,
placebo-controlled pilot study. This study was undertaken to determine
if PEMF could provide pain control after breast augmentation.
METHODS: Forty-two healthy females undergoing breast augmentation for
aesthetic reasons entered the study. They were separated into three
cohorts, one group (n = 14) received bilateral PEMF treatment, the
second group (n = 14) received bilateral sham devices, and in the third
group (n = 14) one of the breasts had an active device and the other a
sham device. A total of 80 breasts were available for final analysis.
Postoperative pain data were obtained using a visual analog scale (VAS)
and pain recordings were obtained twice daily through postoperative day
(POD) 7. Postoperative analgesic medication use was also followed.
RESULTS: VAS data showed that pain had decreased in the active cohort
by nearly a factor of three times that for the sham cohort by POD 3 (p
< 0.001), and persisted at this level to POD 7. Patient use of
postoperative pain medication correspondingly also decreased nearly
three times faster in the active versus the sham cohorts by POD 3 (p
< 0.001).
CONCLUSION: Pulsed electromagnetic field therapy, adjunctive to
standard of care, can provide pain control with a noninvasive modality
and reduce morbidity due to pain medication after breast augmentation
surgery.
Pain Res Manag. 2007 Winter;12(4):249-58.
A randomized, double-blind, placebo-controlled clinical trial using a
low-frequency magnetic field in the treatment of musculoskeletal
chronic pain.
Thomas AW, Graham K, Prato FS, McKay J, Forster PM, Moulin DE, Chari S.
Bioelectromagnetics, Imaging Program, Lawson Health Research
Institute, Department of Medical Biophysics, Schulich School of Medicine
and Dentistry, University of Western Ontario, London, Canada. athomas@lawsonimaging.ca
Abstract
Exposure to a specific pulsed electromagnetic field (PEMF) has been
shown to produce analgesic (antinociceptive) effects in many organisms.
In a randomized, double-blind, sham-controlled clinical trial, patients
with either chronic generalized pain from fibromyalgia (FM) or chronic
localized musculoskeletal or inflammatory pain were exposed to a PEMF
(400 microT) through a portable device fitted to their head during
twice-daily 40 min treatments over seven days. The effect of this PEMF
on pain reduction was recorded using a visual analogue scale. A
differential effect of PEMF over sham treatment was noticed in patients
with FM, which approached statistical significance (P=0.06) despite low
numbers (n=17); this effect was not evident in those without FM (P=0.93;
n=15). PEMF may be a novel, safe and effective therapeutic tool for use
in at least certain subsets of patients with chronic, nonmalignant
pain. Clearly, however, a larger randomized, double-blind clinical trial
with just FM patients is warranted.
Pain Res Manag. 2006 Summer;11(2):85-90.
Exposure to a specific pulsed low-frequency magnetic field: a
double-blind placebo-controlled study of effects on pain ratings in
rheumatoid arthritis and fibromyalgia patients.
Lawson Health Research Institute, St. Joseph’s Health Care, London, Ontario N6A 4V2.
Abstract
BACKGROUND: Specific pulsed electromagnetic fields (PEMFs) have been
shown to induce analgesia (antinociception) in snails, rodents and
healthy human volunteers.
OBJECTIVE: The effect of specific PEMF exposure on pain and anxiety ratings was investigated in two patient populations.
DESIGN: A double-blind, randomized, placebo-controlled parallel design was used.
METHOD: The present study investigated the effects of an acute 30 min
magnetic field exposure (less than or equal to 400 microTpk; less than 3
kHz) on pain (McGill Pain Questionnaire [MPQ], visual analogue scale
[VAS]) and anxiety (VAS) ratings in female rheumatoid arthritis (RA)
(n=13; mean age 52 years) and fibromyalgia (FM) patients (n=18; mean age
51 years) who received either the PEMF or sham exposure treatment.
RESULTS: A repeated measures analysis revealed a significant
pre-post-testing by condition interaction for the MPQ Pain Rating Index
total for the RA patients, F(1,11)=5.09, P<0.05, estimate of effect
size = 0.32, power = 0.54. A significant pre-post-effect for the same
variable was present for the FM patients, F(1,15)=16.2, P<0.01,
estimate of effect size = 0.52, power =0.96. Similar findings were found
for MPQ subcomponents and the VAS (pain). There was no significant
reduction in VAS anxiety ratings pre- to post-exposure for either the RA
or FM patients.
CONCLUSION: These findings provide some initial support for the use
of PEMF exposure in reducing pain in chronic pain populations and
warrants continued investigation into the use of PEMF exposure for
short-term pain relief.
Neurosci Lett. 2001 Aug 17;309(1):17-20.
A comparison of rheumatoid arthritis and fibromyalgia patients and
health controls exposed to a pulsed (200 microT) magnetic field: effects
on normal standing balance.
Thomas AW, White KP, Drost DJ, Cook CM, Prato FS.
The Lawson Health Research Institute, Department of Nuclear Medicine
& MR, St. Joseph’s Health Care, 268 Grosvenor Street, London, N6A
4V2, Ontario, Canada. athomas@lawsonimaging.ca
Specific weak time varying pulsed magnetic fields (MF) have been
shown to alter animal and human behaviors, including pain perception and
postural sway. Here we demonstrate an objective assessment of exposure
to pulsed MF’s on Rheumatoid Arthritis (RA) and Fibromyalgia (FM)
patients and healthy controls using standing balance. 15 RA and 15 FM
patients were recruited from a university hospital outpatient
Rheumatology Clinic and 15 healthy controls from university students and
personnel. Each subject stood on the center of a 3-D forceplate to
record postural sway within three square orthogonal coil pairs (2 m,
1.75 m, 1.5 m) which generated a spatially uniform MF centered at head
level. Four 2-min exposure conditions (eyes open/eyes closed, sham/MF)
were applied in a random order. With eyes open and during sham exposure,
FM patients and controls appeared to have similar standing balance,
with RA patients worse. With eyes closed, postural sway worsened for all
three groups, but more for RA and FM patients than controls. The
Romberg Quotient (eyes closed/eyes open) was highest among FM patients.
Mixed design analysis of variance on the center of pressure (COP)
movements showed a significant interaction of eyes open/closed and
sham/MF conditions [F=8.78(1,42), P<0.006]. Romberg Quotients of COP
movements improved significantly with MF exposure [F=9.5(1,42),
P<0.005] and COP path length showed an interaction approaching
significance with clinical diagnosis [F=3.2(1,28), P<0.09]. Therefore
RA and FM patients, and healthy controls, have significantly different
postural sway in response to a specific pulsed MF.