ArticlePDF Available

Impedance Characterization and Modeling of Electrodes for Biomedical Applications

Authors:

Abstract and Figures

A low electrode-electrolyte impedance interface is critical in the design of electrodes for biomedical applications. To design low-impedance interfaces a complete understanding of the physical processes contributing to the impedance is required. In this work a model describing these physical processes is validated and extended to quantify the effect of organic coatings and incubation time. Electrochemical impedance spectroscopy has been used to electrically characterize the interface for various electrode materials: platinum, platinum black, and titanium nitride; and varying electrode sizes: 1 cm2, and 900 microm2. An equivalent circuit model comprising an interface capacitance, shunted by a charge transfer resistance, in series with the solution resistance has been fitted to the experimental results. Theoretical equations have been used to calculate the interface capacitance impedance and the solution resistance, yielding results that correspond well with the fitted parameter values, thereby confirming the validity of the equations. The effect of incubation time, and two organic cell-adhesion promoting coatings, poly-L-lysine and laminin, on the interface impedance has been quantified using the model. This demonstrates the benefits of using this model in developing better understanding of the physical processes occurring at the interface in more complex, biomedically relevant situations.
Content may be subject to copyright.
IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING, VOL. 52, NO. 7, JULY 2005 1295
Impedance Characterization and Modeling of
Electrodes for Biomedical Applications
Wendy Franks*, Iwan Schenker, Patrik Schmutz, and Andreas Hierlemann
Abstract—A low electrode-electrolyte impedance interface is
critical in the design of electrodes for biomedical applications. To
design low-impedance interfaces a complete understanding of the
physical processes contributing to the impedance is required. In
this work a model describing these physical processes is validated
and extended to quantify the effect of organic coatings and incu-
bation time. Electrochemical impedance spectroscopy has been
used to electrically characterize the interface for various electrode
materials: platinum, platinum black, and titanium nitride; and
varying electrode sizes: 1 cm
, and 900 m . An equivalent
circuit model comprising an interface capacitance, shunted by a
charge transfer resistance, in series with the solution resistance
has been fitted to the experimental results. Theoretical equations
have been used to calculate the interface capacitance impedance
and the solution resistance, yielding results that correspond well
with the fitted parameter values, thereby confirming the validity
of the equations. The effect of incubation time, and two organic
cell-adhesion promoting coatings, poly-L-lysine and laminin, on
the interface impedance has been quantified using the model.
This demonstrates the benefits of using this model in developing
a better understanding of the physical processes occurring at the
interface in more complex, biomedically relevant situations.
Index Terms—Electrochemical impedance spectroscopy, Pt, Pt
black, and TiN bioelectrodes.
I. INTRODUCTION
I
MPEDANCE characterization of the electrode-electrolyte
interface is of paramount importance in the fields of
impedance-based biosensing, neuroprotheses, and in vitro
communication with electrogenic cells. In impedance-based
biosensing, changes in the impedance are correlated to cell
spreading and locomotion [1], to bacterial growth [2], to
DNA hybridization [3] and to antigen-antibody reactions [4].
Neuroprotheses, and in particular cochlear implants, represent
an important application of impedance characterization. The
current applied to stimulate hearing via a cochlear implant is
determined from the known electrode impedance [5], which is
Manuscript received April 13, 2004; revised October 19, 2004. This work
was supported in part by the Information Societies Technology (IST) European
Union Future and Emerging Technologies program, and the Swiss Bundesamt
für Bildung und Wissenschaft (BBW). Asterisk indicates corresponding author.
*W. Franks is with the Physical Electronics Laboratory, ETH, Zurich 8093,
Zurich, Switzerland (e-mail: franks@phys.ethz.ch).
I. Schenker was with the Physical Electronics Laboratory, ETH, Zurich, and is
now with the Nonmetallic Inorganic Materials group, ETH Zurich, 8093 Zurich,
Switzerland (e-mail: iwan.schenker@mat.ethz.ch).
P. Schmutz is with the Laboratory for Corrision and Materials Integrity, Swiss
Federal Institute for Materials Testing and Research (EMPA), 8600 Dübendorf,
Switzerland (e-mail: Patrik.Schmutz@empa.ch).
A. Hierlemann is with the Physical Electronics Laboratory, ETH, Zurich
8093, Zurich, Switzerland (e-mail: hierlema@phys.ethz.ch).
Digital Object Identifier 10.1109/TBME.2005.847523
designed to be as low as possible to avoid cell damage [6]. For
the extracellular, in vitro monitoring of electrogenic cells, where
small microelectrodes are required for high-resolution stimula-
tion and recording, the need for a low interface impedance is
twofold [7]–[10]. During stimulation a certain current density is
necessary to generate activity. A high impedance would result
in a large applied electrode voltage leading to undesirable elec-
trochemical reactions that may be harmful to cellular cultures.
On the recording side, the extracellular signals are low, on
the order of millivolts for cardiomyocytes and microvolts for
neurons. The neural signals will be lost in the noisy, ion-based
electric fluctuations of the surrounding electrolyte media if the
electrode impedance is not low enough. A well-characterized,
fully understood interface impedance leads to an optimized
electrode-electrolyte interface design.
Equivalent circuit models have long been used to model the
interface impedance. In 1899 Warburg first proposed that the
interface could be represented by a polarization resistance in
series with a polarization capacitor [11]. Experimental findings
soon revealed that the polarization capacitance exhibited a
frequency dependency leading to the introduction of Fricke’s
law [12], and the use of a constant phase angle impedance to
represent the impedance of the interface capacitance. Randles’
work with rapid electrode reaction systems resulted in the
well-known Randles model, consisting of an interface capac-
itance shunted by a reaction impedance, in series with the
solution resistance [13]. As the use of platinum electrodes in
medical applications became ubiquitous, more research was
dedicated to the understanding of the electrode—physiological
solution interface. In the case of platinum, the resistive element
due to faradaic current was often omitted as measurement
equipment was not able to measure at low frequencies where
the impedance is finite, and not infinite, as was typically as-
sumed [14]. The work of Schwan and co-workers expanded on
previous work to include low frequency considerations [15]. Of
particular importance to biomedical applications is Schwan’s
limit of linearity: the voltage at which the electrode system’s
impedance becomes nonlinear, which is often exceeded during
stimulation [16], [17]. McAdams and colleagues extensively
studied the platinum pacing electrode (90% platinum and 10%
iridium) in physiological saline, successfully interpreting the
frequency-dependent nonlinear interface impedance [18]–[20].
Kovacs has presented an equivalent circuit model based on the
Randles model, with an additional Warburg impedance due to
the diffusion of faradaic current [7].
In the work presented here, electrochemical impedance spec-
troscopy (EIS) has been used to characterize the electrode-elec-
trolyte interface for various electrode materials commonly used
0018-9294/$20.00 © 2005 IEEE
1296 IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING, VOL. 52, NO. 7, JULY 2005
in biomedical applications: platinum, platinum black and tita-
nium nitride. An equivalent circuit model has been used where
each parameter represents a macroscopic physical quantity con-
tributing to the interface impedance: The model consists of an
interface capacitance, shunted by a charge transfer resistance, in
series with the solution resistance. The model parameters have
been tted to the experimental results. To conrm that the pa-
rameters do indeed represent the physical quantities, theoret-
ical equations have been used to calculate the parameter values
thereby validating the model. With respect to the measurement
technique, the effect of the initial interface conditions on the
charge transfer resistance is demonstrated. Measurements have
been performed for 1 cm
bright Pt, Pt black, and TiN, and for
900
m Pt black. As an extension to models that have been
presented in the past [5], [7], [20], this model has been used to
quantify the effect of organic cell-adhesion promoting coatings,
such as poly-L-lysine (PLL) and laminin, and the effect of incu-
bation time on the interface impedance. This demonstrates the
use of the model to develop a complete understanding of the
physical processes occurring at the interface in more complex,
biomedically relevant situations.
II. M
ETHODS
Both the macroelectrodes and microelectrodes were pro-
duced in-house according to the following procedures. For
the macroelectrodes a bare p-type Si wafer was electrically
isolated with 100 nm SiO
followed by 500 nm of Si N
deposited using plasma enhanced chemical vapor deposition
(PECVD). Sputter deposition was used to coat the substrate
with 50 nm of TiW, an adhesion promoter, followed by 270 nm
of Pt. The wafer was then diced and the chips were cleaned.
The chips were then either coated with Pt black or TiN. Pt
black was deposited using a 0.310 A/cm
current density in
a solution containing 7 mM hexachloroplatinic acid, 0.3 mM
lead acetate and hydrochloric acid to adjust the solution pH
to 1.0. Dendritically structured TiN was vapor deposited by
the Naturwissenschaftliche und Medizinische Institut at the
University of Tübingen, Germany [21]. For the microelectrodes
a p-type Si wafer with 100 nm of SiO
and 500 nm of Si N
was used as the substrate. As above, 50 nm TiW and 270 nm
Pt were sputter deposited, and the microelectrodes with leads
and bond pads were structured in a lift-off process. The wafer
was passivated with 500 nm Si
N . A reactive-ion etch (RIE)
was used to open the electrodes, thereby dening their size
and shape. The wafer was then diced, and the chips cleaned.
The bond wires of the packaged chips were encapsulated in
polydimethylsiloxane (PDMS) for electrical isolation.
Bright Pt macroelectrodes were coated with two different
cell-adhesion promoting coatings, laminin, and poly-L-lysine,
in the following manner. The samples were cleaned with an air
plasma treatment (2 min. at
mbar). It is known
from x-ray photon spectroscopy (XPS) measurements that,
immediately following plasma cleaning, the surface contains
no carbon, indicating that the surface is free of organic residues.
Additionally, plasma cleaning has the effect of activating the
surface and is believed to lead to better quality protein layers
with respect to coverage and adhesion. Within 45 min of
plasma cleaning the protein coatings were applied. Samples
were incubated (37
C, 5% CO ) overnight in laminin (20
mg/ml phosphate buffered saline, PBS). For PLL the samples
were incubated for 30 min in a 0.05% solution. After incubation
the samples were rinsed three times with PBS.
To investigate the effect of time on the interface impedance,
bright Pt samples were incubated in medium containing 10%
horse serum. Impedance measurements were performed after 7,
14, and 35 days using medium with serum as the electrolyte.
Measurements were performed using a commercially avail-
able Autolab PGSTAT30 potentiostat system with Frequency
Response Analysis software (version 4.9, Eco Chemie B.V.,
Netherlands). In this three-electrode system, a standard calomel
electrode (SCE) is the reference electrode, the counter elec-
trode is large-area Pt, and the electrolyte is physiological saline,
0.9% NaCl. In addition to the time dependency measurements,
neuron medium with 10% horse serum was used as the elec-
trolyte in one set of measurements. The perturbation potential
was 10 mV and the scan range was
to Hz, unless oth-
erwise noted. Measurements were typically performed with re-
spect to the open-circuit potential (OCP), the potential naturally
occurring between the working and reference electrodes. The
OCP is a function of the chemical composition of the interface,
and can signicantly affect the impedance results. Accordingly,
special attention was paid to the preparation of the samples to
ensure that the interface was as dened as possible. It is known
that organic residues on the substrate surface can contribute to a
faradaic current, which decreases the charge transfer resistance
and polarizes the electrode. An
in situ cleaning process was,
therefore, used where the Pt and Pt black electrodes (without
coatings) were treated by voltammetric cycling from
to
1.0 V for typically 6 cycles, at which point the measurement sta-
bilized. Cyclic voltammetry potentially results in the formation
and reduction of Pt oxide and Pt dioxide layers, as indicated by
the Pt-H
O Pourbaix diagram [22]. Since the OCP is a function
of the interface composition, the OCP can be altered through
the cycling process. Additionally, the OCP can be used as a
quality control to ensure that the initial conditions are the same
from measurement to measurement. The average OCP values
for Pt, Pt black, and TiN were
,
and
V, respectively. The relatively large range of
OCP values for TiN can be attributed to the fact that the samples
were not treated with cyclic voltammetry for fear of damaging
the dendritic structure.
Measurements were rst performed with relatively large area
samples (1 cm
) to establish the measurement technique and
equivalent circuit model under stable experimental conditions.
Difculties with respect to stability were encountered during the
measurement of the microelectrodes as the current approached
the system measurement limit of 10 nA at low frequencies. As a
result, the frequency range for the microelectrode measurements
was reduced to
to Hz.
III. E
QUIVALENT CIRCUIT MODEL
The equivalent circuit model presented in this work com-
prises a constant phase angle impedance
, that represents
the interface capacitance impedance, shunted by a charge
FRANKS et al.: IMPEDANCE CHARACTERIZATION AND MODELING OF ELECTRODES FOR BIOMEDICAL APPLICATIONS 1297
Fig. 1. Equivalent circuit model of electrode-electrolyte interface.
transfer resistance , together in series with the solution
resistance
, see Fig. 1, [23]. This model is an adaptation
from the theoretical models typically used to represent the
electrode-electrolyte impedance [7], [19], [24]. The Warburg
impedance due to diffusion of the chemical reactants in solution
is not included in this model. For the materials and frequency
range employed here, it was experimentally determined that
the Warburg impedance does not signicantly contribute to the
overall impedance.
A. Interface Capacitance
The constant phase angle impedance is a measure of the non-
faradaic impedance arising from the interface capacitance, or
polarization, and is given by the empirical relation [20]
(1)
where
is a measure of the magnitude of is a con-
stant
representing inhomogeneities in the surface
and
. In a Nyquist plot the angle between the data
and the abscissa axes gives
according to . When
represents a purely capacitive impedance element
corresponding to the interface capacitance.
A theoretical derivation of the interface capacitance is given
by the Gouy-Chapman-Stern model (GCS) [25]. The interface
capacitance is taken to be the series combination of the double-
layer capacitance, termed the Helmholz capacitance
, and
the diffuse layer capacitance, the Gouy-Chapman capacitance
, and is given by the following formula:
(2)
where
is the thickness of the double-layer, is the per-
mittivity of free space,
is the permittivity of the double layer,
is the charge on the ion in solution, is the applied electrode
potential, and
is the thermal voltage. The Debye length, ,
is given by
(3)
where
is the bulk number concentration of ions in solution
and
is the elementary charge. See Table I for the values of the
constants and variables used here.
B. Equilibrium Exchange Current Density and
At equilibrium, equal, and opposite reduction and oxidation
currents ow across the electrode-electrolyte interface. The
magnitude of these currents is termed the equilibrium exchange
current density
and is given by
(4)
TABLE I
S
UMMARY OF THE VALUES USED FOR VARIABLES AND CONSTANTS USED TO
CALCULATE THE THEORETICAL VALUE OF USING (2). INTHENOTE
C
OLUMN ALL ASSUMPTIONS ARE GIVEN
for the reduction reaction where
is Faradays constant, is
reduction reaction rate constant,
is the concentration of elec-
tron-acceptor ions A in solution plane of the interface,
is the
symmetry factor,
is the equilibrium potential, is the gas
constant and
is the temperature [26]. The equilibrium current
is a measure of the electrodes ability to participate in exchange
current reactions, and, hence, is of particular relevance to this
work. For an ideally polarizable electrode
equals zero and for
an ideally unpolarizable electrode
tends to innity; in reality,
lies somewhere between these two extremes. It is tempting
to use literature values of
to compare various electrode ma-
terials, however, since
is a function of ion concentration and
temperature, literature values must be carefully considered.
The equilibrium exchange current and
can be experimen-
tally determined through the application of the low-eld approx-
imation to the Butler-Volmer equation
, which
yields
(5)
where
is the measured current density, is the applied over-
potential,
is the number of electrons involved in the redox
reaction, and
mV at 298 K. Under the low-eld
approximation, the Butler-Volmer equation reduces to Ohms
law. A plot of the current versus overpotential yields a straight
line and the charge transfer resistance is given by the slope as
. Cyclic voltammetry was used to determine
under the following conditions: 5 mV perturbation signal
with respect to the OCP, 0.5 mV/s scan rate, 0.15 mV step po-
tential, averaged over 10 scans. For the case of Pt the charge
transfer arises from the electrolysis of H
O and reduction of O
according to where the equilibrium
potential is
(6)
with respect to the SCE [27]. For the calculation of
was,
therefore, assumed to be 4. It should be noted that the pres-
ence of contaminants at the interface will also contribute to the
faradaic current.
C. Solution Resistance
The resistance measured between the working electrode and
the reference electrode is termed the solution resistance. It can
1298 IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING, VOL. 52, NO. 7, JULY 2005
Fig. 2. Impedance modulus and phase as a function of frequency for 1 cm Pt,
Pt black, and TiN electrode materials [28]. Modeled results are indicated by the
smooth lines.
be determined from the spreading resistance, the resistance en-
countered by current spreading out into solution, under the as-
sumption that the counter electrode is innitely large and the
working electrode is surrounded by electrolyte. The spreading
resistance is given by
(7)
for square electrodes (
is the solution resistivity, 72 cm for
physiological saline [7],
the electrode side length), and
(8)
for round electrodes (
is the radius) [7]. It is worth noting that
unlike
and , which scale with the total electrode area, the
solution resistance is dependent upon the geometric area only
(where geometric area refers to the planar two-dimensional area,
and not the larger total area which increases with roughness).
IV. R
ESULTS
EIS measurements and model results for 1 cm Pt, Pt black,
and TiN are given in Fig. 2, [28]. Table II gives a summary of
the averaged, tted parameter values with corresponding stan-
dard deviations. For
, and the standard deviation is 10%,
or less, in almost all cases. In the case of
the standard devia-
tion is as high as 110%. This large standard deviation can be at-
tributed to the high sensitivity of
to the electrode-electrolyte
interface conditions. Furthermore,
is extrapolated from the
low frequency data which leads to increased uncertainty. The
sharply decreasing phase angles at high frequencies is an arti-
fact of the measurement system.
A. Verification of Model Parameters
To conrm the validity of the model, the model parameters
have been calculated based on the theoretical principles pre-
sented above. A comparison between the tted and theoretical
parameters has been performed for the simplest case of bright
TABLE II
S
UMMARY OF THE AVERAGED,FITTED PARAMETER RESULTS,WITH
CORRESPONDING STANDARD DEVIATIONS, FOR THE FOLLOWING MATERIALS
AND
MEASUREMENT AREAS:BRIGHT Pt 1 cm ,PtBLACK 1cm , TiN 1 cm ,
Pt B
LACK MICROELECTRODE 900 m , PLL-COATED BRIGHT Pt 1
cm
,LAMININ-COATED BRIGHT Pt 1 cm ,BRIGHT Pt WITH NEURON
MEDIUM AS THE ELECTROLYTE,1cm
Fig. 3. Equilibrium exchange current density measurement: current versus
applied electrode overpotential measurement results for 1 cm
bright platinum.
A
mV overpotential with respect to an OCP of 351.5 mV was applied.
Pt. Using (2) and the variable values given in Table I, the theo-
retical interface capacitance was found to be 0.545 F/m
. The
impedance of the interface capacitance was calculated using
, and at an angular frequency of 1 s ,
is (for 1 cm ). Similarly, using the values of and
(given in Table II) for 1 cm bright Pt, and (1), was
found to be
.For , cyclic voltammetry was used
to generate a current versus overpotential plot, see Fig. 3. From
the slope of the plot,
was found to be 300 k ; the tted
value was found to be 450 k
. From (5), the exchange current
density was calculated to be
A/cm . Finally the solu-
tion resistance was determined to be 32.0
using (8). The tted
parameter value is 28.0
.
B. Effect of OCP on Model Results
Throughout the course of this work it has been observed that
the model parameter
is highly dependent on the initial elec-
trode-electrolyte interface conditions. To demonstrate this de-
pendance, the OCP was set to various values, either by cyclic
voltammetry or by an applied potential. The EIS phase results
(the modulus results show no variation) and the tted values for
are given in Fig. 4.
FRANKS et al.: IMPEDANCE CHARACTERIZATION AND MODELING OF ELECTRODES FOR BIOMEDICAL APPLICATIONS 1299
Fig. 4. Effect of OCP on measurements results. Measurements were
performed using 1 cm
bright platinum samples. The table shows the tted
values for
.
Fig. 5. Impedance modulus and phase experimental results for Pt black
microelectrodes (geometrical area 900
m ). The modeled results are indicated
by the solid lines [28].
C. Microelectrode Results
The impedance modulus, phase, and standard deviations
for the Pt black microelectrodes (geometric area 900
m )
are presented in Fig. 5; tted parameter values are given in
Table II. A variation in the total electrode area within the
sample population arising from the platinization procedure
results in a large measurement standard deviation apparent in
the phase measurements at high frequencies. The theoretical
value for the interface capacitance, 0.545 F/m
[from (2)] and
the relation
, which simplies
to
, were used to determine the total area of
each of the 6 microelectrodes measured. The areas ranged
from
to m , with a standard deviation of
m . Development of a more uniform platinization
process is currently under investigation.
D. Effect of Coatings and Neuron Medium as Electrolyte
Fig. 6 shows the impedance modulus and phase results for
the two celladhesion promoting coatings, laminin
and PLL , and for the measurements performed with
neuron medium
as the electrolyte (see Table II for tted
parameter values). The charge transfer resistance of PLL and
laminin is 1.24 and 0.25 M
, respectively, a difference which
quanties the relative reactivity of the materials at the given
conditions. For the PLL and laminin coatings the value for
was found to be 8.6 and 5.0 s/ , respectively. In the case
of the neuron medium as the electrolyte, the parameter value for
is similar to that of bright Pt. The value of was found to
be 14.6
s/ which is smaller than that of bright Pt. The
solution resistance of the neuron medium was found to be 40.4
, which yields a value of 90.3 cm [from (8)] for the neuron
medium resistivity.
Fig. 6. Impedance modulus and phase for the laminin and PLL protein
coatings, and the neuron medium as electrolyte. The modeled results are
indicated by the solid lines.
Fig. 7. Relative change in modeled parameter value as a function of time for
1cm
bright Pt incubated in medium containing 10% horse serum for up to 35
days.
E. Effect of Time
The percent change in model parameter results for impedance
recordings performed at days 1, 7, 14, and 35 in incubation are
given in Fig. 7. In all cases, the parameter value increases from
day 1 to day 7, and then levels off. This demonstrates the effec-
tive encapsulation of the electrodes, as is discussed in the fol-
lowing section.
V. D
ISCUSSION
An equivalent circuit model has been used to describe and
analyze the EIS experimental results. A good match between
the curves generated using the equivalent circuit model and the
measurement results indicates that the appropriate model has
been selected. For the simplest case of bright Pt, equations de-
scribing the macroscopic physical processes occurring at the
electrode-electrolyte interface yield parameter values for
and
that are in good agreement with the experimentally derived
parameters. The difference between
and repre-
sents a 15% deviation from the experimental value, which is
evidence that the GCS model describes well the interface ca-
pacitance. The difference between the tted and calculated the-
oretical value could be a result of impurities at the interface that
are not accounted for in the GCS model. A 14% deviation from
1300 IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING, VOL. 52, NO. 7, JULY 2005
the measured value of the solution resistance indicates that (7)
can be used to calculate
.
Cyclic voltammetry was used to measure
for the bright
Pt-0.9% NaCl system. There is a 50% difference between the
tted
value and that determined from cyclic voltammetry.
The source of this difference is difcult to quantify, however,
is most likely a result of sensitivities, in both the EIS and the
cyclic voltammetry measurements, to the interface conditions.
It is worthwhile to mention that the
value reported here
is four orders of magnitude greater than the values reported
in literature,
cm [10], [29]. The literature values are
erroneously based on the assumption that the faradaic current
owing across the interface is due to a hydrogen redox reaction.
For a neutral pH, the equilibrium potential of the hydrogen re-
action is
V, signicantly negative of the OCP value of
V. Given the positive OCP value, the theoretical
reaction would be the oxidation of hydrogen gas according to:
, although the limited amount of hydrogen
gas available in solution precludes this reaction from occurring.
It is known from literature that, given the experimental condi-
tions used in this work, it is the O
redox reaction that is mainly
responsible for the faradaic current (although contamination at
the interface will also contribute to charge transfer) [27]. Exper-
iment results presented here support this assumption. When the
OCP is set to value higher or lower than the equilibrium value
of approximately 0.6 V [from (6) using a pH of 7.0], the
is
lower than if the OCP were equal to the equilibrium value. The
higher or lower potential shifts the reaction from equilibrium
conditions, more faradaic charge transfer occurs, and
is re-
duced. For example, for an OCP of 0.66 V the change transfer
resistance is 1.0 M
(Fig. 4). For an OCP of 0.39 and 0.92 V the
charge transfer resistance reduces to 0.43 M
and 0.23 k , re-
spectively. Since
depends on , it is interesting to compare
the
value measured here with those reported in literature. The
value reported by Kovacs for the oxygen reaction is
A/cm , which was experimentally determined using a platinum
electrode and aerated frog Ringers solution (115 mM NaCl, 2.0
mM KCl, 1.8 mM CaCl
) [7], [30]. Aeration of the electrolyte
will increase the amount of O
available for reaction, thereby in-
creasing the equilibrium exchange current, which explains the
deviation from the value reported here. The value reported by
McAdams [19] for a platinum pacing electrode in physiological
saline is
A/cm , which is similar to our measured
value of
A/cm , albeit a direct comparison is invalid
due to the 10% iridium content in the pacing electrode.
It has been the goal of this study to use the EIS results
and model parameter values to develop an enhanced under-
standing of the effect of the physical processes on the interface
impedance, with the expressed purpose of improving the inter-
face design. For example, the impedance modulus of Pt black
and TiN is two orders of magnitude smaller than for bright Pt,
clearly demonstrating the effect of increased total surface area.
This nding is expected and is well documented in previous
studies [10], [29], [31]. A more interesting use of the results
can be found in the effects of PLL and laminin on the interface
impedance. Qualitatively, there is little difference between the
impedance modulus of PLL and laminin, however, the phase
results show that the phase decreases more rapidly for laminin
that PLL, indicating a lower
for laminin. Quantitatively,
this difference manifests itself in the tted parameter results,
where
is 1.24 M for PLL versus 0.25 M for laminin.
Although laminin is a thicker coating
nm compared to
nm, it appears to facilitate charge transfer rather than
to impede reactions. The effect of the coating thicknesses is
apparent in the modeled parameter
, which is approximately
of for bright Pt. Since is representative of the interface
capacitance, which is inversely proportional to the dielectric
thickness (in this case the coating), it follows that the coatings
would reduce
. Such effects are important when considering
an optimized interface design for the stimulation and recording
of electrical activity from electrogenic cells, or when designing
biosensors based on changes in the interface capacitance.
Additionally, the model has been used the quantify the ef-
fect of incubation time on the interface impedance. The relative
change in the model parameter results over time show a sharp
increase from day 1 to 7, followed by a plateau. Indeed, the
impedance results themselves (not shown here) do not change
signicantly from day 7 to 35. This indicates that the electrodes
have been encapsulated, most likely by proteins contained in
the medium. For example, the value of
increases from 0.91
at day 1 to 0.95 at day 35, indicating that the interface has be-
come more capacitive. Similarly,
increases from 217 k at
day 1 to 556 k
at day 35, again demonstrating that the inter-
face has become more capacitive, or less conducive to charge
transfer. The same is true of
, which increases from an ini-
tial value of 6.5
s/ to 9.7 s/ by day 35. These
results are relevant to researchers who are performing electro-
physiological measurements over extended periods of time. The
changing electrode impedance of electrodes covered by cell cul-
tures would additionally provide useful insights and shall be in-
vestigated in the future.
The ease in analyzing surface conditions using EIS and its
widespread applicability gives merit to a short discussion on
how this technique may be extended to other electrode-elec-
trolyte systems and smaller electrode sizes. In order to use this
technique with different systems, such as iridium oxide, gold,
stainless steel and other biomedical metals with electrolytes of
varying concentrations and compositions, the appropriate equiv-
alent circuit, which is highly dependent on the equilibrium ex-
change current density, must be selected. For nonpolarizable
electrodes, corresponding to a high equilibrium exchange cur-
rent, the charge transfer resistance becomes low and a War-
burg element, representing the effect of diffusion on the cur-
rent-carrying ions, may be necessary to accurately model the
interface conditions (the reader is referred to general references
on EIS [25], [26], [32]). Due to the sensitivity of
to the inter-
face conditions, as demonstrated in this work, the equilibrium
exchange current varies signicantly with electrode and elec-
trolyte and, hence, care must be taken when using values from
literature. It is, therefore, recommended that
be experimen-
tally determined. Although values for
are reported in litera-
ture, for example Kovacs provides a table with various values
[7], it is unlikely that the experimental conditions are exactly
the same. Pourbaix diagrams may be used to determine the re-
actions giving rise to
, and will provide a value for that is
required to determine
using (5). Many cases of EIS applied
FRANKS et al.: IMPEDANCE CHARACTERIZATION AND MODELING OF ELECTRODES FOR BIOMEDICAL APPLICATIONS 1301
to different electrode-electrolyte systems can be found in litera-
ture, for example: Bates, et al., have analyzed a system of rough-
ened Pt and aqueous H
SO [33], Valoen, et al., have developed
an impedance model for metal hydride electrodes in KOH [34],
and Chou and colleagues have used EIS to investigate the effect
of a self-assembled monolayer on a Au electrode inside a rat
heart [35]. As a nal extension to the model, it should be noted
that while this theory may be applied to innitely small elec-
trodes, the experimental measurement becomes increasingly un-
stable with shrinking electrode dimensions because the current
owing across the interface becomes too small to be measured.
Very small electrodes,
m , must be measured at higher
frequencies requiring special equipment.
VI. C
ONCLUSION
A measurement technique with a corresponding equivalent
circuit model has been established for the quantication of
the electrode-electrolyte interface impedance using electro-
chemical impedance spectroscopy. Equations describing the
macroscopic physical processes occurring at the interface are
presented, and, for the case of bright Pt, yield results that are
in good agreement with the tted parameter values. The effect
of various characteristics, such as total area, protein coating,
and time on the EIS and tted parameter results has been used
to elucidate the processes occurring at the interface, thereby
demonstrating the usefulness of the model in the impedance
analysis of more complex, biomedically relevant situations.
A
CKNOWLEDGMENT
The authors would like to thank Prof. H. Baltes (on leave)
for sharing laboratory resources and for his ongoing stimulating
interest in their work.
R
EFERENCES
[1] J. Wegener, C. R. Keese, and I. Giaever, Electric cell-substrate
impedance sensing (ECIS) as a noninvasive means to monitor the
kinetics of cell spreading to articial surfaces,
Exp. Cell Res., vol. 259,
pp. 158166, 2000.
[2] L. Yang, C. Ruan, and Y. Li, Detection of viable salmonella ty-
phimurium by impedance measurement of electrode capacitance
and medium resistance, Biosensors Bioelectron., vol. 19, no. 5, pp.
495502, 2003.
[3] C. A. Marquette, I. Lawrence, C. Polychronakos, and M. F. Lawrence,
Impedance based DNA chip for direct measurement, Talanta, vol.
56, pp. 763768, 2002.
[4] V. M. Mirsky, M. Riepl, and O. S. Wolfbeis, Capacitive monitoring of
protein immobilization and antigen-antibody reactions on monomolec-
ular alkylthiol lms on gold electrodes, Biosensors Bioelectron., vol.
12, pp. 977989, 1997.
[5] C. Q. Huang, R. K. Shepherd, P. M. Center, P. M. Seligman, and B.
Tabor, Electrical stimulation of the auditory nerve: Direct current
measurement in vivo, IEEE Trans. Biomed. Eng., vol. 46, no. 4, pp.
461469, Apr. 1999.
[6] M. Tykocinski, Y. Duan, B. Tabor, and R. S. Cowan, Chronic electrical
stimulation of the auditory nerve using high surface area (HiQ) platinum
electrodes,Hearing Res., vol. 159, pp. 5368, 2001.
[7] G. T. A. Kovacs, Introduction to the theory, design, and modeling of
thin-lm microelectrodes for neural interfaces, in Enabling Technolo-
gies for Cultured Neural Networks, D. A. Stenger and T. M. McKenna,
Eds. London, U.K.: Academic, 1994, pp. 121165.
[8] G. W. Gross, B. K. Rhoades, D. L. Reust, and F. U. Schwalm, Stim-
ulation of monolayer networks in culture through thin-lm indium-tin
oxide recording electrodes,J. Neurosci. Methods, vol. 50, pp. 13143,
1993.
[9] P. Thiebaud, C. Beuret, M. Koudelka-Hep, M. Bove, S. Martinoia, M.
Grattarola, H. Jahnsen, R. Rebaudo, M. Balestrino, J. Zimmer, and Y.
Dupont, An array of Pt-tip microelectrodes for extracellular monitoring
of activity of brain slices, Biosensors Bioelectron., vol. 14, pp. 6165,
1999.
[10] M. O. Heuschkel, Fabrication of multielectrode array devices for elec-
trophysiological monitoring of in vitro cell/tissue cultures,in Series in
Microsystems, P. A. Besse, M. Gijs, R. S. Popovic, and Ph. Renaud,
Eds. Konstanz, Germany: Hartung-Gorre Verlag, 2001, vol. 13.
[11] E. Warburg, Ueber das Verhalten sogenannter unpolarisbarer Elek-
troden gegen Wechselstrom,Annalen der Physik und Chemie, vol. 67,
pp. 493499, 1899.
[12] H. Fricke, The theory of electrolytic polarization, Philosophical Mag.,
vol. 7, pp. 310318, 1932.
[13] J. E. B. Randles, Kinetics of rapid electrode reactions, Discussions
Faraday Soc., vol. 1, pp. 1119, 1947.
[14] D. Jaron, H. P. Schwan, and D. B. Geselowitz, A mathematical model
for the polarization of cardiac pacemaker electrodes, Med. Biol. Eng.,
vol. 6, p. 579, 1968.
[15] B. Onaral and H. P. Schwan, Linear and nonlinear properties of plat-
inum electrode polarization. I. Frequency dependence at very low fre-
quencies,Med. Biol. Eng. Comput., vol. 20, pp. 299306, 1982.
[16] H. P. Schwan, Electrode polarization impedance and measurements in
biological materials,Ann. New York Acad. Sci., vol. 148, pp. 191209,
1968.
[17]
, Linear and nonlinear electrode polarization and biological mate-
rials,Ann. Biomed. Eng., vol. 20, pp. 26988, 1992.
[18] E. McAdams, Effect of surface topography on the electrode-electrolyte
interface impedance,Surface Topogr., vol. 2, pp. 107122, 1989.
[19] E. T. McAdams and J. Jossinet, Physical interpretation of Schwans
limit voltage of linearity, Med. Biol. Eng. Comp., vol. 32, pp. 12630,
1994.
[20] E. T. McAdams, A. Lackermeier, J. A. McLaughlin, D. Macken, and J.
Jossinet, The linear and nonlinear electrical properties of the electrode-
electrolyte interface, Biosensors Bioelectron., vol. 10, pp. 6774, 1995.
[21] M. Janders, U. Egert, M. Stelzle, and W. Nisch, Novel thin lm titanium
nitride micro-electrodes with excellent charge transfer capability for cell
stimulation and sensing applications, in Proc. IEEE Eng. Med. Biol.
Soc., Amsterdam, The Netherlands, 1997.
[22] M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions,
2nd ed. Brussels, Belgium: National Association of Corrosion Engi-
neers, 1974.
[23] W. Franks, F. Heer, I. McKay, S. Taschini, R. Sunier, C. Hagleitner, A.
Hierlemann, and H. Baltes, CMOS monolithic microelectrode array for
stimulation and recording of natural neural networks, presented at the
Transducers03, Boston, MA, 2003.
[24] D. C. Grahame, Mathematical theory of the faradaic admittance, J.
Electrochem. Soc., vol. 99, pp. 370C385C, 1952.
[25] A. J. Bard and L. R. Faulkner, Electrochemical Methods. New York:
Wiley, 2001.
[26] J. O. M. Bockris and A. K. N. Reddy, Modern Electrochemistry.New
York: Plenum, 1970, vol. 2.
[27] E. Yeager, Electrocatalysts for O2 reduction, Electrochimica Acta, vol.
29, pp. 15271537, 1984.
[28] F. Heer, W. Franks, A. Blau, S. Taschini, C. Ziegler, A. Hierlemann, and
H. Baltes, CMOS microelectrode array for monitoring of electrogenic
cells,Biosensors Bioelectron., vol. 20, pp. 358366, 2004.
[29] D. A. Borkholder, Cell based biosensors using microelectrodes,Ph.D.
Thesis, Stanford Univ., Stanford, CA, 1998.
[30] P. W. Davies, The oxygen cathode,in Physical Techniques in Biolog-
ical Research, W. L. Nastuk, Ed. London, U.K.: Academic, 1962, vol.
IV, pp. 137179.
[31] J. D. Weiland, D. J. Anderson, and M. S. Humayun, In vitro electrical
properties for iridium oxide versus titanium nitride stimulating elec-
trode,IEEE Trans. Biomed. Eng., vol. 49, no. 12, pp. 15741579, Dec.
2002.
[32] J. R. Macdonald, Impedance Spectroscopy. New York: Wiley, 1987.
[33] J. Bates and Y. Chu, Electrode-electrolyte interface impedance: Exper-
iments and model, Annals of Biomed. Eng., vol. 20, pp. 349363, 1992.
[34] L. O. Valeon, S. Sunde, and R. Tunold, An impedance model for elec-
trode processes in metal hydride electrode,J. Alloys Compounds, vol.
253254, pp. 656659, May 20, 1997.
[35] H. A. Chou, D. H. Zavitz, and M. Ovadia, In vivo CH
(CH ) SAu
SAM electrodes in the beating heart: in situ analytical studies relevant
to pacemakers and interstitial biosensors, Biosens. Bioelectron., vol. 18,
pp. 1121, 2003.
1302 IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING, VOL. 52, NO. 7, JULY 2005
Wendy Franks received the B.Sc. degree in chem-
ical engineering in 1998 and the M.Sc. degree in
electrical engineering in 2000, both at the University
of Waterloo, Waterloo, ON, Canada. She received
the Ph.D. degree from the Swiss Federal Institute of
Technology, Zurich (ETHZ) in the eld of bioelec-
tronics.
Iwan Schenker received the Diploma in physics
from the Swiss Federal Institute of Technology
(ETHZ), Zurich, Switzerland, in 2003. He is cur-
rently working towards the Ph.D. degree in the
Nonmetallic Inorganic Materials Group, Department
of Materials, ETHZ, Zurich, Switzerland.
Patrik Schmutz received an undergraduate degree in
solid-state physics in 1991 from the University of Fri-
bourg, Switzerland, and a Ph.D. degree in science in
1996 from the Swiss Federal Institute of Technology
in Lausanne (EPFL).
He is currently Head of Corrosion Research at the
EMPA (National Laboratory for Material Science
and Technology) and is a Lecturer at the Swiss
Federal Institute of Technology in Zurich (ETHZ),
Switzerland. His main research topic is investigation
of localized physico-(electro)chemical processes on
reactive metallic surfaces.
Andreas Hierlemann received the Diploma in
chemistry in 1992 and the Ph.D. degree in physical
chemistry in 1996 from the University of Tübingen,
Tübingen, Germany.
Having been a Postdoc at Texas A&M Uni-
versity, College Station, TX (1997), and Sandia
National Laboratories, Albuquerque, NM (1998),
he is currently Professor at the Physical Electronics
Laboratory at ETH Zurich in Switzerland. The
focus of his research activities is on CMOS-based
microsensors and interfacing CMOS electronics
with electrogenic cells.
... This paper is a follow-up of the work in [1], which aimed to deliver a closed-form analytical formula for the grounding impedance of a hemispherical electrode, accounting for the resistive and reactive contributions of the ground medium. As referred to in [1], the importance of the theme is supported by an enormous variety of applications where metallic electrodes are used, namely for the characterization of soils in geophysics and civil engineering [2,3]; for the characterization of tissue properties and sensing applications in biophysics and bioengineering [4][5][6]; and for grounding purposes aimed to provide electric protection to people and infrastructures in electrical engineering, [7][8][9]. ...
... where the overbars are intended to denote complex amplitudes (phasors) of the sinusoidal time-varying fields they represent, e.g., H(ρ, z, t) = Re H(ρ, z, ω)e jωt . In (5), E denotes the complex amplitude of the electric field intensity and τ is the soil relaxation time. The so-called H-formulation [15][16][17] for the electromagnetic field is a consequence of the previous fundamental field equations: ...
Article
Full-text available
Metallic electrodes are widely used in many applications, the analysis of their frequency-domain behavior is an important subject, particularly in applications related to earthing/grounding systems, from dc up into the MHz range. In this paper, a numerical evaluation of the frequency-dependent complex impedance of the hemispherical ground electrode is implemented. A closed-form solution for non-zero frequencies is still a difficult task to achieve as evidenced in a previous paper dedicated to the subject and, therefore, numerical approaches should be an alternative option. The aim of this article is to present a solution based on a numerical method using finite element analysis. In typical commercial FE tools, electric currents exhibit azimuthal orientation and, as such, the magnetic field has a null azimuthal component but non-null axial and radial components. On the contrary, a dual problem is considered in this work, with a purely azimuthal magnetic field. To overcome the difficulty of directly using a commercial FE tool, a novel formulation is developed. An innovative 2D formulation, the ι-form, is developed as a modification of the H-formulation applied to axisymmetric magnetic field problems. The results are validated using a classical 3D H-formulation; comparisons showed very good agreement. The electrode complex impedance is analyzed considering two different cases. Firstly, the grounding system is constituted by a hemispherical electrode surrounded by a remote concentric electrode; in the second case, the grounding system is constituted by two identical thin hemispherical electrodes. Computed results are presented and discussed, showing how the grounding impedance depends on the frequency and, also, on the radius of the remote concentric electrode (first case) or on the distance between the two hemispherical electrodes (second case).
... We adopted the lumped model introduced in [30] that was used to simulate realistic HFO signals through combination with an electrode-tissue interaction (ETI) model and a 3D neural network model mimicking hippocampal recordings in rodent models. While its ETI model is also an adaptation of common circuit-equivalent representations [22], [55]- [58], this model simulated fast ripples by introducing hyperexcitable neurons in the network and calculating a discretized net potential. Signals recorded from a mouse hippocampus under the Kainate model of temporal lobe epilepsy served to validate this computational simulation. ...
Article
Full-text available
italic xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">Objective: High-frequency oscillations (HFOs) are a promising prognostic biomarker of surgical outcome in patients with epilepsy. Their rates of occurrence and morphology have been studied extensively using recordings from electrodes of various geometries. While electrode size is a potential confounding factor in HFO studies, it has largely been disregarded due to a lack of consistent evidence. Therefore, we designed an experiment to directly test the impact of electrode size on HFO measurement. Methods: We first simulated HFO measurement using a lumped model of the electrode-tissue interaction. Then eight human subjects were each implanted with a high-density 8x8 grid of subdural electrodes. After implantation, the electrode sizes were altered using a technique recently developed by our group, enabling intracranial EEG recordings for three different electrode surface areas from a static brain location. HFOs were automatically detected in the data and their characteristics were calculated. Results: The human subject measurements were consistent with the model. Specifically, HFO rate measured per area of tissue decreased significantly as electrode surface area increased. The smallest electrodes recorded more fast ripples than ripples. Amplitude of detected HFOs also decreased as electrode surface area increased, while duration and peak frequency were unaffected. Conclusion: These results suggest that HFO rates measured using electrodes of different surface areas cannot be compared directly. Significance: This has significant implications for HFOs as a tool for surgical planning, particularly for individual patients implanted with electrodes of multiple sizes and comparisons of HFO rate made across patients and studies.
... Review articles have been written which provide guidance on the application of impedance spectroscopy for biomedical applications, but these do not provide sample data. One previous study [15] suggested a simplified model for the electrode-electrolyte interface that accounts for ohmic resistance in series with a parallel combination of a constant phase element (CPE) and a resistance. They argued that mass transfer was unimportant for their system. ...
Article
Full-text available
Invasive intracranial electrodes are used in both clinical and research applications for recording and stimulation of brain tissue, providing essential data in acute and chronic contexts. The impedance characteristics of the electrode–tissue interface (ETI) evolve over time and can change dramatically relative to pre-implantation baseline. Understanding how ETI properties contribute to the recording and stimulation characteristics of an electrode can provide valuable insights for users who often do not have access to complex impedance characterizations of their devices. In contrast to the typical method of characterizing electrical impedance at a single frequency, we demonstrate a method for using electrochemical impedance spectroscopy (EIS) to investigate complex characteristics of the ETI of several commonly used acute and chronic electrodes. We also describe precise modeling strategies for verifying the accuracy of our instrumentation and understanding device–solution interactions, both in vivo and in vitro. Included with this publication is a dataset containing both in vitro and in vivo device characterizations, as well as some examples of modeling and error structure analysis results. These data can be used for more detailed interpretation of neural recordings performed on common electrode types, providing a more complete picture of their properties than is often available to users.
... The numerical model regards electrochemical processes at the ETI via an equivalent circuit model. The equivalent circuit model used in this study consists of a parallel connection of a constant phase element (CPE) with the impedance Z CPE and a charge transfer resistance R CT [43], [44]. The constant phase element is defined as ...
Article
Full-text available
Objective: Electrical stimulation is known to enhance bone healing. Novel electrostimulating devices are currently being developed for the treatment of critical-size bone defects in the mandible. Previous numerical models of these devices did not account for possible uncertainties in the input data. We present the numerical model of an electrically stimulated minipig mandible, including optimization and uncertainty quantification (UQ) methods that allow us to determine the most influential parameters. Methods: Uncertainties in the optimized finite element model are quantified using the polynomial chaos method that is implemented in the open-source Python toolbox Uncertainpy. The volumes of understimulated, beneficially stimulated, and overstimulated tissue are considered quantities of interest because they may significantly impact the expected healing success. Further, the current is a substantial quantity, limiting the lifetime of a battery-driven stimulation unit. With sensitivity analyses, the most critical parameters in the numerical model can be identified. Thus, we can learn which parameters are particularly relevant, for example, when conceptualizing the stimulation unit or planning the manufacturing process. Results: The results of this study show that the parameters of the electrode-tissue interface (ETI), as well as the conductivity within the defect volume, have the most significant impact on the model results. Conclusions: The UQ results suggest that careful characterization of the ETI and the dielectric tissue properties is crucial to reduce these uncertainties. Significance: The numerical model regarding uncertainties yields important implications for reliable implant design and clinical translation.
... Because of these factors, the process of characterizing biomaterials is essential and entails a thorough examination of the mechanical, chemical, biological, and physical characteristics of materials intended for use in human interaction [29][30][31][32]. To ensure the safety, efficacy, and usefulness of biomaterials in medical applications, this procedure is critical. ...
Article
Full-text available
The need to improve the expectancy and quality of life of subjects affected by disabling pathologies that require the replacement or regeneration of tissues or parts of the body has fueled the development of innovative, better-performing materials that are capable of integrating into and being tolerated by body tissues. Materials with these characteristics, i.e., bio-functionality, bio-safety, and biocompatibility, are defined as biomaterials. One of the many methods for producing such materials is the sol–gel technique. This process is mainly used for the preparation of ceramic oxides at low temperatures, through hydrolysis and polycondensation reactions of organometallic compounds within a hydroalcoholic solution. This study is based on a specific type of biomaterial: organic–inorganic hybrids. The aim of this study is to provide an overview of the advantages and disadvantages of the sol–gel technique, as well as describe the preparation and chemical and biological characterization, uses, and future prospects of these biomaterials. In particular, the use of plant drugs as organic components of the hybrid material is the innovation of this manuscript. The biological properties of plant extracts are numerous, and for this reason, they deserve great attention from the scientific community.
... Previous Shigometers were only able to measure impedance of DC currents, but since tree tissue can be modeled as a combination of resistance and capacitance, the TPAM system requires the ability to change the frequency of the applied current. To measure impedance, a probe is inserted into the tree, and a contact impedance is formed between the metal probe and the tree tissue, as shown in Figure 3b [13]. The symbol CPA represents the interface capacitance, the symbol Rct represents the charge transfer resistance, and the symbol RS represents the solution resistance. ...
Article
Full-text available
A system has been developed to remotely, continuously, and quantitatively measure the physiological activity of trees. The developed tree physiological activity monitoring (TPAM) system is equipped with electrical impedance, temperature, and light intensity measurement functions. In the two-contact impedance measurement method used in the previous plant impedance measurement, errors due to the polarization impedance of the electrodes could not be avoided. The developed TPAM system adopted a four-contact measurement method that could avoid polarization impedance errors, and, with it, the long-term monitoring of zelkova trees was performed. The monitoring of seasonal changes was conducted from July to November, and an impedance change pattern that repeated on a daily basis was observed in the short term, and an overall increase in the impedance was observed in the long term. Impedance changes related to daily temperature changes were observed even after all the tree leaves had fallen, meaning that this effect should be excluded when using impedance to evaluate tree vitality. For this reason, the influence of temperature fluctuations was excluded by using only the impedance values at the same daily temperature of 25 degrees from July to November. The analysis results at 25 degrees showed that the tree impedance value increased linearly by 8.7 Ω per day. The results of this series of long-term monitoring and analysis revealed that the ambient temperature must be taken into account in the evaluation of tree physiological activity based on electrical impedance.
... However, when the magnitude of the applied voltage or current increased beyond those limits, the charge transfer mechanism depended on the activation overpotential generated by the applied current or voltage. A sensitivity analysis of the influence of the exchange current density, confirmed that the polarization curves for both current and voltage-controlled stimulation were relatively insensitive to the value of the exchange current density within the experimentally reported range for platinum and Pt-Ir [14,16,17,[36][37][38], figures 2(C) and (D). This is consistent with Richardot et al who showed that variation in the exchange current density was not the main source of nonlinearity at the electrode but acts as a multiplying coefficient rather than changing the current potential characteristics [16]. ...
Article
Full-text available
Objective: The electrode-tissue interface provides the critical path for charge transfer in neurostimulation therapies and exhibits well-established nonlinear properties at high applied currents or voltages. These nonlinear properties may influence efficacy and safety of applied stimulation but are typically neglected in computational models. In this study, nonlinear behaviour of the electrode-tissue interface was incorporated in a computational model of deep brain stimulation (DBS) to simulate the impact on neural activation and safety. Approach: Nonlinear electrode-tissue interface properties were incorporated in a finite element model of DBS electrodes in vitro and in vivo, in the rat subthalamic nucleus, using an iterative approach. The transition point from linear to non-linear behaviour was determined for voltage and current-controlled stimulation. Predicted levels of neural activation during DBS were examined and the region of linear operation was compared with the Shannon safety limit. Main results: A clear transition of the electrode-tissue interface to the nonlinear region was observed for both current and voltage-controlled stimulation. The transition occurred at lower values of the activation overpotential for simulated in vivo than in vitro conditions (91mV and 165mV current-control; 110mV and 275mV voltage-control), corresponding to applied currents of 30µA and 45µA, voltage of 330mV. Incorporation of nonlinear properties resulted in activation of a higher proportion of neurons under voltage- but not current-controlled stimulation. Under current-controlled stimulation, the transition to Faradaic charge transfer occurred at stimulation amplitudes as low as 30µA, which corresponds to a charge density of 2.29µC/cm2 and charge of 1.8nC, well below the Shannon safety limit. Significance: The results indicate that DBS electrodes may operate within the nonlinear region at clinically-relevant stimulation amplitudes. This affects the extent of neural activation under voltage-controlled stimulation and the transfer to Faradaic charge transfer for both voltage- and current-controlled stimulation with important implications for targeting of neural populations and safety.
Article
This paper presents an ultrahigh input impedance, low noise, and wide bandwidth four‐channels analog front‐end (AFE) for low‐power neural recording systems. To achieve high input impedance, the buffer channels are placed between the electrodes and the main amplifier stage of the AFE. The buffer is designed with low noise and low power consumption to obtain high input impedance of overall AFE while maintaining low input‐referred noise and low power consumption. A chopper capacitively coupled chopper instrumentation amplifier (CCIA) is placed after the buffer as the main amplifier stage of the AFE to improve the common mode rejection ratio (CMRR) and the input‐referred noise of the overall AFE design. A new chopper stabilization control technique is proposed and used in the CCIA stage to reduce the charge injection and clock feedthrough and consequently the high‐frequency ripple of the AFE output signal. A programmable gain amplifier (PGA) is designed as the third stage to adjust the overall gain of the AFE. Benefiting from PGA, the AFE can adapt its gain with both action potential and local field potential signals. To reduce the number of the electrode to be implanted and reduce the impedance mismatch that causes the degradation on overall CMRR performance, two AFE channels shared a reference electrode followed by a reference buffer. The proposed AFE is designed and simulated using a standard 180 nm CMOS process and operates in a wide frequency band of 2.1 to 2500 Hz with low input‐referred noise of 1.6 μVrms and a CMRR over 80 dB at 2.1 Hz. The total power consumption is lower than 4.3 μW per channel. With the proposed structure of AFE, the input impedance is 35 GΩ @ 21 Hz and the minimum impedance over operational bandwidth is 75 MΩ @ 2.5 kHz.
Article
Drug-induced cardiotoxicity is a significant contributor to drug recalls, primarily attributed to limitations in existing drug screening platforms. Traditional heart-on-a-chip platforms often employ metallic electrodes to record cardiomyocyte electrical signals....
Article
Full-text available
A review of various aspects of electrode-electrolyte interface impedance is presented. The effect of electrode topography on the form and magnitude of the interface impedance is discussed. The work of Schwan and his colleagues on the non-linearity of the interface impedance is presented and interpreted. The electrical properties of silver-silver chloride electrodes (much used in a wide range of biomedical applications) are also briefly reviewed.
Article
A model for the impedance of granular metal hydride electrodes is presented. The model includes the rate of hydrogen exchange at the hydride–electrolyte interface and diffusion of hydrogen into the bulk of the hydride. The exchange reaction is assumed to occur in two steps: electrosorption of hydrogen at the interface and hydride formation. The subsequent transport process is approximated as spherical diffusion in the particles. The model was successfully fitted to experimental data for a misch metal/nickel based hydride forming compound (MmNi3.5–3.7Co0.7–0.8Mn0.3–0.4Al0.3–0.4) at different depths of discharge.
Article
The rates of a number of electron exchange reactions at mercury and other noble metal electrodes have been measured. Some interpretation of the rates of such reactions in terms of mechanism has been made; structural change in the reactant accompanying the electron transfer, and any hindrance to the close approach of the reactant ions to the electrode surface, appear to be the most important factors limiting the rates of these reactions.
Article
A microelectrode array (MEA) consisting of 34 silicon nitride passivated Pt-tip microelectrodes embedded on a perforated silicon substrate (porosity 35%) has been realized. The electrodes are 47 microns high, of which only the top 15 microns are exposed Pt-tips having a curvature of 0.5 micron. The MEA is intended for extracellular recordings of brain slices in vitro. Here we report the fabrication, characterization and initial electrophysiological evaluation of the first generation of Pt-tip MEAs.