You are on page 1of 7

Available online at www.sciencedirect.

com

Electrochimica Acta 53 (2008) 7360–7366

An integrated approach to electrochemical impedance spectroscopy


Mark E. Orazem a,∗ , Bernard Tribollet b
a
Department of Chemical Engineering, University of Florida, Gainesville, FL 32611, USA
b UPR15 du CNRS, Laboratoire Interfaces et Systèmes Electrochimiques, Université P. et M. Curie,
CP 133, 4 Place Jussieu, 75252 Paris, France
Received 20 August 2007; received in revised form 25 October 2007; accepted 31 October 2007
Available online 21 December 2007

Abstract
A philosophy for electrochemical impedance spectroscopy is presented which integrates experimental observation, model development, and error
analysis. This approach is differentiated from the usual sequential model development for given impedance spectra by its emphasis on obtaining
supporting observations to guide model selection, use of error analysis to guide regression strategies and experimental design, and use of models to
guide selection of new experiments. These concepts are illustrated with two examples taken from the literature. This work illustrates that selection
of models, even those based on physical principles, requires both error analysis and additional experimental verification.
© 2007 Elsevier Ltd. All rights reserved.

Keywords: Impedance spectroscopy; Modeling; Error analysis

1. Introduction or a comparatively empirical equivalent circuit. The parame-


ters for either model can be estimated by complex nonlinear
While impedance spectroscopy can be a sensitive tool least squares regression. The authors observed that ideal electri-
for analysis of electrochemical and electronic systems, an cal circuit elements represent ideal lumped constant properties;
unambiguous interpretation of spectra cannot be obtained by whereas, the physical properties of electrolytic cells are often
examination of raw data. Instead, interpretation of spectra distributed. The distribution of cell properties motivates use of
requires development of a model which accounts for the distributed impedance elements such as constant-phase elements
impedance response in terms of the desired physical properties. (CPE). An additional problem with equivalent circuit analysis,
Model development should take into account both the impedance which the authors suggest is not shared by the direct compari-
measurement and the physical and chemical characteristics of son to the theoretical model, is that circuit models are ambiguous
the system under study. and different models may provide equivalent fits to a given spec-
It is useful to envision a flow diagram for the measure- trum. The authors suggest that identification of the appropriate
ment and interpretation of experimental measurements such as equivalent circuit can be achieved only by employing physical
impedance spectroscopy. Barsoukov and Macdonald proposed intuition and by carrying out several sets of measurements with
such a flow diagram for a general characterization procedure different conditions.
(Figure 1.2.1 in both references [1] and [2]) consisting of two A similar flow diagram was presented by Huang et al. [3]
blocks comprising the impedance measurement, three blocks for solid-oxide fuel cells (SOFCs). The diagram accounts for
comprising a physical (or process) model, one block for an the actions of measuring impedance data, modeling, fitting the
equivalent electrical circuit, and blocks labeled curve fitting model, interpreting the results, and optimizing the fuel cell for
and system characterization. They suggested that impedance power generation. The authors emphasize that the interpreting
data may be analyzed for a given system by using either an action depends more on the electrochemical expertise of the
exact mathematical model based on a plausible physical theory researchers than on a direct mapping from model parameters to
SOFC properties.
While helpful, the flow diagrams proposed by Barsoukov and
∗ Corresponding author. Tel.: +1 352 392 6207; fax: +1 352 392 9513. Macdonald [2] and Huang et al. [3] are incomplete because they
E-mail address: meo@che.ufl.edu (M.E. Orazem). do not account for the role of independent assessment of exper-

0013-4686/$ – see front matter © 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2007.10.075
M.E. Orazem, B. Tribollet / Electrochimica Acta 53 (2008) 7360–7366 7361

imental error structure and because they do not emphasize the The impedance measurements should also be conducted in
critical role of supporting experimental measurements. Orazem concert with error analysis with emphasis on both stochas-
et al. proposed a flow diagram (Figure 6 in reference [4]) consist- tic and systematic bias errors. The bias errors can be defined
ing of three elements: experiment, measurement model, and pro- to be those that result in data that are inconsistent with the
cess model. The measurement model was intended to assess the Kramers–Kronig relations. An empirical error analysis can
stochastic and bias error structure of the data; thus, their diagram be undertaken using the measurement model approach sug-
accounts for the independent assessment of experimental error gested by Agarwal et al. [5–7] and Orazem [8]. It should be
structure. Their diagram does not, however, account for sup- noted, however, that this approach is not definitive because
porting non-impedance measurements, and the use of regression Kramers–Kronig-consistent artifacts can be caused by electrical
analysis, while implied, is not shown explicitly. The object of leads and the electronics. Use of dummy cells can help identify
this work is to formulate a comprehensive approach which can be artifacts that are consistent with the Kramers–Kronig rela-
applied to measurement and interpretation of impedance spectra. tions. As an alternative approach for identification of stochastic
part of the error structure, Dygas and Breiter have shown
2. Integration of measurements, error analysis, and that impedance instrumentation could, in principle, provide
model standard deviations for the impedance measurements at each
frequency [9].
A refined philosophical approach toward the use of The feedback loop shown in Fig. 1 between EIS Experi-
impedance spectroscopy is outlined in Fig. 1 where the trian- ment and Error Analysis indicates that the error structure is
gle evokes the concept of an operational amplifier for which obtained from the measured data and that knowledge of the
the potential of input channels must be equal. Sequential steps error structure can guide improvements to the experimental
are taken until the model provides a statistically adequate rep- design. The magnitude of perturbation, for example, should be
resentation of the data to within the independently obtained selected to minimize stochastic errors while avoiding induc-
stochastic error structure. The different aspects which comprise ing a nonlinear response. The frequency range should be
the philosophy are presented in this section. selected to sample the system time constants, while avoiding
bias errors associated with non-stationary phenomena. In short,
2.1. Impedance measurement are integrated with the experimental parameters should be selected so as to min-
experimental error analysis imize the stochastic error structure while, at the same time,
allowing for the maximum frequency range that is free of bias
All impedance measurements should begin with measure- errors.
ment of a steady-state polarization curve. The steady-state
polarization curve is used to guide selection of an appropri- 2.2. Process model is developed in concert with other
ate perturbation amplitude and can provide initial hypotheses observations
for model development. The impedance measurements can
then make be performed at selected points on the polarization The model identified in Fig. 1 represents a process model
curve to explore the potential dependence of reaction rate con- intended to account for the hypothesized physical and chemi-
stants. Impedance measurements can be performed as well as cal character of the system under study. From the perspective
a function of other state variables such as temperature, rotation embodied in Fig. 1, the objective of the model is not to
speed, reactant concentration. Impedance scans made at differ- provide a good fit with the smallest number of parameters.
ent points of time can be used to explore temporal changes in The objective is rather to use the model to gain physical
system parameters. Some examples include growth of oxide or understanding of the system. The model should be able to
corrosion-product. account for, or at least be consistent with, all experimental
observations. The supporting measurements therefore provide
a means for model identification. The feedback loop shown in
Fig. 1 between Model and Other Observations is intended to
illustrate that the supporting measurements guide model devel-
opment and the proposed model can suggest experiments needed
to validate model hypotheses. The supporting experiments
can include both electrochemical and non-electrochemical
measurements.
Numerous scanning electrochemical methods such as scan-
ning reference electrodes, scanning tunneling microscopy, and
scanning electrochemical microscopy can be used to explore
surface heterogeneity. Scanning vibrating electrodes and probes
can be used to measure local current distributions. Local
Fig. 1. Schematic flowchart showing the relationship between impedance mea- electrochemical impedance spectroscopy provides a means of
surements, error analysis, supporting observations, model development, and exploring the distribution of surface reactivity. Measurements
weighted regression analysis. can be performed across the electrode at a single frequency to
7362 M.E. Orazem, B. Tribollet / Electrochimica Acta 53 (2008) 7360–7366

create an image of the electrode or, alternatively, performed at


a given location to create a complete spectrum. Other experi-
ments may include in situ and ex situ surface analysis, chemical
analysis of electrolytes, and both in situ and ex situ visualization
and/or microscopy. Transfer-function methods such as electro-
hydrodynamic (EHD) impedance spectroscopy allow isolation
of the phenomena that influence the electrochemical impedance
response [10].

2.3. Regression analysis accounts for error structure

The goal of the operation represented by the triangle in Fig. 1


is to develop a model that provides a good representation of the
impedance measurements to within the noise level of the mea-
surement. The error structure for the measurement clearly plays
a critical role in the regression analysis. The weighting strategy
for the complex nonlinear least squares regression should be
based on the variance of the stochastic errors in the data, and the
frequency range used for the regression should be that which has
been determined to be free of bias errors. In addition, knowledge
of the variance of the stochastic measurement errors is essential
to quantify the quality of the regression.
Sequential steps are taken until the model provides a sta-
tistically adequate representation of the data to within the
independently obtained stochastic error structure. The compari-
son between model and experiment can motivate modifications
to the model or to the experimental parameters.

3. Examples

The systems described below illustrate the approach outlined Fig. 2. Impedance data as a function of frequency for an n-GaAs/Ti Schottky
in Fig. 1. diode with temperature as a parameter: (a) the real part of the impedance; and
(b) the imaginary part of the impedance. Data taken from Orazem et al. [11] and
Jansen et al. [12].
3.1. Deep-level states in GaAs diodes
a normalized frequency defined by
Orazem et al. [11] and Jansen et al. [12] described the
 
impedance response for an n-type GaAs Schottky diode with f E
temperature as a parameter. The system consisted of an n-GaAs f ∗ = ◦ exp (1)
f kT
single crystal with a Ti Schottky contact at one end and a Au,
Ge, Ni Schottky contact at the eutectic composition at the other where E = 0.827 eV, and the characteristic frequency f ◦ was
end. This material has been well characterized in the litera- assigned a value of 2.964 × 1014 Hz such that the imaginary part
ture and, in particular, has a well known EL2 deep-level state of the normalized impedance values reached a peak value near
that lies 0.83–0.85 eV below the conduction band edge [13]. f ∗ = 1. The data collected at different temperatures are reduced
Experimental details are provided elsewhere [11,12].
The experimental data are presented in Fig. 2 (a) and (b)
for the real and imaginary parts of the impedance, respectively,
for temperatures ranging from 320 to 400 K. The logarithmic
scale used in Fig. 2 emphasizes the scatter seen in the imag-
inary impedance at low frequencies. The impedance response
is seen to be a strong function of temperature. The impedance-
plane plots shown in Fig. 3 for data collected at 320 and 340 K
approach the classic semicircle associated with a single relax-
ation process.
Jansen and Orazem showed that the impedance data could be
superposed as given in Fig. 4[12]. The impedance data collected Fig. 3. Impedance data in impedance-plane format for an n-GaAs/Ti Schottky
at different temperatures were normalized by the maximum diode with temperature as a parameter. Data taken from Orazem et al. [11] and
mean value of the real part of the impedance and plotted against Jansen et al. [12]
M.E. Orazem, B. Tribollet / Electrochimica Acta 53 (2008) 7360–7366 7363

Fig. 4. Impedance data from Fig. 2 collected for an n-GaAs/Ti Schottky diode Fig. 5. Impedance data from Fig. 2 collected for an n-GaAs/Ti Schottky diode
as a function of frequency f ∗ = (f/f ◦ ) exp(E/kT ): (a) the real part of the as a function of frequency f ∗ = (f/f ◦ ) exp(E/kT ): (a) the real part of the
impedance; and (b) the imaginary part of the impedance. Data taken from Jansen impedance; and (b) the imaginary part of the impedance.
et al. [12]

electronic transitions [11,12]. The information concerning these


to a single line. The extent to which the data are superposed is transitions could be extracted by regression of an appropriate
seen more clearly on the logarithmic scale shown in Fig. 5. The process model using a weighting strategy based on the error
superposition shown in Figs. 4 and 5 suggest that the system is structure of the measurement.
controlled by a single-activation-energy-controlled process. Two models have been proposed for the data presented above.
A closer examination of Fig. 4(b) reveals that the maxi- Macdonald proposed a distributed-time-constant model which
mum magnitude of the scaled imaginary impedance is slightly accounts for distributed relaxation processes [14]. This model
less than 0.5; whereas the corresponding value for a single- fits the data very well and has the advantage that it requires
activation-energy-controlled process should be identically 0.5. a minimal number of parameters. A second model, presented
Regression analysis using the traditional weighting strategies in Fig. 6, accounts for discrete energy levels and provides a
under which the standard deviation of the experimental values fit of equivalent regression quality at each temperature [11,12].
was assumed to be proportional to the modulus of the impedance, In Fig. 6, Cn is the space-charge capacitance, Rn is a resis-
σr = σj = α|Z|, to the magnitude of the respective components tance that accounts for a small but finite leakage current, and
of the impedance, σr = αr |Zr | and σj = αj |Zj |, or independent the parameters R1 . . . Rk and C1 . . . Ck are attributed to the
of frequency, σr = σj = α, yielded one dominant RC time con- response of discrete deep-level energy states. Parameters cor-
stant with only a hint that other parameters could be extracted. responding to deep-level states were added sequentially to the
Orazem et al. [11] and Jansen et al. [12] used the measurement model, subject to the constraint that the 2σ (95.4%) confi-
model approach described by Agarwal et al. [5–7] and Orazem dence interval for each of the regressed parameters may not
[8] to identify the stochastic error structure for the impedance include zero. Including the space charge capacitance and leak-
data. When the data were regressed using this error structure for age resistance, four resistor–capacitor pairs could be obtained
a weighting strategy, additional parameters could be resolved from the impedance data collected at 300, 320, and 340 K; three
revealing additional activation energies. Thus, while the data resistor–11/23/2007capacitor pairs could be obtained from the
do superpose nicely in Figs. 4 and 5, the impedance data do in data collected at 360, 380, and 400 K; and two resistor-capacitor
fact contain information on minor activation-energy-controlled pairs could be obtained from the data collected at 420 K [12].
7364 M.E. Orazem, B. Tribollet / Electrochimica Acta 53 (2008) 7360–7366

Fig. 7. Electrochemical impedance results obtained for a single-cell PEM fuel


cell with current density as a parameter. Results taken from Roy et al. [18].

reactants [18]. The measurements were conducted in galavanos-


tatic mode for a frequency range of 1 kHz to 1 mHz with a 10 mA
Fig. 6. Electrical circuit corresponding to the model presented by Jansen et al. peak-to-peak sinusoidal perturbation. Roy and Orazem [19] used
[12] in which Cn is the space-charge capacitance, Rn is a resistance that accounts the measurement model approach developed by Orazem and
for a small but finite leakage current, and the parameters R1 . . . Rk and C1 . . . Ck coworkers [5–8] to demonstrate that, for the fuel cell under
are attributed to the response of discrete deep-level energy states. steady-state operation, the low-frequency inductive loops seen
in Fig. 7 were consistent with the Kramers–Kronig relations.
This model has the disadvantage that up to eight parameters Therefore, the low-frequency inductive loops could be attributed
are required, depending on the temperature, as compared to to process characteristics and not to non-stationary artifacts.
the three parameters required for the distributed-time-constant Roy et al. [18] proposed that that the low-frequency inductive
model. The question to be posed then, “which is the better model loops observed in PEM fuel cells could be caused by parasitic
for the measurements?” reactions in which the Pt catalyst reacts to form PtO and sub-
If the goal of the regression is to provide the most parsimo- sequently forms Pt+ ions. They also showed that a reaction
nious model for the data, the model with the smallest number of involving formation of hydrogen peroxide could yield the same
parameters and a continuous distribution of activation energies inductive features. A comparison between the two models and
is the best model. If the goal of the regression is to provide a the experimental results is shown in Fig. 8. and the correspond-
quantitative physical description of the system for which the data ing values for the real and imaginary parts of the impedance are
were obtained, additional measurements are needed to determine presented as a function of frequency in Fig. 9(a) and (b), respec-
whether the activation energies are discrete or continuously dis- tively. The model calculations were not obtained by regression
tributed. In this case, deep-level transient spectroscopy (DLTS) but rather by simulation using approximate parameter values.
measurements indicated that the n-GaAs diode contained dis- Regression was not used because the model was based on the
crete deep-level states. In addition, the energy levels and state assumption of a uniform membrane-electrode assembly (MEA);
concentrations obtained by regression of the second model were whereas, the use of a serpentine flow channel caused the reactiv-
consistent with the results obtained by DLTS [12]. Thus, the sec- ity of the MEA to be very nonuniform. The parameters were first
ond model with a larger number of parameters provides the more selected to reproduce the current–potential curve and then the
useful description of the GaAs diode. The choice between the same parameters were used to calculate the impedance response
two models could not be made without the added experimental
evidence.

3.2. Side reactions in PEM fuel cells

Low-frequency inductive features [15–17] are commonly


seen in impedance spectra for PEM fuel cells. Makharia et al.
[15] suggested that side reactions and intermediates involved
in the fuel cell operation can be possible causes of the induc-
tive loop seen at low frequency. However, such low-frequency
inductive loops could also be attributed to non-stationary behav-
ior, or, due to the time required to make measurements at low
frequencies, non-stationary behavior could influence the shapes
Fig. 8. Comparison of the impedance response for a PEM fuel cell operated at
of the low-frequency features. 0.2 A/cm2 to model predictions generated using a reaction sequence involving
A typical result is presented in Fig. 7 for the impedance formation of hydrogen peroxide and a reaction sequence involving formation of
response of a single 5 cm2 PEMFC with hydrogen and air as PtO. Results taken from Roy et al. [18]
M.E. Orazem, B. Tribollet / Electrochimica Acta 53 (2008) 7360–7366 7365

with a decrease in the active catalytic surface area and a loss of


Pt ions in the effluent. Cyclic voltammetry after different peri-
ods of fuel cell operation could be used to explore reduction
in electrochemically active area. Inductively coupled plasma
mass spectroscopy (ICPMS) could be used to detect residual
Pt ions in the fuel cell effluent, and ex situ techniques could
detect formation of PtO in the catalyst layer. A different set of
experiments could be performed to explore the hypothesis that
peroxide formation is responsible for the inductive loops. Plat-
inum dissolution has been observed in PEM fuel cells, [21] and
peroxide formation has been implicated in the degradation of
PEM membranes [22–24]. Thus, it is likely that both reactions
are taking place and contributing to the observed low-frequency
inductive loops.

4. Discussion

The example presented in Section 3.1 demonstrates the


importance of coupling experimental observation, model devel-
opment, and error analysis. The measurements conducted
at different temperatures allowed identification of discrete
activation energies for electronic transitions. Use of a weight-
ing strategy based on the observed stochastic error structure
increased the number of parameters that could be obtained from
the regression analysis. Thus, four discrete activation-energy-
controlled processes could be identified, but at the expense of a
corresponding model that required eight parameters. A regres-
sion of almost the same quality could be obtained under the
assumption of a continuous distribution of activation energies,
and this model required only three parameters. Discrimination
Fig. 9. Comparison of the impedance response for a PEM fuel cell operated at
0.2 A/cm2 to model predictions generated using a reaction sequence involving
between the two models requires additional experimental obser-
formation of hydrogen peroxide and a reaction sequence involving formation of vations, such as the DLTS identification of electronic transitions
PtO: (a) real part of the impedance; and (b) imaginary part of the impedance. involving discrete deep-level states.
Results taken from Roy et al. [18]. The example presented in Section 3.2 demonstrates the
utility of the error analysis for determining consistency
for each value of current density. The potential (or current) with the Kramers–Kronig relations. In this case, the low-
dependence of model parameters was that associated with the frequency inductive loops were found to be consistent with the
Tafel behavior assumed for the electrochemical reactions. Kramers–Kronig relations at frequencies as low as 0.001 Hz so
While reaction parameters were not identified by regres- long as the system had reached a steady-state operation. The
sion to impedance data, the simulation presented by Roy et error analysis employed regression of a measurement model.
al. [18] demonstrates that side reactions proposed in the liter- The mathematical process models that were proposed to account
ature can account for low-frequency inductive loops. Indeed, for the low-frequency features were based on plausible physical
the results presented in Figs. 8 and 9 show that both models and chemical hypotheses. Nevertheless, the models are ambigu-
can account for low-frequency inductive loops. Other models ous and require additional measurements and observations to
can also account for low-frequency inductive loops so long as identify the most appropriate for the system under study.
they involve potential-dependent adsorbed intermediates [20]. The philosophy described here cannot always be followed
It is generally understood that equivalent circuit models are not to convergence. Often the hypothesized model is inadequate
unique and have therefore an ambiguous relationship to phys- and cannot reproduce the experimental results. Even if the
ical properties of the electrochemical cell. As shown by Roy proposed reaction sequence is correct, surface heterogeneities
et al. [18], even models based on physical and chemical pro- may introduce complications that are difficult to model. The
cesses are ambiguous. In the present case, the ambiguity arises accessible frequency range may be limited at high frequency for
from uncertainty as to which reactions are responsible for the systems with a very small impedance. The accessible frequency
low-frequency inductive features. range may be limited at low frequency for systems subject
Resolution of this ambiguity requires additional experiments. to significant non-stationary behavior. The experimentalist
The processes and reactions hypothesized for a given model can may need to accept a large level of stochastic noise for a
suggest experiments to support or reject the underlying hypoth- system with a large impedance. In cases where the models are
esis. For example, the proposed formation of PtO is consistent unable to explain all features of the experiment, the graphical
7366 M.E. Orazem, B. Tribollet / Electrochimica Acta 53 (2008) 7360–7366

methods presented by Orazem et al. [25] can nevertheless yield [10] C. Gabrielli, B. Tribollet, A transfer-function approach for a generalized
quantitative information. electrochemical impedance spectroscopy, J. Electrochem. Soc. 141 (1994)
1147.
[11] M.E. Orazem, P. Agarwal, A.N. Jansen, P.T. Wojcik, L.H. Garcı́a-Rubio,
5. Conclusions Development of physico-chemical models for electrochemical impedance
spectroscopy, Electrochim. Acta 38 (1993) 1903.
The philosophy embodied in Fig. 1 integrates experimental [12] A.N. Jansen, P.T. Wojcik, P. Agarwal, M.E. Orazem, Thermally-stimulated
observation, model development, and error analysis. It takes into deep-level impedance spectroscopy: application to an n-GaAs Schottky
account the observations that impedance spectroscopy is not a diode, J. Electrochem. Soc. 143 (1996) 4066.
[13] G.M. Martin, S. Makram-Ebeid, The mid–gap donor level el2 in GaAs, in:
stand-alone technique and that other observations are required to S.T. Pantelides (Ed.), Deep Centers in Semiconductors, Gordon and Breach
validate a given interpretation of the impedance spectra. Within Science Publishers, New York, 1986, p. 455.
the context of the work presented here, the objective of modeling [14] J.R. Macdonald, Power-law exponents and hidden bulk relaxation in
is not to provide a good fit with the smallest number of parame- the impedance spectroscopy of solids, J. Electroanal. Chem. 378 (1994)
ters, but, rather, to use the model to learn about the physics and 17.
[15] R. Makharia, M.F. Mathias, D.R. Baker, Measurement of catalyst layer
chemistry of the system under study. electrolyte resistance in PEFCs using electrochemical impedance spec-
troscopy, J. Electrochem. Soc. 152 (5) (2005) A970.
References [16] O. Antoine, Y. Bultel, R. Durand, Oxygen reduction reaction kinetics and
mechanism on platinum nanoparticles inside Nafion, J. Electroanal. Chem.
[1] J.R. Macdonald (Ed.), Impedance Spectroscopy: Emphasizing Solid Mate- 499 (2001) 85.
rials and Systems, John Wiley & Sons, New York, 1987. [17] Y. Bultel, L. Genies, O. Antoine, P. Ozil, R. Durand, Modeling impedance
[2] E. Barsoukov, J.R. Macdonald (Eds.), Impedance Sprectroscopy Theory, diagrams of active layers in gas diffusion electrodes: diffusion, ohmic
Experiment, and Applications, 2nd ed., Wiley-Interscience, New York, drop effects and multistep reactions, J. Electroanal. Chem. 527 (2002)
2005. 143.
[3] Q.-A. Huang, R. Hui, B. Wang, J. Zhang, A review of ac impedance mod- [18] S.K. Roy, M.E. Orazem, B. Tribollet, Interpretation of low-frequency
eling and validation in SOFC diagnosis, Electrochim. Acta (2007) 8144. inductive loops in PEM fuel cells, J. Electrochem. Soc. 154 (2007) B1378.
[4] M.E. Orazem, P. Agarwal, L.H. Garcı́a-Rubio, Critical issues associated [19] S.K. Roy, M.E. Orazem, Error analysis for the impedance response of PEM
with interpretation of impedance spectra, J. Electroanal. Chem. Interfacial fuel cells, J. Electrochem. Soc. 154 (8) (2007) B883.
Electrochem. 378 (1994) 51. [20] I. Epelboin, C. Gabrielli, M. Keddam, H. Takenouti, A model of the anodic
[5] P. Agarwal, M.E. Orazem, L.H. Garcı́a-Rubio, Measurement models for behavior of iron in sulphuric acid medium, Electrochim. Acta 20 (1975)
electrochemical impedance spectroscopy: I. Demonstration of applicabil- 913.
ity, J. Electrochem. Soc. 139 (7) (1992) 1917. [21] R.M. Darling, J.P. Meyers, Kinetic model of platinum dissolution in PEM-
[6] P. Agarwal, O.D. Crisalle, M.E. Orazem, L.H. Garcı́a-Rubio, Measurement FCs, J. Electrochem. Soc. 150 (2003) A1523.
models for electrochemical impedance spectroscopy: 2. Determination of [22] C.F. Zinola, J. Rodriguez, G. Obal, Kinetic of molecular oxygen electrore-
the stochastic contribution to the error structure, J. Electrochem. Soc. duction on platinum modified by tin underpotential deposition, J. Appl.
(1995) 4149. Electrochem. 31 (2001) 1293.
[7] P. Agarwal, M.E. Orazem, L.H. Garcı́a-Rubio, Measurement models for [23] A. Damjanovic, V. Brusic, Electrode kinetic of oxygen reduction on oxide-
electrochemical impedance spectroscopy: 3. Evaluation of consistency with free Pt electrode, Electochim. Acta 12 (1967) 615.
the kramers-kronig relations, J. Electrochem. Soc. 142 (1995) 4159. [24] V.O. Mittal, H.R. Kunz, J.M. Fenton, Is H2 O2 involved in the membrane
[8] M.E. Orazem, A systematic approach toward error structure identification degradation mechanism in pemfc, Electrochem. Solid State Lett. 9 (6)
for impedance spectroscopy, J. Electroanal. Chem. 572 (2004) 317. (2006) A229.
[9] J.R. Dygas, M.W. Breiter, Measurements of large impedances in a wide [25] M.E. Orazem, N. Pébère, B. Tribollet, A new look at graphical representa-
temperature and frequency range, Electrochim. Acta 41 (7/8) (1996) 993. tion of impedance data, J. Electrochem. Soc. 153 (2006) B129.

You might also like