You are on page 1of 22

ELEC-8900 Special Topics:

Advanced Energy Storage Systems


Lecture 08: Electrochemical Impedance
Spectroscopy (EIS)

Instructor: Dr. Balakumar Balasingam


July 25, 2022

Contents
1 Introduction 2

2 Frequency Response of a Battery 2

3 Computing Frequency Response Using DFT 3

4 ECM Parameter Estimation Problem 6

5 Approximate Estimation of ECM Parameters 6


5.1 Effect of Approximation . . . . . . . . . . . . . . . . . . . . . . . 8
5.2 Effect of Measurement Noise . . . . . . . . . . . . . . . . . . . . 8

6 Improved Approach for Parameter Estimation 8


6.1 Estimation of Warburg Coefficient . . . . . . . . . . . . . . . . . 10
6.2 Estimation of CT Components . . . . . . . . . . . . . . . . . . . 11
6.3 Estimation of SEI Components . . . . . . . . . . . . . . . . . . . 13
6.4 Estimation of Resistance and Inductance . . . . . . . . . . . . . . 14
6.5 Feature Point Extraction . . . . . . . . . . . . . . . . . . . . . . . 15

7 Demonstration 15
7.1 Demonstration Using Simulated Data . . . . . . . . . . . . . . . 15
7.2 Demonstration Using Real Data . . . . . . . . . . . . . . . . . . . 16

8 Summary 20

1
1 Introduction
Electrochemical impedance spectroscopy (EIS) is a widely used approach for
battery analysis. In EIS, the frequency response of the battery, known as the
Nyquist plot, is studied for insights about battery health. Numerous works in
the literature have employed Nyquist plots to understand and visualize battery
aging. This chapter provides insights about the relationship between EIS and
the ECM parameters of the battery; it is shown how to estimate the ECM
parameters once the Nyquist plot is obtained. The estimated ECM parameters
can then be related to SOH using the ideas provided in Chapter ??.
This chapter also illustrates how the parameter estimation becomes challeng-
ing with measurement noise in the sensors. Measurement noise is a significant
problem in practical systems that are preferred in general to be low-cost; it is
also possible that the measurement noise increases as the system ages. With
measurement noise, the uncertainty about the (estimated) ECM parameters
increases. It is important to have an understanding about the performance
of ECM parameter estimation at various noise levels. This chapter introduces
performance analysis of ECM parameter estimation, based on the Nyquist spec-
trum, at various signal to noise ratio (SNR) levels.
Despite numerous literature that employed EIS to analyze battery charac-
teristics, there is only a sparse work in literature that enlightens the signal
processing aspects and challenges involved in battery analysis. The present
chapter tries to paint a signal processing view of EIS analysis for battery engi-
neers by highlighting the ECM parameter estimation aspects and by presenting
performance analysis at various SNR levels.

2 Frequency Response of a Battery


Randles circuit models are one of the widely used ECM in battery analysis.
Figure 1(a) shows a diagram of the adaptive Randles equivalent circuit model.
This model consists of the following elements:
• Voltage source, Ecell
• Stray inductance, L
• Ohmic resistance, RΩ

• Solid electrolyte interface (SEI) resistance, RSEI


• SEI capacitance, CSEI
• Charge transfer(CT) resistance, RCT

• Double layer (DL) capacitance, CDL


• Warburg impedance, Zw

2
where the Warburg impedance defined as
σ
Zw (jω) = (1 − j) √ (1)
ω

where σ is referred in this chapter as the Warburg coefficient. The Warburg


impedance describes a phenomenon observed at very low frequencies. It can
be noticed that as the frequency decreases, the Warburg impedance increases.
When there is a significant frequency in the system, which is the case in practical
applications, the Warburg impedance effectively becomes zero. On the other
hand, at low frequencies, only the effects of resistors and the Warburg element
remain.
In EIS, an AC perturbation signal (current or voltage) is applied to a battery
and its response (voltage or current) is recorded; by jointly analyzing the applied
signal and its response, the parameters of the battery ECM can be estimated.
Figure 1(b) shows the typical response of a battery in the frequency domain —
this impedance plot is generally known as the Nyquist plot.
Using the AR-ECM shown in Figure 1, the AC impedance Z(ω) can be
written as

Z(ω) ,Z(jω)
1 1
=jωL + RΩ + 1 + 1
RSEI + jωCSEI RCT +Zw (jω) + jωCDL
RSEI RCT + Zw (jω)
=jωL + RΩ + + (2)
1 + jωRSEI CSEI 1 + jω (RCT + Zw (jω)) CDL

where the parameters are as indicated in Figure 1(a). Likewise, the qualitative
impedance plot shown in Figure 1(b) has four branches associated with four
specific electrochemical processes. In the first branch, denoted in this chapter
as the ‘RL Arc’, the effect of the inductive behavior at high frequencies (ω > ωk3 )
as well as the ohmic resistance (RΩ ) can be seen. The second branch (ωk2 < ω <
ωk3 ), consists of a semi-circle, denoted in this chapter as the ‘SEI Arc’, related
to the solid electrolyte interface. The effect of the double-layer capacitance
and charge transfer resistance at the electrodes represents the second semi-
circle (denoted as ‘CT Arc’) in the third branch (ωk1 < ω < ωk2 ). Finally,
the constant slope (denoted as ‘Diffusion Arc’) in the impedance plot in the last
branch, represents the diffusion processes in the active material of the electrodes;
it has a significant effect at very low frequencies (ω0 < ω < ωk1 ) only.

3 Computing Frequency Response Using DFT


The frequency response of the system, described in Section 2 can be computed
using the discrete Fourier transform (DFT) technique. In order to compute the
Nyquist plot, the voltage and current signals, measured in the time domain, are
converted to the frequency domain through the DFT technique. Let us assume

3
(a) Adaptive Randles equivalent circuit model (AR-ECM)

(b) Impedance Spectrum (Nyquist Plot)

Figure 1: AR-ECM and related impedance spectrum

4
that zv (k) and zc (k) are the measured voltage and current, respectively, from
the battery over a certain time window L, i.e.,

zv (k) = v(k) + nv (k), k = 1, 2, . . . , L


(3)
zc (k) = i(k) + nc (k), k = 1, 2, . . . , L

where k indicates time, v(k) is the true voltage, and i(k) is the true current.
The voltage and current measurement noise nv (k) and nc (k) are assumed to be
zero-mean i.i.d. with standard deviation σv and σc , respectively. The Fourier
transform of the voltage and current measurements in (3) are defined as
L
X −i2πkω
Zc (ω) = FFT(zc (k)) = zc (k)e L = I(ω) + Nc (ω) (4)
k=1
L
X −i2πkω
Zv (ω) = FFT(zv (k)) = zv (k)e L = V (ω) + Nv (ω) (5)
k=1

where Zv (ω) and Zc (ω) are voltage and current measurement in the frequency
domain. Here, V (ω) and I(ω) indicate the Fourier transforms of the noiseless
voltage and current, respectively. The Fourier transforms of the voltage and
current measurement noises are given by Nv (ω) and Nc (ω), respectively. The
impedance at frequency ω is now written as
 
Zv (ω) V (ω) + Nv (ω) 1
Z(ω) = = = (V (ω) + Nv (ω))
Zc (ω) I(ω) + Nc (ω) I(ω) + Nc (ω)
 
1 Nc (ω)
= (V (ω) + Nv (ω)) − (6)
I(ω) I(ω)2

where the following Taylor series approximation was used


 
1 1 Nc (ω)
≈ − (7)
I(ω) + Nc (ω) I(ω) I(ω)2

Let us re-write the impedance in (6) in the following format

V (ω)
Z(ω) = + Nz (ω) (8)
I(ω)

where it can be shown that the noise Nz (ω) is zero-mean. Let us denote the
real and imaginary parts of the frequency response at ωk as

zr (k) = zr (ωk ) = Re(Z(ωk ))


zi (k) = zi (ωk ) = Im(Z(ωk )) (9)

We will make use of this notation to describe the ECM parameter estimation
approaches in the subsequent sections.

5
4 ECM Parameter Estimation Problem
The ECM parameter estimation in the frequency domain can be formally stated
as follows: given the frequency response of the system Z(ω) at the frequencies
ω1 , ω2 , . . . , ωL , estimate the ECM parameters.

Θ̂ = arg min Z(ωi ) − Ẑ(ωi ) , i = 1, . . . , L (10)


Θ

where k·k denotes two-norm,

Θ = {RΩ , L, RSEC , CSEC , RCT , CDL , ZW } (11)

An approximate solution to the above optimization problem can be obtained


using non-linear least squares approach.

5 Approximate Estimation of ECM Parameters


Consider the impedance in (2) at very high frequencies ω > ω3 . The capacitive
reactance approaches to zero at very high frequencies and only the inductive
and ohmic resistance remain dominant. Hence, one can write

Z(ω) ≈ jωL + RΩ ω > ω k3 (12)

From this, the estimates of L and RΩ can be obtained as

R̂Ω = min (zr (k)) k > k3 (13)


|min (−zi (k))|
L̂ = k > k3 (14)
ω
At very low frequencies the Warburg impedance (1) becomes dominant. The
resistive part of the Warburg impedance is written as
σ
Rw = √ (15)
ω

Consequently, the real part of Z(ω), when ω < ωk1 , can be written as

zr (k) ≈ RΩ + RSEI + RCT + Rw k < k1 (16)

From (16), σ can be calculated as the slope of zr (k) versus √1ω such that k < k1 .
k
To compute σ, two low frequencies are chosen and the corresponding resistance
value is taken from the impedance plot. Let us select these two frequencies as
follows

ωa = ω0 (17)
ωb = ω s.t. ω0 < ω < ωk1 (18)

6
Then, the Warburg coefficient can be written as follows:

( ωa ωb )(zr (a) − zr (b))
σ̂ = √ √ (19)
ωb − ωa

Now, let us consider the two arcs (SEI Arc and CT Arc) in the Nyquist
plot to determine the value of RSEI , CSEI , RCT and CDL . First, consider the
CT Arc which occurs in lower frequencies, i.e., ωk1 < ω < ωk2 . The (Faradaic)
impedance due to RCT and CDL , in this region is
1
ZF (ω) = 1
RCT +Zw (jω) + jωCDL
1
RCT +Zw (jω) − jωCDL
= 1 ωk1 < ω < ωk2 (20)
( RCT +Z w (jω)
)2 + ω 2 C 2 DL

By ignoring the effect of Warburg impedance, the impedance corresponding to


the CT Arc can be written as
1
RCT− jωCDL
ZCT (ω) ≈  2 ωk1 < ω < ωk2 (21)
1
RCT + ω 2 C 2 DL

At the peak of CT Arc, one can observe (see Figure 1(b))


1
= RCT at ω = ωCT,peak (22)
ωCDL
1
|Im(ZCT (ω))| = RCT at ω = ωCT,peak (23)
2
and the following two estimates can be obtained:

R̂CT = 2|Im(ZCT (ω))| at ω = 2πfCT,peak (24)


1
ĈDL = at ω = 2πfCT,peak (25)
ω R̂CT
where fCT,peak denotes the frequency corresponding to the peak of the CT Arc.
Let us now consider the SEI Arc in the range of ωk2 < ω < ωk3 . In this
region, the impedance is given by
1
1 RSEI − jωCSEI
ZSEI (ω) = 1 = 1 ωk2 < ω < ωk3 (26)
RSEI + jωCSEI R2 SEI + ω 2 C 2 SEI

At the peak of SEI Arc, we have (see Figure 1(b))


1
= RSEI at ω = ωSEI,peak (27)
ωCSEI

7
Based on the above observation, we have
1
|Im(ZSEI (ω))| = RSEI at ω = ωSEI,peak (28)
2
and the following two estimates can be obtained:

R̂SEI = 2|Im(ZSEI (ω))| at ω = 2πfSEI,peak (29)


1
ĈSEI = at ω = 2πfSEI,peak (30)
ω R̂SEI
where fSEI,peak denotes the frequency corresponding to the peak of the SEI Arc.
Despite its simplicity and low processing time, the algorithm reviewed in
this section has some drawbacks due to the effect of approximation and due to
the effect of measurement noise. These effects are described next.

5.1 Effect of Approximation


For the approximation parameter estimation presented in this section, the peak
values of SEI Arc and CT Arc are found and the value of RC elements are
calculated using (23) to (30). However, it should be noted that, for these cal-
culations, the effect of estimated Warburg impedance on CT Arc is neglected.
Figure 2 illustrates how this approximation will affect the parameter estimation.
For the case shown in Figure 2 the frequency at which the CT Arc reaches its
maximum value is changed more than 50 percent (from 0.59 to 0.91); that will
affect the values of RCT and CDL according to (24).

5.2 Effect of Measurement Noise


The measurement noise also plays a major role in the accuracy of estimated
ECM parameters. In order to analyze the performance, let us first define the
SNR as
 
PSignal
SNR = 10 log (31)
PNoise

where the unit of SNR is decibels (dB). Figure 3 shows the Nyquist plots at four
different SNR values (0, 5, 15, 30 dB). The algorithm summarized in Section
5 is applied to estimate the AR-ECM parameters in each case. The estimated
parameters are then used to generate the Nyquist plot. Ideally, both plots
should coincide. It can be noticed that with increasing noise, the discrepancies
become prominent.

6 Improved Approach for Parameter Estimation


Let us denote the frequencies at different branches of the Nyquist plot as follows:
• Warburg: ω0 , ω1 , . . . , ωk1

8
0.16
With Warburg Impedance
0.14 Without Warburg Impedance

0.12

0.1

0.08

0.06

0.04

0.02

-0.02
0.55 0.6 0.65 0.7 0.75 0.8 0.85

Figure 2: Effect of approximation in performance

• CT: ωk1 +1 , ωk1 +2 , . . . , ωk2


• SEI: ωk2 +1 , ωk2 +2 , . . . , ωk3

• RL: ωk3 +1 , ωk3 +2 , . . . , ωk4


In total, there are k4 frequency pairs at which the impedance measurements
were computed. Let us denote the real and imaginary parts of the impedance
measurements at each of the above frequencies as follows:

• Warburg:

[zr (0), zi (0)], [zr (1), zi (1)], . . . , [zr (k1 ), zi (k1 )] (32)

• CT:

[zr (k1 + 1), zi (k1 + 1)], [zr (k1 + 2), zi (k1 + 2)], . . . , [zr (k2 ), zi (k2 )] (33)

• SEI:

[zr (k2 + 1), zi (k2 + 1)], [zr (k2 + 2), zi (k2 + 2)], . . . , [zr (k3 ), zi (k3 )] (34)

9
0.15 0.15
Approximate Approximate
Measurements 0.1 Measurements
0.1

0.05
0.05
0

0
-0.05
0.5 0.6 0.7 0.8 0.5 0.6 0.7 0.8 0.9

0.15 0.15
Approximate Approximate
Measurements Measurements
0.1 0.1

0.05 0.05

0 0
0.5 0.6 0.7 0.8 0.9 0.5 0.6 0.7 0.8

Figure 3: Effect of measurement noise on performance

• RL:

[zr (k3 + 1), zi (k3 + 1)], [zr (k3 + 2), zi (k3 + 2)], . . . , [zr (k4 ), zi (k4 )] (35)

We will use the above notations to discuss an improved version of the approxi-
mate parameter estimation approach discussed in Section 5.

6.1 Estimation of Warburg Coefficient


Several frequencies in the Diffusion Arc can be selected to write
 
1 1
zr (0) − zr (k1 ) = σ √ − √
ω0 ωk1
 
1 1
zr (1) − zr (k1 − 1) = σ √ −√
ωk1 ωk1 −1 (36)
..
.
 
1 1
zr (n) − zr (k1 − n) = σ √ −√
ωn ωk1 −n

10
where n < k1 /2. The observations in (36) were selected in such a way that the
quantity zr (i) − zr (j) could be as high as possible — this strategy is designed
to reduce the effect of noise in the observations. The observations (36) can be
written in matrix form as

z̃ = bσ (37)

where
   
  √1 − √
1
zr (0) − zr (k1 )  ω0  ωk1
 zr (1) − zr (k1 − 1)   √1 1
− √

 ωk1 ωk1 −1 
z̃ =  , b = (38)
 
..  
..

.
 
.
   
zr (n) − zr (k1 − n)
 
1 1
ωn − ωk −n
√ √
1

The least square estimate of σ is



ˆ bT z̃
σ̂ = T (39)
(b b)

6.2 Estimation of CT Components


Let us denote an impedance measurement in the CT Arc as

zr , zr (k) s.t. k1 < k ≤ k2


(40)
zi , zi (k) s.t. k1 < k ≤ k2

The measurements in (40) will satisfy the following circular equation

zr2 + zi2 + azr + b = 0


 a 2 a2
zr + + zi2 = −b (41)
2 4
where it was assumed that the center of the circle lies on the real axis (see Figure
1). The center of the circle (41) is denoted as (xCT , 0) where
a
xCT = − (42)
2
and the radius of the circle (41) is
r
a2
rCT = −b (43)
4
Based on (41), one can notice that the argument of the square root is always
positive. It is now easy to see that the estimate of the resistance RCT is
r
ˆ â2
R̂CT = 2 − b̂ (44)
4

11
where â and b̂ are estimates of a and b, respectively.
In order to estimate a and b, the pairs of impedance measurements shown
in (33) can be substituted in (41) to get the following sets of equations
−(zr (k1 + 1)2 + zi (k1 + 1)2 ) = azr (k1 + 1) + b
−(zr (k1 + 2)2 + zi (k1 + 2)2 ) = azr (k1 + 2) + b
.. (45)
.
−(zr (k2 )2 + zi (k2 )2 ) = azr (k2 ) + b
the above can be written in matrix form as
z = BxCT (46)
where
−(zr (k1 + 1)2 + zi (k1 + 1)2 )
   
zr (k1 + 1) 1
−(zr (k1 + 2)2 + zi (k1 + 2)2 ) zr (k1 + 2) 1  
a
z= , B =   , xCT = (47)
   
.. .. b
 .   . 
−(zr (k2 )2 + zi (k2 )2 ) zr (k2 ) 1
The least-square estimate of xCT is
−1 T
x̂CT = BT B B z (48)
and the estimates of a and b are
â = x̂CT (1), b̂ = x̂CT (2) (49)
which will be substituted in (44) to estimate RCT . Then, for estimating CDL ,
the inverse of (20) is calculated as:
1 1 1
= + jωCDL = + jωCDL
ZCT (ω) RCT + Zw (jω) RCT + σω − j √σω

RCT + √σω + j √σω


= 2 + jωCDL (50)
2
RCT + √σω + σω

Hence,
√σ
 
1 ω
Im = 2 + ωCDL (51)
ZCT (ω) √σ σ2
RCT + ω
+ ω

and by substituting the estimate of RCT we get (for ω = ωk )


 
√σ
 
1  1 ωk
C̄DL (k) = Im −

ωk ZCT (ωk ) 2 
ˆ σ2
R̂CT + √σωk + ωk

k = k1 + 1, k1 + 2, . . . , k2 (52)

12
Finally, all estimates of C̄DL (k) are averaged to obtain
k2
ˆ 1 X
ĈDL = C̄DL (k) (53)
k2 − k1
k=k1 +1

6.3 Estimation of SEI Components


Let us denote an impedance measurement in the SEI Arc as

yr , zr (k) s.t. k2 < k ≤ k3


(54)
yi , zi (k) s.t. k2 < k ≤ k3

The measurements in (54) will satisfy the following circular equation

yr2 + yi2 + cyr + d = 0 (55)

where it was assumed that the center of the circle lies on the real axis (see Figure
1). The center of the circle (41) can be denoted as (xSEI , 0) where
c
xSEI = − (56)
2
and the radius of the circle (41) is
r
c2
rSEI = −d (57)
4
By the same reasoning in Subsection 6.2, the argument of the square root can
be shown to be always positive. It now is easy to see that the estimate of the
resistance RCT is
r
ˆ ĉ2
R̂SEI = 2 − dˆ (58)
4

where ĉ and dˆ are estimates of c and d, respectively. In order to estimate c and


d, the pairs of impedance measurements shown in (34) can be substituted in
(55) to get the following sets of equations

−(yr (k2 + 1)2 + yi (k2 + 1)2 ) = cyr (k2 + 1) + d


−(yr (k2 + 2)2 + yi (k2 + 2)2 ) = cyr (k2 + 2) + d
.. (59)
.
−(yr (k3 )2 + yi (k3 )2 ) = cyr (k3 ) + d

The above can be written in matrix form as

y = AxSEI (60)

13
where
−(yr (k2 + 1)2 + yi (k2 + 1)2 )
   
yr (k2 + 1) 1
−(yr (k2 + 2)2 + yi (k2 + 2)2 ) yr (k2 + 2) 1  
c
y= , A =   , xSEI =
   
.. .. d
 .   . 
−(yr (k3 )2 + yi (k3 )2 ) yr (k3 ) 1
(61)

The least-square estimate of xSEI is


−1
x̂SEI = AT A AT y (62)

and the estimates of c and d are

ĉ = x̂SEI (1), dˆ = x̂SEI (2) (63)

which will be substituted in (58) to estimate RSEI . Then, CSEI is simply driven
from the inverse of (26):
1 1
= + jωCSEI (64)
ZSEI (ω) RSEI

From the above we get (for ω = ωk )


   
1 1
C̄SEI (ωk ) = Im (65)
ωk ZSEI (ωk )

Finally, all possible estimates of C̄SEI are averaged to obtain


k3
ˆ 1 X
ĈSEI = C̄SEI (k) (66)
k3 − k2
k=k2 +1

6.4 Estimation of Resistance and Inductance


Finally, for a better estimation of the ohmic impedance and the inductance, the
values are averaged over a range of high frequencies (ωk3 < ω ≤ ωk4 ).

k4
ˆ 1 X
R̂Ω = zr (k) (67)
k4 − k3
k=k3 +1
k4
ˆ 1 X zi (k)
L̂ = (68)
(k4 − k3 ) ωk
k=k3 +1

in which, zr (k) = Re (Z(ωk )) and zi (k) = Im (Z(ωk )).

14
6.5 Feature Point Extraction
It should be stressed that the least squares based (approximate) parameter esti-
mation algorithm presented in this section needs to know the critical frequency
values ωk0 , . . . , ωk4 . It is easy to know the values of ωk0 and ωk4 as these are
the lowest and highest frequencies, respectively, in the Nyquist spectrum. An
approach to estimate ωk1 , ωk2 , and ωk3 is presented in [1]. In this approach,
a straight line is fitted to the Diffusion Arc with progressively increasing data
starting from the lower frequency. As the Diffusion Arc turns into CT Arc,
the correlation of fitting starts to drop. The critical point k1 is detected by
observing the drop in k1 relative to a predefined threshold. Similarly, a circular
curve is fitted to the data starting from the (detected) critical point k1 and the
correlation coefficient is monitored to detect the critical point k2 based on a
predefined threshold. The critical point k3 can be detected based on the fact
that the imaginary part of the Nyquist curve changes its sign at k3 .

7 Demonstration
In this section, computational demonstration of the proposed ECM parame-
ter identification techniques are presented using simulated and real-world data.
[selected codes will be added later.]

7.1 Demonstration Using Simulated Data


In order to obtain the impedance spectrum, simulation is done for a 1000 mAh
lithium-ion battery, parameters of which are taken from a real battery through
experiment. For simulation, a 200 mA (C/5 Rate) DC charging current and an
AC perturbation current with a peak of 70 mA are used. A series of 901 single-
sine waves was used with different frequencies in the range of 0.01 Hz to 10 kHz
(k = 901 frequencies) for better resolution and more precise demonstration of
an industrial Li-ion battery. The battery charging current, IB (t), is represented
as

 Idc + Im sin(ω1 t), t < t1

 Idc + Im sin(ω2 t), t1 ≤ t < t2

IB (t) = .. (69)


 .
Idc + Im sin(ωk t), tk−1 ≤ t < tk

where Idc is the DC current, Im is the peak of perturbation current and each tk
is selected such that tk − tk−1 = f1k i.e., each frequency is made to have one full
cycle of data.
For measuring the noise effects it is assumed that the measured current and
voltage are corrupted with Gaussian noise of the same noise variances, i.e.,

σv = σi = Im 10(− )
SNR
20 (70)

15
where SNR varies from 0 to 50 dB (0, 10, 20, 30, 40, 50 dB).
The Nyquist plot is derived for each case (each level of noise) and algorithms
explained in sections 4, 5, and 6 are applied to estimate ECM parameters.
The performance of each algorithm is quantified in terms of the normalized
percentage mean square error, simply referred hereafter as Error%. For example,
the Error (%) of estimating the Ohmic resistance by the proposed approach is
defined as
ˆ
RΩ − R̂Ω
Error(%) = 100 (71)
RΩ
Each reported Error measure is averaged over 100 Monte-Carlo runs.
Figures 4-7 presents the percentage error for each of the approaches discussed
in this chapter. Figure 4 compares the performance in estimating the Warburg
coefficient (σ). An explanation of the performance loss by the approximate
approach (Section 5) could be that it used only two points to find the slope of
a line to estimate σ. On the other hand, the improved approach (Section 6)
used many pairs of points and resulted in better estimation error. The reason
for the failure of the non-linear LS approach could be attributable to the severe
non-linearity in the model when it comes to estimating the Warburg coefficient.

Figures 5(a) and 5(b) show the performance comparison of the three ap-
proaches presented in this chapter for RCT and CDL estimation. And figures
6(a) and 6(b) show similar comparison for RSEI and CSEI estimation. These two
figures exhibit the performance trade-off of the different approaches presented
in this chapter and highlight the need to develop robust approaches to estimate
ECM parameters.
Figure 7(a) shows the estimation errors corresponding to the Ohmic resis-
tance, RΩ . All the algorithms have an excellent performance (error rate lower
than 1%) in estimating RΩ . This is due to the linear relationship of RΩ to the
measurements in (2). Figure 7(b) shows the comparison of different estimators
in the estimation of stray inductance L.
For a comprehensive analysis of the performance of each three approaches
presented in this chapter, the extracted parameters from each method are used
to generate the Nyquist plot. Figure 8(a) compares the performance of the
previous approach and the proposed algorithm in estimating the Nyquist plot
at SNR = 30 dB. Figure 8(b) shows the performance comparison of all three
approaches in a severe but practical case (low SNR). In the presence of a high
level of noise, the approximate LS approach is observed to outperform both the
non-linear LS approach and the approximate approach, as shown in Figure 8(b).

7.2 Demonstration Using Real Data


In this section, ECM parameter extraction is demonstrate using a real world
experiment. For this, an Arbin battery cycler with Gamry interface 5000P

16
70
Approximate
Approximate LS
60 Non-linear LS

50

40

30

20

10

0
0 5 10 15 20 25 30 35 40 45 50

Figure 4: Parameter estimation error the Diffusion Arc.

40 50
Approximate Approximate
Approximate LS 45 Approximate LS
35
Non-linear LS Non-linear LS
40
30
35
25
30

20 25

20
15
15
10
10
5
5

0 0
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50

(a) Charge transfer resistance (b) Double layer capacitance

Figure 5: Parameter estimation of the CT Arc.

EIS device were used. Fig. 9 shows these devices in an experimental setup —
the computer screen shows the interface of the Arbin software that allows to
collect Nyquist data from the Gamry device. A relatively new cylindrical Li-ion

17
400 2000
Approximate Approximate
Approximate LS 1800 Approximate LS
350
Non-linear LS Non-linear LS
1600
300
1400
250
1200

200 1000

150 800

600
100
400
50
200

0 0
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50

(a) Solid electrolyte interface resistance (b) Solid electrolyte interface capacitance

Figure 6: Parameter estimation of the SEI Arc.

1 25
Approximate Approximate
0.9 Approximate LS Approximate LS
Non-linear LS Non-linear LS
0.8 20

0.7

0.6 15

0.5

0.4 10

0.3

0.2 5

0.1

0 0
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50

(a) Ohmic resistance (b) Stray inductance

Figure 7: Parameter estimation of the RL Arc.

battery (LG INR18650 MJ1) is used in the experiment. The specifications of


this battery are shown in Table 1. The state of charge of the battery is nearly
empty when the experiment was performed.
First, the Gamry EIS device was programmed to use zero DC current with
a 50 mA sinusoid superimposed on it. In a second experiment, right after the
first experiment, the experiment was repeated with 200 mA charging current
and 50 mA sinusoid superimposed on it. The output of the Gamry EIS device is
the real and imaginary values of the measured impedance, as shown in Fig. 10.
The parameter estimation algorithms proposed in sections 4 and 6 were applied
on the impedance measurement to estimate the battery’s AR-ECM parameters.
Table 2 shows the estimated values of the AR-ECM parameters based on the
nonlinear LS approach. It can be seen in Table 2 that the estimated parameters

18
0.16 0.18

0.14 0.16

0.12 0.14

0.12
0.1
0.1
0.08
0.08
0.06
0.06
0.04
0.04
0.02
Approximate 0.02 Approximate
Approximate LS Approximate LS
0 Non-Linear LS 0 Non-Linear LS
Measurements Measurements
-0.02 -0.02
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95

(a) High SNR (b) Low SNR

Figure 8: Performance comparison where the estimated parameters were used


to reconstruct Nyquist plot.

Figure 9: Experimental setup

are slightly different when there is a DC current compared to the zero-mean


current experiment. It is assumed in this chapter that the DC values of both
voltage and current are zero. However, in reality, the non-zero DC current is
more practical (that can be superimposed during battery charging) and this
results in performance loss. Detailed analysis of such performance losses are
not discussed in this chapter. Instead, the experiment presented in this section
is intended to make one aware of the discrepancies of estimated parameters in
the presence of non-zero DC current. Table 3 summarizes the results of ECM
parameter estimation using the approximation LS approach. It can be noticed
that the estimated Warburg coefficient is significantly different in Tables 2 and
Table 3. In the absence of true values, according to the simulation analysis pre-
sented in Section 7.1, Table 3 con be considered to provide the most approximate

19
ECM parameters of the battery.

Table 1: Battery specifications

Specification Value (unit)


Nominal capacity 3500 mAh
Max. current 10 A
Nominal voltage(Vnom ) 3.7 V
Height 65 mm
Diameter 18 mm
Weight 46.5 g

0.02

0.01

-0.01 With non-zero DC current


With zero DC current
-0.02

-0.03

-0.04

-0.05

-0.06

-0.07
0.405 0.41 0.415 0.42 0.425 0.43 0.435

Figure 10: Nyquist plot for real battery. The nyquist plot in blue depicts the
impedance response of the battery with 0.2 A DC current. Impedance response
of the battery with 0 DC Current is shown in green.

8 Summary
This chapter introduced frequency domain approaches to battery equivalent cir-
cuit model parameter estimation using EIS. In EIS, an excitation signal (either

20
Table 2: Estimated parametes (Nonlinear LS)

Paramter Value (0 DC current) Value (-0.2 DC current)


RΩ 412.5 mΩ 411.3 mΩ
RCT 2.6848 mΩ 3.22 mΩ
CDL 1.000 F 1.000 F
L 1.1588 × 10−06 H 1.1534 × 10−06 H
σ 0.0584 0.0656

Table 3: Estimated parametes (Approximate LS)

Paramter Value (0 DC current) Value (-0.2 Dc current)


RΩ 410.7 mΩ 409.7 mΩ
RCT 3.1 mΩ 3.3 mΩ
CDL 1.0278 F 0.8740 F
L 1.1142 × 10−06 H 1.1107 × 10−06 H
σ 0.0020 0.0041

voltage or current) is applied to the battery and its response (current or voltage)
is measured. This procedure is repeated and the amplitude and phase of the
frequency response is computed at various (fixed) frequencies spanning very low
frequency in fractions of Hz and very high frequency in several MHz. Based
on the obtained responses at wide ranging frequencies the ECM parameters
can be estimated. This chapter outlines three different approaches to estimate
the ECM parameters based on the frequency response. The first approach is
based on the non-linear least squares estimation which requires significant com-
putational resources. Also, the non-linear least squares approach is shown to
be susceptible local convergence even at high SNR regions. It is also shown
that distinct portions of the Nyquist plot can be extracted to estimate ECM
parameters.
The EIS approach to battery analysis is a time consuming process. Es-
pecially, measurements at low frequencies incurs significant delay. Significant
recent research work is focused on reducing the experimental time of the EIS
approach for battery analysis.

References
[1] Marzieh Abaspour, “Performance Analysis and Improvement of Electro-
chemical Impedance Spectroscopy for Online Estimation of Battery Param-
eters,” Dissertation, University of Windsor (Canada), 2021.

21
[2] Mark E. Orazem and Bernard Tribollet, “Electrochemical Impedance Spec-
troscopy,” John Wiley & Sons, Inc., New Jersey, 2008.

22

You might also like